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 [2] The structural style of mountain belts greatly affects foreland basin development, mountain building, exhumation and sediment recycling. Although each orogenic system is unique, all mountain belts are subject to the same fundamental tectonic processes of continental contraction and lateral transport of upper crustal material. On the most fundamental scale, the transported rocks of a fold‐and‐thrust belt act like a wedge of sand or snow in front of a moving bulldozer, incorporating new sediments and rocks at the frontal tip across the basal thrust, and sometimes deforming internally within the wedge to maintain a critical taper [Bally et al., 1966; Dahlstrom, 1970]. [3] Thin‐skinned systems are characterized by low‐angle thrust faults, multiple shallow décollements, ramp flat 1 Department of Geosciences, University of Arizona, Tucson, Arizona, USA. 2 Department of Geology and Geophysics, University of Wyoming, Laramie, Wyoming, USA. 3 Department of Geology, Kansas University, Lawrence, Kansas, USA. geometries, duplex structures and a systematic progression of deformation that generally propagates toward the foreland through a thick package of sedimentary rocks with relatively low mechanical strength [e.g., Bally et al., 1966; Davis et al., 1983]. Conversely, thick‐skinned belts are characterized by steeper faults in crystalline basement, deeper décollements, and often experience more erratic, out‐of‐sequence development [e.g., Jordan, 1995]. Because of the relatively high mechanical strength of basement rocks, faults in thick‐ skinned systems often nucleate along preexisting heterogeneities, such as fabrics, ancient shear zones, pervasive fracture patterns, or inherited fault systems [Allmendinger et al., 1983; Schmidt et al., 1995]. [4] The Eastern Cordillera of the central Andes involves substantial amounts of mechanical ‘basement’, in this case previously deformed low‐grade metasedimentary rocks [e.g., McQuarrie et al., 2005]. A key question addressed in this paper is whether or not the timing of deformation in the Eastern Cordillera (EC) of NW Argentina is similar to what observed along strike to the north in Bolivia, where the EC is 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 information on the magnitude and timing of cooling and exhumation 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 Figure 1. Shaded regional topography of the central Andes of northern Argentina and southern Bolivia, showing tectonomorphic zones (labeled), thrust faults (barbed lines), and volcanoes (circles) [modified after Sobel et al., 2003]. Black box shows area covered in Figure 2. [5] The Eastern Cordillera (EC) in Argentina and Bolivia is a large retroarc fold‐and‐thrust belt that has developed during the Cenozoic along the eastern margin of the central Andean Plateau from ∼13 to 26°S (Figure 1). The basement of the EC in NW Argentina is mainly characterized by high‐ angle reverse faults cutting into Precambrian‐Cambrian crystalline and metamorphic rocks, and locally Cretaceous rift sedimentary rocks, with a décollement at >30 km depth [Ramos, 2002]; whereas the basement in Bolivia is characterized by thrust faults cutting into Ordovician and Silurian metasedimentary rocks [McQuarrie, 2002]. This along‐ strike variability makes the EC a prime laboratory to examine the factors that control the sequence in which faults activate within the orogenic system, the rate of exhumation associated to deformation on those faults, and the style of foreland basin fragmentation. [6] This study combines detailed mapping of the Calchaquí Valley (Figure 1) with apatite (U‐Th)/He thermochronology and zircon U‐Pb geochronology of sandstone 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 orogenic 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 largest plateau in a retroarc setting. The Puna‐Altiplano plateau is bounded to the west by the magmatic arc and monoclinal flexure of the Western Cordillera and to the east by the varied contractional systems of the Eastern Cordillera (EC), Santa Barbara System, sub‐Andes, and Sierras Pampeanas (Figure 1). [9] Foreland basin deformation in the central Andes exhibits a variety of structural styles that can be broadly correlated along strike to the angle of subduction of the Nazca plate under the South American plate [Isacks, 1988; Jordan and Allmendinger, 1986]. Thick‐skinned tectonics deformation in the Sierras Pampeanas is associated with the 10° “flat slab” subduction under central Argentina, whereas the thin‐skinned fold‐and‐thrust belts of the EC and sub‐ Andes are associated with the 30° “normal” subduction under Bolivia [Cahill and Isacks, 1992]. However, this relationship breaks down between 25° and 28°S where the subducting plate is represented by a broad transitional flexure at an angle between 10° and 30° [Ramos, 2002]. In this region, the angle of the subducting slab appears to have less influence on the structural style of foreland deformation than mechanical weaknesses in the crust related to preexisting basement anisotropy. This study focuses on the southern EC in this transitional interval, where structural style is controlled by a complex interplay between inherited structures, inversion of Cretaceous rift structures and foreland basin 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 Salfity, 1995]. 2.2. Eastern Cordillera [10] The Eastern Cordillera bounds the eastern margin of the plateau for over 1,500 km along strike, and it varies greatly in topographic morphology, structural style, and magnitude of shortening over that distance (Figure 1) with average total shortening up to 270 km in Bolivia and 70 km in NW Argentina [Kley et al., 1999; Mon and Drozdzewski, 1999; Elger et al., 2005; McQuarrie et al., 2005, 2008]. Both the thick‐skinned and the thin‐skinned portions of the EC are bivergent fold‐and‐thrust belts, with the west half of the cordillera verging westward and the east half of the 3 of 30 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 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION 4 of 30 TC3003 TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION cordillera verging eastward. There is also a difference in timing, with deformation in the Bolivian EC initiating earlier than the Argentinean EC [McQuarrie et al., 2005]. [11] From ∼15 to 23°S, the Bolivian EC and sub‐Andes are connected today as part of a single, retroarc fold‐and‐ thrust belt developed along the eastern margin of the Central Andean Plateau (Figure 1), although the structures observed today are a combination of pre‐Andean and Cenozoic deformation. The fold‐and‐thrust belt evolved in places to form separate east and west verging systems; deformation in the EC and sub‐Andes has overall propagated eastward into the foreland during the Cenozoic [Echavarria et al., 2003; Uba et al., 2009; McQuarrie et al., 2005, 2008]. Deformation in the Argentinean EC from 23 to 26°S is markedly different from Bolivia. It exhibits a complex structural style dominated by high‐angle reverse faults involving intrusive and metamorphic basement and Cretaceous sedimentary rocks [Grier et al., 1991; Mon and Salfity, 1995]. [12] There is some evidence from sedimentology and magnetostratigraphy to suggest that the range uplift in this part of the EC may have swept generally eastward as it did in the Bolivian EC [Carrera and Muñoz, 2008; Reynolds et al., 2000], but these studies are based on indirect proxies, and the sequence of deformation and associated exhumation remains unresolved in this part of the EC. This study focuses on the Angastaco area in the southernmost EC (Figures 1 and 2), where Cenozoic shortening has inverted older north‐south trending normal fault‐bounded rocks of the Cretaceous Salta Rift. In this region basement anisotropies and abrupt changes in sedimentary thickness of Cretaceous rift strata profoundly affect the style and geometry of recent contractional deformation (Figure 3) [Mon and Salfity, 1995; Carrera et al., 2006]. Seismic reflection studies and regional balanced cross sections suggest that the ranges of the southwestern EC have been uplifted by the domino rotation of semirigid basement blocks along spoon‐shaped faults, with net shortening of 14–25% [Grier et al., 1991; Cristallini et al., 1997; Mon and Salfity, 1995; Cristallini et al., 2004; Kley and Monaldi, 2002]. Major folds run parallel to fault trends in the southern EC, and synrift rocks are often thrust over postrift and Neogene foreland strata [Grier et al., 1991; Carrera et al., 2006]. These high‐angle reverse faults often have limited displacement, creating broad to overturned anticlines in their hanging walls and synclines in their footwalls (Figure 5) [Grier et al., 1991; Carrera et al., 2006]. Kinematic analysis of major and minor structures has revealed a Mio‐Pliocene interval of WNW‐ESE shortening that created the major structures exposed at the surface, followed by a minor secondary stage of Pliocene‐Quaternary WSW‐ESE shortening [Grier et al., 1991; Marrett et al., 1994]. Kley and Monaldi [2002] note that narrow folds with shallow detachments are rare within this part of the EC, but major structures are much more closely spaced than in the nearby thick‐skinned Sierras Pampeanas to the south [Ramos, 2002]. The Argentinean EC represents a type of “fold‐and‐thrust” belt that is transitional to the classic thin‐skinned and thick‐skinned models [e.g., Kley et al., 1999], and therefore likely contains a hybrid of the structural and kinematic features commonly found in these two deformation regimes. TC3003 2.3. Stratigraphy [13] There are three major stratigraphic components in the southern EC (Figure 2), including (1) the metamorphic Precambrian‐Cambrian Puncoviscana Formation, (2) rift fill sedimentary rocks of the Cretaceous Salta rift, and (3) a thick Cenozoic sequence of synorogenic Andean foreland basin strata (Figure 4) [Grier et al., 1991]. The Puncoviscana Formation represents a thick passive margin sequence that was deposited during the Neoproterozoic to Cambrian, shortened and metamorphosed during the Pampean orogeny, and later intruded by multiple granitic bodies [Jezek and Miller, 1985; Willner and Miller, 1985; Grier et al., 1991]. The early Paleozoic metamorphism resulted in crustal imbrication and in the development of east dipping shear belts and axial planar cleavages that permeate the formation today [Willner et al., 1987]. [14] The Cretaceous‐Eocene Salta group rests unconformably on the Puncoviscana Formation. The Salta Group is over 5 km thick in places and is divided into three subgroups: the Pirgua, Balbuena and Santa Barbara subgroups (Figure 4). The Pirgua subgroup is composed of over 3 km of synrift sandstones, conglomerates, siltstones, and volcanics of the La Yesera, La Curtiembres, and Los Blanquitos Formations [Marquillas et al., 2005]. The Balbuena subgroup is composed of 400–500 m of postrift sandstones, limestones, and pelites. The Santa Barbara subgroup is composed of early foreland basin sandstones, siltstones, limestones, and shales [Marquillas et al., 2005]. Early foreland basin strata (Paleogene) were deposited on a mixed substratum within a continuous regional basin that was later divided into smaller isolated basins by intrabasin range uplift in the late Miocene [Jordan and Alonso, 1987; Grier et al., 1991; Starck and Vergani, 1996; Bosio et al., 2009]. These strata are represented in the southern EC by four formations of the Eocene‐Pliocene Payogastilla group [Grier and Dallmeyer, 1990]. In the study area, the Eocene‐ Oligocene Quebrada de los Colorados Formation is composed of approximately 350 m of coarse sandstone and conglomerate deposited in a braided fluvial to alluvial plain system in a distal foredeep depocenter [Starck and Vergani, 1996]. The Miocene Angastaco Formation is composed of ∼4.2 km of proximal fluvial sandstones and conglomerates that generally coarsen and thicken upward [Starck and Vergani, 1996]. The Miocene‐Pliocene Palo Pintado Formation is composed of ∼2 km of cross‐stratified medium grained sandstones and mudstones, topped by ∼450 m of mudrocks, channel sandstones, pebble conglomerates, and intercalated tuffs [Starck and Vergani, 1996]. The Pliocene San Felipe Formation is a ∼650 m thick succession of pebble to cobble conglomerates, medium‐grained sandstones, and sparse green mudrocks [Starck and Vergani, 1996]. The reported thickness of the Neogene formations should be considered minimums because unit thicknesses vary along strike, likely thicken in the subsurface, and are eroded and crosscut by the Calchaquí Fault to the east (Figure 2). 2.4. Chronology of Deformation in NW Argentina [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 Maastrichtian time [Grier et al., 1991], and although Neogene deformation has overprinted the intricate geometry of the original rifts [Kley and Monaldi, 2002], extensional structures 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 deformation 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 segmented by inherited structural highs and geographically complex [e.g., Carrapa et al., 2005; Carrapa and DeCelles, 2008; Hongn et al., 2007]. Original paleodrainages 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 documented that the regional foreland basin became successively 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 instantaneously, which is a similar pattern to the one proposed for the Altiplano in Bolivia [McQuarrie et al., 2005]. Alternatively, 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 sedimentation 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 manuscript, 2011] with approximate stratigraphic location of the analyzed samples. 6 of 30 TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION of a regional retroarc foreland basin that formed in response to subsidence and sediment influx from the rising Puna and Cumbres de Luracatao range (EC) directly to the west (Figure 2) [Coutand et al., 2006]. [18] Features typical of growth strata, tilted unconformities, and thermochronological data in the southern EC have suggested a west‐to‐east and north‐to‐south migration of deformation across the EC and Santa Barbara System starting from Miocene time [Acocella et al., 2007; Carrera and Muñoz, 2008; Deeken et al., 2006]. The northern tip of the Sierra de Quilmes range (Figure 2) (referred to as Cerro Negro by Coutand et al. [2006]) began exhuming by 13–10 Ma, resulting in the removal of Pirgua and early Payogastilla Group strata from the range and deposition into the Angastaco and Pucará areas (Figure 4) [Coutand et al., 2006]. From 14.5 Ma to 5.5 Ma, the EC has been the primary source of sediment for the Angastaco area, [Coutand et al., 2006]. Multiple apatite fission track data sets support the exhumation of the Cerro Runno and Cerro Durazno range (Figure 2) during the interval of 13–10 Ma [Coutand et al., 2006; Deeken et al., 2006]. Paleodrainage reorganization in the Angastaco‐La Viña areas suggest that the Angastaco basin was fully separated from the rest of the foreland and was in an intermontane position by ∼4 Ma (B. Carrapa et al., Cenozoic basin evolution in the Eastern Cordillera of northwestern Argentina (25°–26°S): Regional implications for Andean orogenic wedge development, submitted to Basin Research, 2011). [19] As the sedimentary system was fragmented into a series of isolated basins, each of those depocenters became progressively deformed. Synclinal folding of the Pucará area strata occurred during southward propagation of an adjacent anticline (Figure 3), and was determined to be active between 13.4 ± 0.4 Ma and 12.1 ± 0.1 Ma by U‐Pb geochronology on a welded tuff, lahar deposit, and lower Payogastilla strata [Grier and Dallmeyer, 1990; Marrett et al., 1994]. However, the timing of deformation and exhumation in the Angastaco area and Lerma Valley has so far been based on indirect proxies (i.e., provenance, undated 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 by ∼4 Ma [Bywater‐Reyes et al., 2010]. Growth strata relationships suggest that the Cumbres de Peñas Blancas range east of the Lerma Valley (Figure 2) was uplifted and exhumed during the past 5 Ma [Carrera and Muñoz, 2008]. After deformation swept across the southern EC during the Miocene and Pliocene, out‐of‐sequence deformation and minor strike‐slip movement occurred throughout the EC during the Pleistocene and Holocene [Carrera and Muñoz, 2008; Marrett et al., 1994]. [20] Further south the Sierras Pampeanas, including areas all along the southeastern margin of the Puna Plateau, experienced a transition from regional foreland to compartmentalized basin at ∼6 Ma [Mortimer et al., 2007; Carrapa et al., 2008a]. The most recent stratigraphic record in the region is represented by the flat lying Pleistocene lake deposits in the Calchaquí Valley near the town of Angastaco, which lie in angular unconformity atop the deformed Payogastilla group and document the end of deformation [Salfity et al., 2004]. TC3003 2.5. Study Area: Calchaquí Valley [21] The Calchaquí Valley is an elongate depression located in the southwestern EC near the eastern margin of the Puna Plateau (Figure 2). It is subdivided into the Angastaco and Tin‐Tin areas (or subbasins), and separates the Alemanía and Brealito subbasins of the Cretaceous Salta rift (Figure 2) [Deeken et al., 2006]. The valley’s linear shape is controlled in part by the long N‐S trending Calchaquí fault on its eastern margin (Figure 2). The Cerro Negro, Payogasta, and Tin‐Tin reverse faults in the northern part of the Calchaquí valley expose synrift and postrift rocks in their hanging walls, but there is no evidence to suggest that synrift sedimentary rocks exist under the southern part of Calchaquí valley, and in places foreland basin strata lie directly on crystalline basement (Figure 2) [Carrera and Muñoz, 2008; Mon and Salfity, 1995; G. Vergani and D. Starck, unpublished map, 1988]. Because of its excellent exposures and relative accessibility, the Angastaco area of the Calchaquí Valley is an ideal location to study the details of intrabasin deformation that can occur during and after fragmentation of the larger foreland basin system (Figure 2). The Angastaco area is part of a valley that continues ∼100 km north as the Calchaqui Valley and south ∼130 km into the Santa Maria Valley; it is bounded to the east by the inverted Calchaquí Fault, to the west by the Sierras Pampeanas style Sierra de Quilmes uplift, (Figure 2). The Angastaco area contains evidence of syntectonic sedimentation, intrabasin faulting and folding, and is characterized in the southern part by large drape folds and thick‐skinned basement uplifts [Carrera and Muñoz, 2008]. Many of these features are common in other parts of the southern EC, such as the Luracatao Valley, the Tin‐ Tin and Amblayo areas, and the Lerma Valley (Figure 2) [Vergani and Starck, 1988]. Detailed sampling of discrete structures for thermochronology in the Angastaco area can thus provide cooling ages, and by inference the timing of deformation and exhumation, in a variety of Cenozoic basins within the southern EC. 3. Deformation in the Calchaqui Valley [22] In order to have a reliable stratigraphic and structural frame for our sampling field mapping of geologic structures within the Angastaco area was conducted using 1:30,000 scale stereoscopic aerial photographs and form the basis of Figure 5. In particular, detailed mapping was conducted to identify possible growth in the Angastaco Formation, which records syndepositional deformation and thus the age of strata can be used to determine the timing of deformation within the basin. [23] Evidence of synorogenic sedimentation within the Angastaco area was observed in one location (Figure 6). To identify growth, we were looking for truncation, thinning or thickening of beds across strike. A 24° angular unconformity was found near the town of Angastaco in the lower fluvial Angastaco Formation (Figure 6), with a well‐ developed paleosol interval along the unconformity (∼1 m thick). Detrital U‐Pb geochronology on a sandstone sample immediately below the unconformity constrains the maximum 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 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 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 maximum 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 youngest 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 discordance) in the lower fluvial Angastaco Formation at the northern edge of the town of Angastaco. A paleosol horizon is well developed along the unconformity. Note person for scale. Formation yielded no other conclusive evidence for growth. [24] Eastward shallowing of bedding within the lower/ middle Angastaco Formation near the contact with the basement is suggestive of basin‐scale growth or can alternatively be attributed to the synclinal flexure of the basin (Figure 2). The contacts between the Angastaco/Palo Pintado and Palo Pintado/San Felipe Formations were observed to be conformable at every observed location within the study area. However, the intense folding and possible imbrication of the sedimentary section could easily obscure the shallow fanning of bedding that would provide more conclusive evidence of large‐scale growth. The angular unconformity in the lower Angastaco Formation represents the earliest evidence of syndepositional deformation in the area and is interpreted here to have formed by movement along the high‐angle Sierra de Quilmes and Cerro Negro faults to the west (Figure 2) right after ∼14 Ma. 4. Detrital U‐Pb Geochronology 4.1. Detrital U‐Pb Methods [25] Detrital zircon U‐Pb analysis has proven a reliable technique for determining provenance [Gray and Zeitler, 1997] and maximum depositional age of sedimentary rocks [DeCelles et al., 2007; Dickinson and Gehrels, 2009]. Four detrital sandstone samples from the Cenozoic basin fill preserved in the Angastaco and Pucará areas have been analyzed in this study to better constrain the maximum depositional age of the lower Angastaco and upper San Felipe Formations in the region where ashes were not present (Figure 2). In all samples, elongate euhedral zircon crystals were preferentially analyzed to increase the likelihood of documenting a young population of ashfall zircons, and therefore a statistically significant provenance analysis cannot be made from this data set. All analyses were conducted by laser ablation multicollector inductively coupled plasma mass spectrometry at the University of Arizona 4.2. Detrital U‐Pb Sampling [27] In order to discern the age of the uppermost San Felipe Formation, sample 04.14.08‐02 was collected in a 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 timing of recent movement on the fault as well. In order to discern the age of the angular unconformity in the Angastaco Formation near the town of Angastaco previously in this paper (Figure 6), two detrital sandstone samples were collected. Sample 05.07.08‐B.Unc was collected one meter below the base of the angular unconformity shown in Figure 6, and sample 05.07.09‐A.Unc was collected one meter above the same unconformity. In order to better define the age of the Angastaco Formation in the Pucará area, a boulder of andesite (sample 04.27.08‐AngB) was collected within an alluvial fan conglomeratic interval of the lower Angastaco Formation in the Pucará syncline (Figure 2). 4.3. Detrital U‐Pb Results [28] Of the four detrital zircon samples analyzed (Table A1), only samples 05.07.08‐B.Unc and 04.14.08‐02 had a young population of ages capable of constraining the maximum depositional age of the Angastaco and San Felipe Formations. For sample 04.14.08‐02, 40 detrital zircons were analyzed and 34 grains yielded concordant ages (Figure 7a). Zircon ages are distributed in four clusters, with major peaks around 1–9 Ma, 460–590 Ma, 1.0–1.1 Ga, and one old grain at 1.9 Ga. A young population of 14 grains yielded ages younger than 3 Ma. The Tuffzirc routine of Isoplot [Ludwig, 2008] was used to calculate a mean age of 2.3 ± 0.1 Ma from a coherent group of six detrital zircons, and the youngest single grain age is 1.8 ± 0.1 Ma (Figure 7a). We interpret this age to represent the formation and deposition of volcanogenic zircons during the deposition of the uppermost San Felipe Formation at ∼2.3 Ma. Because the formation here is crosscut by a major fault, this age also demonstrates that the Calchaquí fault has experienced some reverse offset since ∼2.3 Ma. The Paleozoic and Precambrian ages are typical of first cycle Paleozoic granites and detrital ages in the Puncoviscana Fm. [29] For sample 05.07.08‐A.Unc, 115 detrital zircons were analyzed and 102 grains yielded concordant ages (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 purposes, 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 Formation, 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) constrains 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 diffuses out of apatite at shallow crustal temperatures, the AHe system can be used as a precise low‐temperature thermochronometer sensitive to as little as 2–3 km of exhumation [Ehlers and Farley, 2003; Farley, 2002; Stockli et al., 2000; Zeitler et al., 1987] and >4 km in case of low geothermal gradients (<20°C). The AHe closure temperature is well constrained to be 75 ± 5°C in Durango apatite [Wolf et al., 1996; Ehlers and Farley, 2003], but can vary significantly depending on grain size, crystal morphology, cooling rate, and presence of radiation damage [Shuster et al., 2006; Ehlers and Farley, 2003; Wolf et al., 1996]. 5.2. AHe Sampling [33] Ten detrital sandstone and conglomerate rock samples were collected in the deepest exposed strata of five fault‐bounded segments of the southern EC (Figure 2 and Table 1). The fault‐bounded segments were sampled along a roughly east‐west oriented transect in the Pucará Valley, the Angastaco area of the Calchaquí Valley, the Sierra de los Colorados anticline, the Rio de las Conchas, and off transect in the Tin‐Tin area of the Calchaquí Valley (Figure 2). Two samples were collected from fault‐bounded segments of the EC, in multiple formations at each location where possible, and are therefore described in the following as sample pairs. All analyses were conducted using the ultra high vacuum noble gas extraction and purification line and ICP‐MS at the University of Kansas (U‐Th)/He laboratory, following analytical procedures of Stockli et al. [2000] and Biswas et al. [2007]. For complete AHe laboratory methods, consult the appendix to this paper. 5.2.1. Pucará Valley Sample Locations [34] The Pucará valley consists of a large N‐S trending syncline that contains Cretaceous through Neogene sedimentary rocks. Two samples were collected on the west limb of the syncline, in the deepest parts of the Cenozoic strata. Sample QCPucara1 was collected near the base of the Quebrada de los Colorados Formation, and Sample AngPuc06 was collected at the base of the Angastaco Formation just above the contact that separates these two formations (Figure 3). 5.2.2. Angastaco Area Sample Locations [35] The Angastaco area is a structurally complex segment of the Calchaquí Valley that contains Oligocene‐Pliocene synorogenic sedimentary rocks. Two samples were collected on the westernmost part of the valley, just south of the town of Angastaco, in the footwall of a large reverse fault that bounds the northeastern tip of the Sierra de Quilmes range. Sample QC1EB080 was collected in the upper Quebrada de los Colorados Formation, and sample Ang2EB was collected just above in the lower eolian facies of the Angastaco Formation (Figure 2) (B. Carrapa et al., submitted manuscript, 2011). 5.2.3. Sierra de los Colorados Sample Locations [36] The Sierra de los Colorados is a broad, west vergent anticline developed in the hanging wall of the Calchaquí Fault (Figure 3). The ridge is composed of Pirgua Subgroup rift sediments, and is dissected in several places by antecedent drainage canyons. Sample 04.29.08‐01 was collected in the core of the anticline where it is cut by a narrow canyon, in a conglomeratic interval of the Los Blanquitos Formation within the Pirgua Subgroup (Figure 2). Sample 04.14.08‐01 was also collected in upper Pirgua Subgroup TC3003 conglomerate, approximately three meters east of the Calchaquí Fault in the overturned west limb of the same anticline. Because sample 04.29.08‐01 was collected in the core of the anticline, it lies stratigraphically ∼400 m below sample 04.14.08‐01. 5.2.4. Rio de Las Conchas Sample Locations [37] Two samples were collected along the Rio de las Conchas, and represent two different fault‐bounded segments of the EC. Sample LY is from the lowermost part of the Pirgua Subgroup in the La Yesera Formation. It was collected in the hanging wall of the southernmost tip of the Calchaquí Fault, also known as the Zorrito Fault [Mon and Salfity, 1995], at the south tip of the Sierra de Leon Muerto where Cretaceous rift rocks are thrust over Cenozoic sedimentary rocks (Figure 2). Sample ALP is from the upper Pirgua Subgroup in the Los Blanquitos Formation, and was collected in the hanging wall of the La Viña Fault near the town of Alemanía along the southwestern margin of the Lerma Valley (Figure 2). 5.2.5. Tin‐Tin Sample Locations [38] Two samples were collected in the northern Calchaquí Valley southeast of the town of Cachi. Sample QCTin250, was collected in the Quebrada de los Colorados Formation in the northern Calchaquí Valley in the hanging wall of the Payogasta Fault (Figure 2). Sample 1TT18FT was collected in the Santa Barbara Subgroup at the northeast corner tip of Cerro Tin‐Tin in the hanging wall of the Tin‐ Tin fault (Figure 2). 5.3. AHe Results [39] Fifty individual grain analyses from ten individual rock samples show a scatter of raw AHe dates that systematically vary with the effective uranium concentration of the grain (Table 1) (eU = U + 0.235Th) [Shuster et al., 2006]. For detrital grains, the presence of a strong age‐eU (Figure 8) relationship implies an episode of partial He loss [Flowers et al., 2007]. Although these detrital samples all come from different parts of the EC, geologic constraints suggest that they were all buried to depths greater than 3 km and subsequently exhumed as indicated by >6 km of Cenozoic section that sits stratigraphically above the collected samples [Uliana et al., 1989; Coutand et al., 2006]. Assuming a standard geothermal gradient and standard Durango apatite diffusion kinetics, all these grains would have been thermally reset during burial, and therefore AHe ages should represent cooling following basin incision or exhumation, presumably related to deformation. [40] However, most samples have two groups of grain ages, including a scatter of dates older than depositional age and a cluster of dates younger than depositional age (Figure 8 and Table 1). AHe dates that are older than depositional age (DA) of the formation in which they were collected routinely correlate with high [eU], suggesting that higher radiation damage increased He retentivity and raised the effective closure temperature (Tec) in these grains, allowing only partial resetting, and resulting in older AHe dates. In contrast, the young clusters of dates in many samples correspond to low [eU], indicating that these grains have less radiation damage. The clustering of these young populations suggests that they were more thermally reset during burial. Thus, the older, partially annealed populations can be momentarily disregarded, and the 11 of 30 TC3003 TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION 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 (Ma) (Ma) (Ma) Longitude (mg) (ppm) (ppm) (ppm) (ppm) (nmol/g) (mm) (mm) Fta (mm) (Ma) Pucará Valley Samples AngPuc06, base of Angastaco Fm., eolian sandstone, deposited since 21.4 ± 0.8 Mac a1 a2 a3 a4 a5 a6 QCPucara1, base of Quebrada de Los Colorados Fm, sandstone, deposited ∼37.6 ± 1.2 Mac a1 a2 a3 a4 a5 2.0 2.7 11.6 5.5 16.4 0.65 136 1.5 13.8 1.5 62.9 2.7 6.6 2.0 2.4 1.2 29.3 25.2 19.7 109.5 25.8 68.9 17.6 5.4 7.8 18.2 9.7 4.7 23.4 9.1 31.4 94.1 0.61 0.64 0.70 0.65 0.61 145 121 136 125 127 1.4 30.6 9.7 32.9 37.0 0.64 1.4 5.5 1.5 16.2 1.1 19.4 1.1 8.3 11.6 8.3 59.7 20.6 21.0 54.8 25.7 25.4 55.1 17.1 12.3 19.5 0.60 0.62 0.56 0.59 86 0.65 44 8.5 0.5 13.2 0.8 39 40 49 44 36 47.1 36.2 24.6 8.6 38.3 2.8 2.2 1.5 0.5 2.3 77.0 56.4 35.3 13.3 62.7 4.6 3.4 2.1 0.8 3.8 114 79 0.64 40 31.7 1.9 49.7 3.0 156 132 150 109 67 76 60 72 37 39 33 36 22.1 20.1 26.8 15.8 1.3 1.2 1.6 0.9 37.2 32.3 47.4 26.7 2.2 1.9 2.8 1.6 72 79 100 89 69 0.61 0.64 0.70 0.65 0.61 0.60 0.62 0.56 0.59 25.794°S, 66.282°W 25.784°S, 66.291°W Angastaco Area Samples Ang2EB, base of Angastaco Fm., eolian sandstone, deposited since 19.2 ± 0.7 Mac a2 a3 a4 a5 QC1EB080, top of Quebrada de los Colorados Fm., sandstone, deposited since 37.6 ± 1.2 Mac a1 a2 a3 a4 a5 a6 a7 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 2.0 15.2 1.1 2.9 1.0 2.3 18.8 19.6 53.9 21.8 8.1 66.7 17.3 6.3 44.2 5.89 0.27 0.22 145 102 99 83 0.65 43 71 0.57 36 72 0.57 36 54.3 5.7 6.1 3.3 0.3 0.4 83.0 9.9 10.6 5.0 0.6 0.6 40.1 15.6 110.0 0.59 232 157 0.79 79 6.6 0.4 8.4 0.5 56.9 0.76 83.7 7.85 77.8 23.68 41.8 1.51 61.8 7.82 80.9 0.71 222 231 247 172 160 186 152 147 163 167 141 152 11.5 52.1 72.6 38.2 52.8 9.3 0.7 3.1 4.4 2.3 3.2 0.6 14.6 65.1 88.6 47.9 67.8 11.9 0.9 3.9 5.3 2.9 4.1 0.7 11.5 6.1 10.4 4.2 10.1 25.7 13.2 57.9 9.6 6.0 6.4 26.0 8.7 4.9 32.0 6.2 6.9 4.0 3.5 36.5 11.7 27.1 59.6 7.0 26.9 13.5 0.79 0.80 0.82 0.80 0.78 0.78 76 75 82 76 66 73 25.714°S, 66.175°W 25.715°S, 66.178°W Sierra de los Colorados Samples (Alemania Subbasin) 04.29.08‐01, Pirgua Subgroup, conglomerate, core of anticline, deposited 128–70 Mad a1 a2 a3 a4 04.14.08‐01, Pirgua Subgroup, conglomerate, east of Calchaqui Fault, deposited 128–70 Mad a1 a2 a3 a4 a5 a6 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 0.6 11.8 0.9 7.9 0.8 20.5 44.6 22.3 56.5 26.2 14.1 97.4 63.1 35.4 63.6 0.55 0.46 0.98 70 80 92 65 0.51 30 76 0.57 35 64 0.54 32 4.5 5.7 5.1 0.3 0.3 0.3 8.7 9.9 9.4 0.5 0.6 0.6 2.2 42.8 19.6 127.2 1.98 113 99 0.66 46 17.7 1.1 26.6 1.6 3.7 61.1 23.5 3.8 12.9 5.35 0.21 0.30 0.99 0.38 176 69 134 144 98 60 56 76 88 66 45.3 1.1 3.8 16.8 11.5 2.7 0.1 0.2 1.0 0.7 76.2 2.4 6.1 24.8 21.0 4.6 0.1 0.4 1.5 1.3 9.6 1.3 20.6 0.4 17.7 1.6 8.8 2.3 9.9 0.9 2.8 21.5 32.0 14.3 10.8 5.9 41.2 163.3 18.7 14.2 33.3 12 of 30 0.59 0.47 0.62 0.68 0.55 34 27 40 45 33 25.624°S, 65.991°W 25.725°S, 65.967°W TC3003 TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION Table 1. (continued) 4 Mass U Th eU Sm He Length Width rb Raw Date Est 1s Corr Date Est 1s Latitude (Ma) (Ma) (Ma) Longitude (mg) (ppm) (ppm) (ppm) (ppm) (nmol/g) (mm) (mm) Fta (mm) (Ma) Rio de las Conchas Samples LY, La Yesera Fm, sandstone, hangingwall of Zorrito Fault, deposited 128–90 Mad a3 a4 a5 a6 ALP, Los Blanquitos Fm., sandstone, hangingwall of La Viña Fault, deposited 76–70 Mad a1 a2 a3 a4 a5 a6 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 1.4 5.5 15.0 9.1 95.6 1.4 33.8 183.5 76.9 321.1 0.7 7.0 35.9 15.5 166.7 0.1 1.2 0.2 122 147 86 75 0.61 38 69 0.59 37 62 0.52 31 2.6 2.8 1.7 0.2 0.2 0.1 4.3 4.7 3.4 0.3 0.3 0.2 2.5 60.4 1.1 179 83 0.68 45 3.1 0.6 4.5 0.8 0.5 0.3 0.0 1.1 1.3 112 154 220 121 150 33 50 75 42 58 2.4 2.8 2.3 3.2 4.3 0.1 0.2 0.1 0.2 0.3 4.2 4.0 3.0 4.8 5.7 0.3 0.2 0.2 0.3 0.3 0.9 3.0 10.0 1.7 4.4 37.3 15.4 0.7 60.2 55.1 7.9 62.3 69.0 14.5 5.7 1.9 4.3 4.0 40.8 16.8 1.1 61.2 56.0 60.9 70.0 1.2 54.7 88.1 63 99 150 84 120 0.57 0.71 0.79 0.67 0.75 25.962°S, 65.751°W 25.632°S, 65.629°W Tin‐Tin Samples QCTin250, Quebrada de los Colorados Fm, sandstone, deposited since ∼37.6 ± 1.2 Mad a1 a2 a3 a4 1TT18FT, Pirgua Subgroup, sandstone, deposited 128–70 Mad a1 a2 a3 a4 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 4.7 15.2 4.0 30.2 3.8 5.4 12.8 18.2 22.7 5.4 31.5 61.5 8.3 7.3 7.6 2.0 9.0 0.8 206 160 169 107 0.73 57 112 0.74 56 106 0.71 54 20.2 51.9 19.5 1.2 3.1 1.2 27.7 70.4 27.3 1.7 4.2 1.6 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 1.0 63.0 236.0 118.5 82.7 0.9 65.3 250.9 124.2 97.7 1.2 71.6 158.9 109.0 100.9 2.1 36.5 2.6 85 92 106 75 0.57 35 71 0.56 34 76 0.60 38 3.3 53.8 4.4 0.2 3.2 0.3 5.7 95.4 7.4 0.3 5.7 0.4 25.254°S, 66.120°W 25.113°S, 66.005°W a Ft is alpha ejection correction. r is spherical radius. Max depositional age based on detrital U‐Pb geochronology (B. Carrapa et al., submitted manuscript, 2011). d Salfity and Marquillas [1994]. b c young clustered populations of dates can be interpreted to represent the most recent signal of exhumation (Figure 9). Overall, there is a remarkable eastward younging trend among the young populations of AHe ages. Raw AHe dates systematically young eastward from ∼13 Ma in the Pucará area to ∼4 Ma in the La Viña area, suggesting an eastward younging wave of exhumation that propagated across the southern EC during the Mio‐Pliocene at a rate of ∼8.3 mm/yr (Figure 9). 5.3.1. Pucará Valley Sample Results [41] Six apatites were analyzed from sample AngPuc06. Two of the six grains were younger than the DA of ∼21.4 Ma (B. Carrapa et al., submitted manuscript, 2011), with comparable cooling ages of 13.2 ± 0.8 and 13.3 ± 0.8 Ma (Table 1) and a weighted mean age of 13.3 ± 1.1 Ma. Five apatites were analyzed from sample QCPucara1. Three of the five grains were younger than DA, and have disparate grain ages of ca. 26.7 ± 1.6 Ma, 32.3 ± 1.9 Ma, and 37.2 ± 2.2 Ma. [42] QCPucara1 is stratigraphically lower than AngPuc06 (Figure 4), and therefore should have been heated to a greater temperature than AngPuc06 during burial, but QCPucara1 shows AHe dates considerably older than AngPuc06. We interpret the two youngest AngPuc06 grain dates to represent exhumation of the Pucará area because they have comparable ages, show the lowest eU of any apatites from the Pucará area, and are consistent with other constraints on basin folding at ∼13.4–12.1 Ma [Marrett et al., 1994]. Although we recognize that model results for sample QCPucara1 are equivocal, the fact that this age match well with evidence of growth in the adjacent Angastaco area at ∼14 Ma supports our interpretation that this part of the basin was exhuming at this time possibly as a result of tilting and uplift of the east verging limb of the Sierra de Quilmes hanging wall (Figure 3). 5.3.2. Angastaco Area Sample Results [43] Four apatites were analyzed from sample Ang2EB (Table 1). Two of the four grain ages are younger than DA, and have similar grain ages of 9.9 ± 0.6 Ma and 10.6 ± 0.6 Ma. Seven apatites were analyzed from sample QC1EB080. Three of the seven grain ages are younger than DA, with dates of 8.4 ± 0.5 Ma, 11.9 ± 0.7 Ma, and 14.6 ± 0.9 Ma. Young AHe grain ages from these two samples correlate strongly with eU, suggesting a basin history where 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 represents depositional age. Formations spent a significant amount of time within the HePRZ. Additionally, because these two samples are stratigraphically close to one another, it can be assumed that they underwent similar thermal histories. Whereas the oldest group of ages (Table 1) can be explained by partial annealing of different closure temperatures depending on different compositions, the weighted mean of the youngest five grain ages is used to approximate the timing of exhumation at 10.4 ± 2.5 Ma. 5.3.3. Sierra de los Colorados Sample Results [44] Four apatites were analyzed from sample 04.29.08‐ 01, all of which are significantly younger than DA, with ages between 5.7 ± 0.3 Ma and 9.4 ± 0.6 Ma. Six grains were analyzed from sample 04.14.08‐01. Five of the six grains give ages younger than DA, and cluster into two distinct populations. The two youngest ages of 2.4 ± 0.1 Ma and 6.1 ± 0.4 Ma generally correlate with the young exhumation signal of sample 04.29.08‐01, but three other grains yield significantly older ages, which cluster around 21.0 ± 1.3 to 26.6 ± 1.6 Ma. The grain yielding an age of 2.4 ± 0.1 Ma is the smallest crystal of the entire study, with crystal dimensions of 56 mm × 69 mm; this is likely the cause for its anomalously young AHe date. Both of these samples were collected within the same structural, and topographic feature (the Sierra de los Colorados anticline), however sample 04.29.08‐01 despite being farther from the fault, with respect to sample 04.14.08‐01, it is from a deeper stratigraphic level (Figure 2). The emplacement of a relatively warm hanging wall over a cool footwall causes thermal perturbations that depress isotherms and reduce lateral thermal homogeneity, and there is some evidence that frictional heating along the fault plane can locally affect thermochronological systems [Ruppel and Hodges, 1994; Ehlers and Farley, 2003]. This could explain the wide spread of AHe dates in sample 04.14.08‐01. Because sample 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 sample good cooling paths through the HePRZ as modeled in HeFTy. pest stratigraphic interval of the anticline, it was likely subjected to greater temperatures, and it is not surprising that it represents a simpler, less altered, and more thermally reset signal of cooling and exhumation. 5.3.4. Rio de Las Conchas Sample Results [45] Four apatites were analyzed from sample LY, all of which are younger than DA, with dates between 3.4 ± 0.2 Ma and 6.0 ± 0.4 Ma. Six apatites were analyzed from sample ALP, all of which are younger than DA, with ages between 3.0 ± 0.2 Ma and 5.7 ± 0.3 Ma. The fact that grain ages are significantly younger than DA suggest that both samples were completely thermally reset and that all these ages represent a signal of recent cooling. 5.3.5. Tin‐Tin Sample Results [46] Four apatites were analyzed from sample QCTin250. Two of the four grain ages are younger than DA, with similar ages of 27.3 ± 1.6 Ma and 27.7 ± 1.7 Ma. Four apatites were analyzed from sample 1TT18FT. Three of 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 young 6–12 Ma signal of sample 1TT18FT can thus be interpreted to represent recent main exhumation of the northern Calchaquí Valley. 5.4. Modeling of AHe Data [47] Sample thermal history information is extracted employing an inverse modeling approach using a constrained Monte Carlo simulations that takes into account the effects of grain size, radiation damage, and cooling rate on the thermal history of each sample (HeFTy) [Ketcham, 2005; Flowers et al., 2009]. Sandstone samples in this study show a relationship between grain ages and [eU], making them prime candidates for inverse modeling using the radiation damage trapping model of Shuster et al. [2006]. Radiation damage impedes the diffusion of 4He from the apatite crystal lattice, and therefore diffusion coefficient (D) of an apatite varies between grains and evolves within grains as a function of time. In a series of laboratory diffusion experiments, Shuster et al. [2006] empirically calculated the relationship between the diffusion coefficient (D) and the volume of radiation damage (Vrd) within the grain, and the radiation damage trapping model exploits this laboratory constrained age‐Vrd relationship to greatly restrict the number of possible time‐temperature (t‐T) paths that could reproduce the AHe dates observed in nature. Inverse modeling of AHe grain ages predicts t‐T paths that would produce the observed relationship between Tc and radiation damage. Numerical inverse modeling can thus help determine the detailed thermal history of a sample. 5.5. HeFTy Modeling Constraints [48] Thermal models were run for multiple grain ages from six AHe samples (Table 2). In all cases, He models were run using the radiation damage trapping model calibration of Shuster et al. [2006] (Do/a2) and the static alpha ejection correction of Farley et al. [1996]. Gradual node randomization was chosen to reflect steady burial, and episodic node randomization was used to represent exhumation accommodated by potentially erratic fault movements. For each sample, wide model constraints based on geological plausible 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 depositional 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 represent 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 parameters can be found in Table 2. Sample locations are indicated on Figures 2, 3, and 4. 16 of 30 TC3003 TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION Table 2. HeFTy Modeling Constraints Deposition Burial Present Sample Age (Ma) T (°C) Age (Ma) T (°C) Age (Ma) T (°C) AngPuc06 Ang2EB QC1EB080 04.29.08‐01 LY ALP 18–21.4 14–19.2 19.2–37.6 70–128 90–128 70–76 20 20 20 20 20 20 0–21.4 0–19.2 0–19.2 0–128 0–128 0–76 30–100 40–100 40–100 40–200 40–200 40–200 0 0 0 0 0 0 10–25 10–25 10–25 10–25 10–25 10–25 most recent cooling and exhumation (Figure 10). Although thermal models cannot determine the exact path of exhumation, they can greatly constrain a window of time and temperature when the sample began cooling through the HePRZ (40°–80°C). For all models, this window of exhumation has been interpreted, and it superimposed upon Figure 9 as a cloud of time during which exhumation through the HePRZ was occurring for each location. Model AngPuc06 was conducted using grains a1 and a5, and suggests that early cooling likely initiated by ∼14 Ma, and that the sample had cooled through the HePRZ by ∼12 Ma (Figure 10a). These results and presence of an age‐eU relationship suggests that sample AngPuc06 cooled slowly, and likely spent an extended period of time in the HePRZ. [51] Model Ang2EB was conducted using grains a4 and a5, and suggests that early cooling initiated ∼11 Ma, with cooling through the HePRZ by ∼8 Ma (Figure 9). These results and the presence of an age‐eU relationship suggests that sample Ang2EB spent significant time in the HePRZ. [52] Model QC1EB080 was conducted using grains a1 and a7. Although HeFTy was never able to find a “good” path for these grains, 1500 “acceptable” paths were found that suggest cooling began by ∼13 Ma, with cooling through 70°C by 12 Ma (Figure 10c). Modeling of samples 04.29.08‐01, LY, and ALP all suggest rapid cooling. Model 04.29.08‐01 was conducted using grains a2, a3, and a4, and suggests that early cooling began ∼10 Ma, with cooling through the HePRZ at ca. 6–7 Ma (Figure 10d). Model LY was conducted using grains a4 and a5, and suggests that early cooling began by ∼6 Ma, with final cooling through the HePRZ at ∼2–4 Ma (Figure 10e). Model ALP was conducted using grains a1, a2, a3, and a5, and suggests that early cooling through the HePRZ had initiated at ∼35– 22 Ma, with final cooling through the 80°–40° T‐window at ca. 0–4 Ma (Figures A1 and 10f). 5.7. Interpretation of AHe Ages and HeFTy Modeling Results [53] The detrital AHe system provides information on the latest stage of exhumation in the uppermost crust, and by exploiting the effects of radiation damage on multiple crystals using HeFTy modeling, we can expand our confidence in the interpretation of recent cooling. While AHe thermochronology cannot tell us about events that occurred while the sample was buried beneath the HePRZ, incorporating the data with reasonable assumptions and known geological constraints allows us to make interpretations that greatly elucidate the recent patterns of exhumation in the region. [54] For the six Cenozoic samples that contain an old population of partially to nonreset grain ages (Pucará, Grains Modeled a1, a4, a1, a2, a4, a1, a5 a5 a7 a3, a4 a5 a2, a3, a5 Total Paths Acceptable Paths Good Paths 2,003 36,892 308,236 120,547 34,514 27,180 168 2,871 1,500 1,381 357 237 100 100 0 100 100 100 Angastaco, and Tin‐Tin samples), it can be confidently interpreted that the samples were not completely buried beneath the HePRZ, and therefore young populations of AHe dates closely represent the onset of exhumation. In the Angastaco area, where corresponding apatite fission track ages have not been annealed during burial [Coutand et al., 2006], there is further confidence in interpreting young AHe ages as reflecting the onset of exhumation of the basin strata. [55] For the two Cretaceous sandstone samples collected in the Sierra de los Colorados anticline, just east of Angastaco area, the sample collected in the deepest stratigraphic portion of anticline has been fully reset during burial (sample 04.29.08‐01), whereas the shallower sample contains a mixture of nonreset, partially reset, and fully reset ages (sample 04.14.08‐01) (see Table 1). Thus, the shallower sample was likely buried to a maximum depth within the HePRZ, while the deeper sample experienced temperatures below the HePRZ; differences in composition have contributed to the observed scatter of ages. Because these samples are stratigraphically ∼400 m apart, sample 04.29.08‐01 was buried to a maximum depth just below the HePRZ, and therefore should have entered the HePRZ shortly to immediately after the onset of Neogene exhumation. HeFTy modeling suggests that this sample cooled through the HePRZ at rate ≥0.5 mm/yr, and therefore the ∼400 m stratigraphic offset between the two samples can be used to infer that the maximum time between onset of exhumation and entering the HePRZ is 0.8 Myr. This suggests that the AHe ages of sample 04.29.08‐01 do represent the onset of exhumation for this range, and that any variability is within error of the AHe signal. [56] The Cretaceous samples LY and ALP have been completely thermally reset, and all grain ages are significantly younger than DA. While the AHe data for these samples do not necessarily constrain the onset of exhumation for these samples, HeFTy modeling shows with confidence that they have experienced rapid cooling during the last ∼6 Myr, which we interpret to reflect major range exhumation caused by the onset of major active deformation on adjacent reverse faults. [57] The majority of the analyzed grains give ages that can be interpreted to roughly represent the onset of basin exhumation in the southern EC. The AHe grain ages and HeFTy model results indicate that the sequence of exhumation in the southern EC has swept consistently eastward during the Miocene and Pliocene (Figure 10). We interpret the younging in cooling ages eastward as the result of erosion related to progressive eastward younging of fault activation. A linear regression placed through the combined young grain age populations suggests that this wave of 17 of 30 TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION exhumation (and inferred deformation) propagated eastward at a rate of ∼8.3 mm/yr (r2 = 0.7) (Figure 10). There is an interesting recurrence of a second population of young grain ages between 21 and 28 Ma in multiple parts of the study area. We interpret this to represent a possible early, minor exhumation signal between deposition of the Quebrada de los Colorados and Angastaco Formations. 6. Discussion 6.1. Angastaco Area [58] This study confirms that deposition of the Miocene‐ Pliocene sedimentary package in the Angastaco area was contemporaneous with local deformation. U‐Pb dating in this study and B. Carrapa et al. (submitted manuscript, 2011) limits the age range of Cenozoic strata in the Angastaco area to ∼37–2.3 Ma. The dominance of flexural slip folding implies that the exposed strata experienced brittle deformation at low temperature and confining pressure. Based on detrital U‐Pb geochronology from strata below the angular unconformity shown in Figure 6, synsedimentary deformation initiated in the area by ∼14 Ma. AHe thermochronology documents significant rapid cooling along the western margin of the Angastaco area by no later than 10 Ma, demonstrating that the western part of the basin was exhuming coevally with active deposition in the central and eastern parts of the basin. Thus, much of the sediment in the Palo Pintado and San Felipe Formations may have been recycled from the Quebrada de los Colorados and Angastaco Formations just a few kilometers to the west. It is interesting to note that samples 04.29.08‐01 and 04.29.08‐01 from the Sierra de los Colorados yield cooling ages between 6 and 10 Ma. This indicates that cooling and exhumation in, at least a portion of, this range initiated before the change in provenance and basin reorganization documented by paleocurrent and conglomerate deposition at 4 Ma [Bywater‐Reyes et al., 2010]. In turn, this suggests that despite being deformed and eroded as early as ∼10 Ma, the Sierra de los Colorados did not have enough topographic relief to shed coarse material to the surrounding areas and that it was likely not an orographic barrier. More likely this earlier exhumation is related to the formation of local, low relief. This is supported by fine‐grained Miocene deposits of the Anta and Jesus Maria formations in the La Viña area, which constitute the distal equivalents of the Angastaco Formation [Starck and Anzótegui, 2001; B. Carrapa et al., submitted manuscript, 2011]. Overall the data indicate that the main river system was still draining to the east and was not disrupted by the regional uplift of ranges to the east of the Angastaco area, which did not occur until ∼4 Ma when conglomerates derived from the Sierra de los Colorados range were deposited in the Angastaco and La Viña areas. 6.2. Recent Exhumation in the Southernmost EC [59] AHe data show an eastward younging of cooling ages that suggests an eastward migrating wave of exhumation and related deformation during the Miocene‐Pliocene. These data document middle Miocene exhumation of the Pucará area and initiation of structural growth within the Angastaco area at ∼14 Ma (Figure 10). The western Angastaco area experienced significant exhumation by TC3003 ∼10 Ma, at which point deformation and exhumation migrated eastward as the Sierra de los Colorados anticline experienced rapid exhumation ∼9 Ma. The Pucará, Angastaco, and Sierra de los Colorados areas continued to exhume as the Sierra de Leon Muerto and western Lerma Valley experienced significant exhumation ∼6–3 Ma (Figure 10). This timeline corresponds to the timing of shortening proposed by Marrett et al. (∼13–1 Ma) [1994], as well as the eastward propagation of growth structures proposed in the Mio‐Pliocene by Carrera and Muñoz [2008]. This pattern of eastward younging deformation involved mainly westward verging thrust faults, and suggests that, regardless of local structural geometry, overall orogenic strain propagated eastward as expected during the development of a tapered orogenic wedge. 6.3. Evolution of Orogenic Propagation Across the Southern Central Andes [60] Our new data set, when compared to existing data, documents a wave of exhumation that propagated across the southern EC of NW Argentina at 8.3 mm/yr between ∼14 and ∼3 Ma (Figure 11), which is roughly in agreement with the average rate of strain propagation across the entire central Andes (∼7.8 mm/yr) at the same latitude (Figure 11). Major development of the southern part of the central Andes began during the mid to late Cretaceous in the Salar de Atacama basin [Arriagada et al., 2006; Mpodozis et al., 2005]. Later, during the middle to late Eocene, the deformation front jumped eastward across the width of the Puna region, causing a roughly synchronous onset of deformation from ∼40 to 38 Ma in the Salar de Antofalla, Salar de Arizaro and Salar de Pastos Grandes (and possibly La Poma) basins (Figure 11) [Kraemer et al., 1999; Voss, 2002; Carrapa et al., 2005; Mpodozis et al., 2005; Arriagada et al., 2006; Jordan and Mpodozis, 2006; Hongn et al., 2007; Carrapa and DeCelles, 2008]. Deformation was localized within the eastern Puna/western EC of NW Argentina between ∼39 and ∼22 Ma and swept eastward across the eastern EC during the Miocene‐Pliocene (Figure 11). Evidence of deformation and exhumation in the Puna Plateau at ca. 25 Ma [Carrapa et al., 2005] suggests out‐of‐sequence deformation of the hinterland at this time. [61] Despite similarities of deformation during the early Cenozoic, the Oligocene and Neogene kinematic histories in Bolivia and NW Argentina are significantly different. Shortening in Bolivia continued to propagate eastward through the EC and inter‐Andean zone throughout the Oligocene [McQuarrie et al., 2005; Ege et al., 2007], whereas the Puna experienced internal deformation and exhumation [Carrapa et al., 2005] with very limited eastward propagation during this time interval (Figure 11). During Miocene‐Pliocene time (∼22–3 Ma) deformation was concentrated within the EC/Santa Barbara system of NW Argentina (∼26°S, 66°W) and it was migrating through the study area in sequence between ∼14 and 3 Ma. In the sub‐Andes of NW Argentina‐S Bolivia [Echavarria et al., 2003] deformation was migrating through the region between ∼8 and 3 Ma and it was in the sub‐Andes of Bolivia by ∼10 Ma [Gubbels et al., 1993]. [62] Interesting questions that arise from this comparison are (1) why did different segments of the central Andes 18 of 30 TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION behave differently during the Neogene and (2) what are the possible controlling mechanisms? Mass removal underneath 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 Cordilleran type orogenic systems such as the Andes [DeCelles et al., 2009]. In particular the abrupt eastward shifts of the deformation front observed in Bolivia at ca. 45–40 Ma and again at 10 Ma have been explained as a response to isostatic elevation gain following the removal of dense arc roots beneath the orogenic hinterland [DeCelles et al., 2009; Kay et al., 1999]. However, significant along‐strike variations in the Neogene kinematic history between the study area and regions farther north suggest differences in orogenic processes. For example the Altiplano plateau to the north has been explained in terms of orogenic‐scale lithospheric removal [Garzione et al., 2006]. Alternatively the whole region has been explained by continuous shortening and gradual uplift [Ehlers and Poulsen, 2009]. The Puna Plateau to the south has been associated with smaller‐scale, and perhaps diachronous, lithospheric removal [Carrapa et al., 2009]. Thus, orogenic cyclicity has likely not operated uniformly over along‐strike distances of >500 km, perhaps because of differences in shortening rates, arc melt and dense residue production rates resulting in different upper crustal to surficial structural responses. A younger lithospheric removal event in the Puna Plateau at ∼3 Ma as proposed by DeCelles et al. [2009] is consistent with TC3003 migration of the deformation front into the Santa Barbara system after 4 Ma. 7. Conclusions [63] Although the central Andes express great variability in timing and structural style along strike, remarkable similarities exist between the timing and rates of propagation of deformation and exhumation in the northern (Bolivian) and southern (Argentinean) portions of the central Andes, particularly when major structural features are followed along strike regardless of the established tectonomorphic nomenclature. The development of the Bolivian EC was roughly coeval with the Eocene‐Oligocene internal deformation of the Puna region, and the development of the Argentinean EC was coeval with the development of the sub‐Andean system (Figure 11) [Carrera and Muñoz, 2008; Echavarria et al., 2003; McQuarrie et al., 2005, 2008; Uba et al., 2009]. Many of the geomorphic and structural features of the Puna can be traced northward along strike into the Bolivian EC (Figure 11), and because both provinces contain copious evidence of Oligocene deformation, it appears that the eastern part of the Puna correlates both geologically and temporally with the Bolivian EC and inter‐Andean zone. [64] If we temporarily overlook the geographic constrictions implied by predefined tectonomorphic provinces, and instead focus on regions of the central Andes that correlate 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, stratigraphy, 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 multidisciplinary 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 Figure 11 20 of 30 TC3003 TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION mation) of the NW Argentinean EC is comparable to propagation of deformation rates in the thin‐skinned sub‐ Andes during Mio‐Pliocene time. Evidence for this comes from a variety of sources (Figure 11). [65] In summary, whereas the Mio‐Pliocene rate of eastward propagation of the orogenic front varies along strike, the propagation rate in the study area of NW Argentina (8.3 mm/yr) falls within the same range of rates for the EC of Bolivia (2.1–15 mm/a), and is remarkably similar to the long‐term average rate of orogenic propagation in the central Andes (7.8 mm/yr) [McQuarrie et al., 2005]. More importantly, similar rates of propagation of the deformation front along the plateau margin characterized by significantly different structural styles suggests that the rate at which the deformation front has driven into the foreland is determined more by the dynamics of regional eastward expansion of the Andean orogeny than the by local structural style, and therefore may develop independently of features that might control local structural style (e.g., the nature of the local basement). It also interesting to note that rates of propagation of deformation are similar along strike despite of the significant differences in amounts of shortening between Bolivia (>300 km) [Kley et al., 1997; McQuarrie, 2002] and Argentina (<100 km) [Kley and Monaldi, 1998, and references therein]. This may be explained by the fact that the Altiplano and Eastern Cordillera of Bolivia have deformed mainly by simple shear whereas the Puna Plateau and Eastern Cordillera of NW Argentina by pure shear [Allmendinger and Gubbels, 1996]. [66] The differences in the timing of late Miocene‐ Pliocene deformation between the study area and regions farther north suggest different structural and surficial responses to lithospheric processes. For example the late Cenozoic uplift history of the Altiplano in Bolivia has been associated with large‐scale lithospheric removal [Garzione et al., 2006], which has been used to explain changes in taper angle and subsequent migration of the orogenic front outward in the sub‐Andes at ∼8 Ma [DeCelles et al., 2009]. In NW Argentina, small‐scale lithospheric removal has been proposed to have influenced Puna Plateau evolution [Kay and Coira, 2009], late Cenozoic sedimentary basin evolution and paleogeography [Strecker et al., 2009; Carrapa et al., 2008b]. However, in this last case the size, volume and possible diachrounous removal of lithosphere has probably prevented a syncronous response of deformation forelandward. Appendix A A1. U‐Pb Lab Methods [67] Mineral separation was conducted at the University of Wyoming using a jaw crusher, disk mill, wilfley table, Frantz electromagnetic separator, and methyl iodide heavy liquids. Detrital zircon U‐Pb analyses were conducted at the Arizona LaserChron center using standard methods of laser ablation multicollector inductively coupled plasma mass spectrometry (LA‐MC‐ICP MS). U‐Pb data are presented in Table A1. Zircons were ablated with a New Wave/Lambda Physic DUV193 Excimer laser using a spot diameter of 35 mm, material was transported by He gas to the plasma source of a GVI Isoprobe, and isotopes were measured TC3003 simultaneously in static mode. Measurement of 238U, 232U, 208 Pb, and 206Pb was conducted with eight 1011 W Faraday detectors. A 1012 W Faraday collector measured 207Pb, and 204 Pb was measured with an ion‐counting channel. Ion yields were ∼1.0 mv per ppm. A typical analysis included 20 s of background measurement, 12 s integrated measurements as the laser fired, and 30 s to purge the instrument and prepare for the next analysis. Unknowns were calibrated against a Sri Lanka standard of known age after every fifth zircon analysis to correct for apparent fractionation of Pb isotopes and inter element fractionation of Pb/U. Corrections for common Pb were calculated using measured 204Pb and an assumed initial Pb composition from Stacey and Kramers [1975]. Measurement uncertainty was generally 1–2% (2s). Because of their sensitivity to young ages, 206 Pb/238U ages were used, and analyses with more than 30% discordance between 206Pb/238U and 206Pb/207Pb ages were removed from consideration. For more information concerning lab methods at the Arizona LaserChron center, see Gehrels et al. [2008]. A2. (U‐Th)/He Lab Methods [68] Mineral separation and grain picking were conducted at the University of Wyoming. Apatites were separated from ten detrital samples using a jaw crusher, disk mill, wilfley table, Frantz electromagnetic separator, and methyl iodide heavy liquids. Using a Leica M165C stereomicroscope, each apatite fraction was thoroughly screened for relatively euhedral apatites that were free of cracks and inclusions and had minimum dimensions greater than 60 mm. Unfortunately, detrital apatites are often small, broken, or abraded, and thus pose a significant challenge during grain picking. Although six individual grains were desired, in some samples only four acceptable grains could be found. All grains were photographed and measured for Ft calculations using the method of Farley et al. [1996]. [69] Apatite U‐Th/He analyses were conducted using the ultra high vacuum noble gas extraction and purification line and ICP‐MS at the University of Kansas. For each single‐ grain aliquot, 4He was extracted with a continuous mode Nd‐YAG laser, spiked with 3He for isotope dilution, concentrated and purified with in a Janis cryogenic trap, released, and the 3He/4He ratio was subsequently measured with a Blazers Prisma QMS‐200 quadrupole mass spectrometer. After degassing, each sample was analyzed for U, Th, Sm, and selected REE using a Fisons/VG PlasmaQuad II Inductively Coupled Plasma Mass Spectrometer (ICP‐ MS). Samples were routinely calibrated against Durango apatite standard to ensure laboratory accuracy. A3. Additional (U‐Th)/He Details [70] Radiation damage introduces isolated structural defects and vacancies into the apatite crystal structure that can substantially retard the diffusion of 4He and lead to anomalously old ages [Shuster et al., 2006; Shuster and Farley, 2009]. Radiation damage controls AHe dates more than any other known factor, is twice as influential as grain size, can cause expansion of the HePRZ to 50°–115°C, and generally correlates with [eU], the effective uranium concentration of the grain (eU = U + 0.235Th) [Shuster et al., 2006]. Because radiation damage accumulates over geo- 21 of 30 SAMPLE 04.14.08‐02 04140802‐1 04140802‐3 04140802‐4 04140802‐5 04140802‐6 04140802‐7 04140802‐8 04140802‐9 04140802‐10 04140802‐11 04140802‐12 04140802‐13 04140802‐15 04140802‐16 04140802‐17 04140802‐18 04140802‐19 04140802‐21 04140802‐22 04140802‐23 04140802‐24 04140802‐25 04140802‐27 04140802‐28 04140802‐29 04140802‐30 04140802‐31 04140802‐33 04140802‐34 04140802‐35 04140802‐36 04140802‐38 04140802‐39 04140802‐40 SAMPLE 05.07.08‐A.Unc 050708‐AUNC‐1 050708‐AUNC‐2 050708‐AUNC‐3 050708‐AUNC‐4 050708‐AUNC‐5 050708‐AUNC‐6 050708‐AUNC‐7 050708‐AUNC‐8 050708‐AUNC‐9 050708‐AUNC‐10 050708‐AUNC‐11 050708‐AUNC‐12 Analysis Isotope Ratios Apparent Ages (Ma) 22 of 30 17.2735 17.7149 17.6687 17.6122 17.6579 17.8076 17.8131 17.6944 17.5969 17.4002 17.9142 16.0399 11614 6966 15192 11976 17030 15958 15190 22476 14036 60018 14664 43230 600 322 424 312 352 352 366 597 421 574 403 891 1.7 1.5 2.2 1.8 1.6 2.2 1.9 3.3 2.1 1.6 1.5 5.8 2.1 24.1 1.3 95.1 1.9 2.2 1.9 6.8 6.1 2.6 79.2 15.3 30.9 60.8 9.3 11.6 2.4 1.4 14.9 2.7 40.2 2.8 24.0 6.1 1.8 12.5 16.3 22.0 4.3 3.8 20.8 60.6 29.2 1.8 0.6104 0.6014 0.5939 0.6150 0.6102 0.5912 0.5904 0.5948 0.6041 0.6069 0.5906 0.7311 0.6082 0.0022 0.7920 0.0021 0.6636 1.6792 0.6016 0.0023 0.0026 1.5330 0.0023 0.0049 0.0015 0.0024 0.0030 0.0082 0.5868 5.2650 0.0021 1.5470 0.0014 0.5110 0.0022 0.0074 0.6831 0.0027 0.8628 0.0023 0.7019 0.6144 0.0023 0.0021 0.0018 1.8811 2.4 2.0 2.7 2.4 2.1 3.0 2.2 3.5 2.4 2.2 2.0 7.0 2.3 24.4 1.7 95.5 2.5 3.8 2.1 7.3 6.4 2.7 79.3 15.7 30.9 60.9 9.4 11.6 3.2 1.7 15.2 4.8 40.2 3.9 24.1 6.1 2.5 12.5 16.3 22.1 4.4 4.0 20.9 60.6 29.5 2.0 0.0765 0.0773 0.0761 0.0786 0.0781 0.0764 0.0763 0.0763 0.0771 0.0766 0.0767 0.0851 0.0773 0.0004 0.0959 0.0003 0.0839 0.1667 0.0777 0.0003 0.0004 0.1548 0.0005 0.0006 0.0003 0.0004 0.0004 0.0013 0.0740 0.3329 0.0003 0.1447 0.0003 0.0649 0.0004 0.0011 0.0850 0.0004 0.0891 0.0004 0.0834 0.0792 0.0004 0.0004 0.0003 0.1810 1.7 1.3 1.6 1.6 1.4 2.0 1.1 1.1 1.3 1.5 1.3 3.8 0.9 3.9 1.1 8.3 1.6 3.2 0.9 2.6 2.1 0.5 4.0 3.4 1.1 4.4 1.6 0.6 2.2 1.1 3.1 3.9 2.2 2.6 2.1 0.7 1.8 1.5 0.9 1.2 0.8 1.1 1.8 1.5 4.0 0.9 0.69 0.64 0.60 0.66 0.68 0.67 0.50 0.32 0.52 0.68 0.66 0.55 0.41 0.16 0.66 0.09 0.65 0.82 0.41 0.36 0.32 0.19 0.05 0.22 0.04 0.07 0.16 0.05 0.68 0.63 0.20 0.82 0.06 0.69 0.09 0.11 0.71 0.12 0.05 0.05 0.19 0.27 0.09 0.02 0.14 0.46 475.0 479.8 472.8 487.5 485.0 474.3 473.9 474.2 478.8 475.7 476.6 526.2 479.8 2.3 590.1 1.8 519.4 993.9 482.2 2.1 2.5 927.8 3.4 4.1 2.0 2.3 2.8 8.7 460.4 1852.6 2.2 871.3 2.2 405.4 2.3 7.4 525.8 2.7 550.0 2.4 516.2 491.2 2.3 2.6 2.1 1072.7 7.7 5.9 7.4 7.4 6.7 9.1 5.1 5.1 5.8 6.8 6.2 19.2 4.3 0.1 6.2 0.1 8.1 29.0 4.0 0.1 0.1 4.3 0.1 0.1 0.0 0.1 0.0 0.0 9.8 17.6 0.1 31.9 0.0 10.4 0.0 0.1 9.1 0.0 4.6 0.0 4.1 5.1 0.0 0.0 0.1 9.0 483.8 478.1 473.3 486.7 483.7 471.6 471.1 473.9 479.8 481.6 471.2 557.2 482.5 2.2 592.3 2.1 516.8 1000.7 478.3 2.3 2.6 943.7 2.3 5.0 1.5 2.4 3.0 8.3 468.8 1863.2 2.2 949.3 1.4 419.1 2.2 7.5 528.6 2.7 631.7 2.3 539.9 486.3 2.4 2.1 1.8 1074.5 9.4 7.6 10.2 9.2 8.2 11.2 8.4 13.3 9.2 8.3 7.7 29.8 8.8 0.5 7.5 2.0 10.2 24.5 8.1 0.2 0.2 16.4 1.8 0.8 0.5 1.5 0.3 1.0 12.1 14.9 0.3 29.5 0.6 13.2 0.5 0.5 10.4 0.3 77.0 0.5 18.5 15.4 0.5 1.3 0.5 13.1 525.7 470.1 475.9 483.0 477.2 458.5 457.8 472.6 484.9 509.6 445.3 686.0 495.2 −116.2 600.9 417.0 505.5 1015.6 459.5 277.6 103.5 981.1 −1059.5 425.5 −846.2 96.0 161.1 −106.1 510.2 1875.1 −23.9 1134.8 −1275.4 495.5 −106.4 57.1 540.9 29.7 935.8 −115.8 641.3 463.4 66.5 −491.4 −275.1 1078.2 38.4 34.0 47.6 39.8 34.7 49.0 42.8 73.9 45.7 35.1 34.0 124.3 46.1 601.0 27.3 673.7 42.1 44.8 43.1 156.1 143.2 53.4 1167.2 344.0 901.6 1583.6 218.4 287.0 51.8 24.5 363.3 54.6 1309.6 61.8 598.3 145.7 38.6 299.5 336.8 548.4 93.4 85.0 500.7 1759.7 756.6 35.2 475.0 479.8 472.8 487.5 485.0 474.3 473.9 474.2 478.8 475.7 476.6 526.2 479.8 2.3 590.1 1.8 519.4 1015.6 482.2 2.1 2.5 981.1 3.4 4.1 2.0 2.3 2.8 8.7 460.4 1875.1 2.2 1134.8 2.2 405.4 2.3 7.4 525.8 2.7 550.0 2.4 516.2 491.2 2.3 2.6 2.1 1078.2 7.7 5.9 7.4 7.4 6.7 9.1 5.1 5.1 5.8 6.8 6.2 19.2 4.3 0.1 6.2 0.1 8.1 44.8 4.0 0.1 0.1 53.4 0.1 0.1 0.0 0.1 0.0 0.0 9.8 24.5 0.1 54.6 0.0 10.4 0.0 0.1 9.1 0.0 4.6 0.0 4.1 5.1 0.0 0.0 0.1 35.2 90.4 102.1 99.4 100.9 101.6 103.4 103.5 100.3 98.7 93.3 107.0 76.7 96.9 −2.0 98.2 0.4 102.7 97.9 105.0 0.8 2.4 94.6 −0.3 1.0 −0.2 2.4 1.8 −8.2 90.2 98.8 −9.1 76.8 −0.2 81.8 −2.2 12.9 97.2 9.0 58.8 −2.1 80.5 106.0 3.5 −0.5 −0.8 99.5 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION 2.8 12.2 12.1 3.3 6.3 7.0 15.8 11.6 12.6 11.6 4.6 47.7 18315 5.0 17.5152 1017 4.2 22.7779 54378 2.2 16.6871 411 1.2 18.1429 27078 3.5 17.4331 52539 3.4 13.6872 31068 2.5 17.8001 888 2.1 19.2971 1713 5.2 20.7960 12228 3.7 13.9223 729 5.8 32.2610 399 1.2 18.0740 525 0.6 29.9985 705 4.1 20.8624 1137 3.5 20.2923 3363 4.5 22.6853 10602 2.7 17.3961 21492 0.8 8.7188 1359 7.2 21.9338 67716 1.9 12.8987 474 6.2 34.6046 11025 4.5 17.5126 861 3.2 22.6877 5718 99.6 21.2060 16416 2.1 17.1542 1317 3.3 21.4510 7506 3.3 14.2336 915 3.7 22.7749 8193 2.8 16.3782 15750 2.4 17.7686 930 4.4 21.1231 384 0.6 26.3769 903 4.4 24.2709 45420 3.7 13.2694 227 2640 607 1068 284 216 370 3532 5073 53 1064 1057 2654 2237 2810 2660 189 69 3409 437 1228 161 2204 5541 162 2962 255 2294 257 156 2298 1312 2504 178 U 206Pb/ 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Error 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Best Age Standard Concentration (ppm) 204Pb U/Th 207Pb* Error (%) 235U* Error (%) 238U Error (%) Correlation 238U* Error (Ma) 235U Error (Ma) 207Pb* Error (Ma) (Ma) Error (Ma) (%) Table A1. Detrital U‐Pb Data TC3003 TC3003 050708‐AUNC‐13 050708‐AUNC‐14 050708‐AUNC‐16 050708‐AUNC‐17 050708‐AUNC‐18 050708‐AUNC‐19 050708‐AUNC‐20 050708‐AUNC‐21 050708‐AUNC‐22 050708‐AUNC‐23 050708‐AUNC‐24 050708‐AUNC‐25 050708‐AUNC‐26 050708‐AUNC‐27 050708‐AUNC‐28 050708‐AUNC‐29 050708‐AUNC‐30 050708‐AUNC‐31 050708‐AUNC‐32 050708‐AUNC‐34 050708‐AUNC‐35 050708‐AUNC‐36 050708‐AUNC‐37 050708‐AUNC‐38 050708‐AUNC‐39 050708‐AUNC‐41 050708‐AUNC‐42 050708‐AUNC‐43 050708‐AUNC‐44 050708‐AUNC‐45 050708‐AUNC‐46 050708‐AUNC‐47 050708‐AUNC‐48 050708‐AUNC‐49 050708‐AUNC‐50 050708‐AUNC‐51 050708‐AUNC‐53 050708‐AUNC‐54 050708‐AUNC‐55 050708‐AUNC‐56 050708‐AUNC‐58 050708‐AUNC‐59 050708‐AUNC‐60 050708‐AUNC‐61 050708‐AUNC‐62 050708‐AUNC‐63 050708‐AUNC‐64 050708‐AUNC‐65 Analysis Table A1. (continued) Isotope Ratios Apparent Ages (Ma) 140 330 379 287 263 313 340 358 152 512 421 284 511 273 322 335 310 568 630 566 375 513 198 291 292 408 431 818 244 238 403 617 320 282 193 377 505 237 397 446 386 248 297 460 465 205 87 386 6538 6706 16902 14598 22614 23586 14650 17568 8546 22046 19122 17236 22184 16428 8660 30910 16152 26808 20292 24030 8542 8118 6292 9520 6258 13642 31526 19732 10462 19686 19976 28992 19986 8120 18154 14656 20230 9770 24606 35004 18142 13252 28910 26582 31740 27948 8706 23894 5.1 6.1 8.6 2.3 2.3 15.2 8.9 14.4 5.7 25.4 3.8 8.0 14.5 4.9 8.1 5.9 3.0 11.7 15.2 2.3 1.3 10.3 9.7 3.0 3.8 10.5 7.6 2.4 6.3 6.4 6.1 2.9 9.0 2.2 3.9 2.0 7.4 2.2 5.8 8.4 1.6 6.0 4.0 13.5 5.9 1.4 2.5 13.9 17.5819 16.4621 17.8625 17.3930 13.5773 17.5765 17.7632 17.8406 15.6312 17.6155 17.8672 17.1379 17.5288 14.8111 17.4928 13.8530 17.5750 17.5409 17.4866 17.5901 17.5430 17.4223 17.6830 17.6084 18.0625 17.7339 14.2932 17.4104 17.8770 14.8534 14.6417 17.3411 15.1012 19.6097 13.6823 17.2716 17.5961 17.7330 15.4115 15.0197 16.3304 15.7770 13.4216 17.7904 14.4374 13.4373 15.3626 16.3018 3.3 2.0 1.6 2.4 1.3 2.2 1.4 1.1 3.3 1.6 1.3 1.1 2.0 1.8 1.1 2.0 1.4 1.0 1.9 1.7 2.2 0.7 1.8 1.9 2.5 2.1 2.5 1.0 2.6 3.0 1.4 2.1 5.2 2.8 1.6 1.3 1.3 1.6 4.7 2.1 1.6 3.0 1.8 1.9 0.7 1.6 1.3 2.9 0.6253 0.6811 0.5848 0.6619 1.5816 0.6028 0.5914 0.5917 0.7964 0.5855 0.5912 0.6874 0.6137 1.0120 0.6050 1.5746 0.6047 0.6179 0.6158 0.6061 0.6337 0.6173 0.5914 0.6630 0.5831 0.5871 1.2476 0.5900 0.5788 1.0099 1.1280 0.6913 0.7951 0.2926 1.6547 0.6814 0.5898 0.5871 0.8447 0.9829 0.8495 0.7537 1.6919 0.6036 1.2154 1.7904 1.1927 0.7138 3.3 4.3 2.9 3.1 2.5 2.6 1.8 1.6 4.4 2.4 1.6 1.6 2.2 4.2 2.3 3.2 1.7 1.7 2.6 1.8 2.5 1.1 2.1 1.9 2.7 2.7 2.9 1.1 3.2 4.0 2.0 2.3 6.1 2.9 2.1 1.5 1.5 1.7 7.0 2.9 1.9 7.3 2.8 2.0 2.3 1.9 1.5 3.9 0.0797 0.0813 0.0758 0.0835 0.1557 0.0768 0.0762 0.0766 0.0903 0.0748 0.0766 0.0854 0.0780 0.1087 0.0768 0.1582 0.0771 0.0786 0.0781 0.0773 0.0806 0.0780 0.0758 0.0847 0.0764 0.0755 0.1293 0.0745 0.0750 0.1088 0.1198 0.0869 0.0871 0.0416 0.1642 0.0854 0.0753 0.0755 0.0944 0.1071 0.1006 0.0862 0.1647 0.0779 0.1273 0.1745 0.1329 0.0844 0.5 3.8 2.4 1.9 2.2 1.4 1.2 1.1 2.9 1.8 1.0 1.2 1.0 3.8 2.1 2.5 1.0 1.4 1.8 0.6 1.0 0.8 1.1 0.5 0.8 1.8 1.6 0.5 1.8 2.7 1.5 0.8 3.2 0.6 1.4 0.7 0.7 0.7 5.2 2.1 1.0 6.7 2.2 0.7 2.2 1.0 0.6 2.6 0.15 0.88 0.84 0.62 0.86 0.54 0.66 0.68 0.66 0.75 0.63 0.72 0.45 0.91 0.89 0.79 0.59 0.81 0.69 0.35 0.41 0.75 0.50 0.26 0.31 0.66 0.53 0.45 0.56 0.66 0.74 0.34 0.53 0.22 0.64 0.49 0.44 0.39 0.74 0.71 0.54 0.91 0.77 0.34 0.96 0.54 0.43 0.66 494.5 504.0 470.8 517.0 933.1 477.3 473.3 475.6 557.2 465.0 475.9 528.5 484.3 665.2 476.7 946.8 478.7 487.8 484.8 480.2 499.9 484.2 471.3 524.0 474.5 469.3 784.0 463.2 466.5 665.7 729.3 537.4 538.3 262.9 980.1 528.0 467.9 469.3 581.6 655.7 618.0 533.3 982.8 483.4 772.3 1036.8 804.3 522.3 2.4 18.2 11.0 9.6 18.9 6.4 5.5 4.9 15.4 8.0 4.7 5.8 4.7 24.1 9.6 21.8 4.8 6.4 8.3 3.0 4.9 3.9 4.8 2.5 3.8 8.2 11.4 2.3 8.1 16.8 10.1 4.0 16.7 1.6 12.3 3.7 3.0 3.1 28.8 13.0 6.0 34.2 19.6 3.2 16.3 10.0 4.8 12.8 493.1 527.4 467.6 515.8 963.0 479.0 471.7 471.9 594.8 468.0 471.7 531.2 485.9 709.9 480.4 960.2 480.2 488.5 487.2 481.1 498.4 488.2 471.7 516.5 466.5 469.0 822.3 470.9 463.7 708.8 766.8 533.6 594.1 260.6 991.4 527.6 470.8 469.0 621.7 695.1 624.4 570.4 1005.5 479.5 807.7 1042.0 797.2 547.0 13.0 17.6 10.9 12.6 15.7 9.9 6.8 5.9 19.8 8.9 6.2 6.6 8.6 21.5 9.0 19.6 6.7 6.5 9.9 7.0 9.7 4.3 8.0 7.8 10.0 10.3 16.4 4.3 11.9 20.6 10.8 9.4 27.6 6.7 13.4 6.1 5.6 6.5 32.5 14.8 8.8 32.0 17.8 7.6 13.0 12.5 8.1 16.3 486.8 630.3 451.7 510.5 1031.9 487.4 464.1 454.4 740.8 482.6 451.1 542.9 493.4 853.8 497.9 991.2 487.6 491.9 498.7 485.7 491.6 506.8 474.1 483.5 426.9 467.7 927.3 508.4 449.9 847.8 877.6 517.1 813.4 240.6 1016.4 525.9 485.0 467.8 770.6 824.6 647.5 721.1 1055.3 460.7 906.6 1052.9 777.3 651.3 72.4 43.8 35.0 53.7 25.9 48.0 30.0 25.2 70.3 34.7 28.4 24.5 43.8 36.9 23.2 39.6 31.2 21.9 40.8 38.1 49.6 16.0 40.7 41.1 56.5 45.4 50.7 22.3 58.5 63.3 28.0 46.8 109.0 65.5 32.9 28.3 29.5 35.2 99.0 43.5 34.0 63.8 35.9 41.5 14.1 32.5 27.9 61.9 494.5 504.0 470.8 517.0 1031.9 477.3 473.3 475.6 557.2 465.0 475.9 528.5 484.3 665.2 476.7 991.2 478.7 487.8 484.8 480.2 499.9 484.2 471.3 524.0 474.5 469.3 927.3 463.2 466.5 665.7 729.3 537.4 538.3 262.9 1016.4 528.0 467.9 469.3 581.6 655.7 618.0 533.3 1055.3 483.4 772.3 1052.9 804.3 522.3 2.4 18.2 11.0 9.6 25.9 6.4 5.5 4.9 15.4 8.0 4.7 5.8 4.7 24.1 9.6 39.6 4.8 6.4 8.3 3.0 4.9 3.9 4.8 2.5 3.8 8.2 50.7 2.3 8.1 16.8 10.1 4.0 16.7 1.6 32.9 3.7 3.0 3.1 28.8 13.0 6.0 34.2 35.9 3.2 16.3 32.5 4.8 12.8 101.6 80.0 104.2 101.3 90.4 97.9 102.0 104.7 75.2 96.4 105.5 97.3 98.1 77.9 95.7 95.5 98.2 99.2 97.2 98.9 101.7 95.5 99.4 108.4 111.2 100.3 84.6 91.1 103.7 78.5 83.1 103.9 66.2 109.3 96.4 100.4 96.5 100.3 75.5 79.5 95.4 74.0 93.1 104.9 85.2 98.5 103.5 80.2 U 206Pb/ 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Error 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Best Age Standard Concentration (ppm) 204Pb U/Th 207Pb* Error (%) 235U* Error (%) 238U Error (%) Correlation 238U* Error (Ma) 235U Error (Ma) 207Pb* Error (Ma) (Ma) Error (Ma) (%) TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION 23 of 30 TC3003 050708‐AUNC‐66 050708‐AUNC‐67 050708‐AUNC‐68 050708‐AUNC‐69 050708‐AUNC‐70 050708‐AUNC‐71 050708‐AUNC‐72 050708‐AUNC‐73 050708‐AUNC‐74 050708‐AUNC‐75 050708‐AUNC‐76 050708‐AUNC‐77 050708‐AUNC‐78 050708‐AUNC‐79 050708‐AUNC‐80 050708‐AUNC‐81 050708‐AUNC‐82 050708‐AUNC‐83 050708‐AUNC‐84 050708‐AUNC‐85 050708‐AUNC‐86 050708‐AUNC‐87 050708‐AUNC‐88 050708‐AUNC‐90 050708‐AUNC‐91 050708‐AUNC‐92 050708‐AUNC‐93 050708‐AUNC‐94 050708‐AUNC‐95 050708‐AUNC‐96 050708‐AUNC‐97 050708‐AUNC‐98 050708‐AUNC‐99 050708‐AUNC‐100 SAMPLE 05.07.08‐B.Unc 050708BUNC‐1 050708BUNC‐2 050708BUNC‐4 050708BUNC‐5 050708BUNC‐6 050708BUNC‐8 050708BUNC‐9 050708BUNC‐10 050708BUNC‐11 050708BUNC‐12 050708BUNC‐13 050708BUNC‐14 050708BUNC‐17 Analysis Table A1. (continued) Isotope Ratios Apparent Ages (Ma) 7354 2.5 17.7912 13110 4.0 17.0601 15368 7.2 18.1485 16520 7.1 17.6770 12174 2.5 16.7117 7484 1.7 16.4425 32092 2.9 14.4929 16164 9.7 17.8554 8832 5.6 17.8501 26196 9.7 13.4664 10454 6.7 17.5396 14218 4.4 15.4779 18578 8.4 17.8661 7628 4.7 16.8421 4310 12.5 16.7121 14566 4.4 17.4333 9334 2.5 17.3488 8812 3.0 17.6661 14904 17.2 15.8897 13436 5.5 17.9244 3402 1.8 17.2524 13846 2.9 17.7121 17668 1.6 17.7922 13994 1.5 17.6028 14078 8.7 17.4607 10700 6.0 17.5084 16312 11.6 17.5718 5792 6.0 17.6977 11290 8.4 17.5228 17756 4.0 14.6760 3380 1.4 16.1641 4998 8.8 17.8253 9214 7.0 17.3671 10922 2.2 17.1253 76935 7.0 7.7557 23646 23.7 17.6252 13419 1.8 17.2050 25272 7.7 17.5059 15780 4.5 17.6861 18144 2.5 13.6948 18042 3.3 17.9257 25203 10.2 17.5751 14259 5.3 17.6516 39015 6.3 16.7757 9942 1.7 17.1931 23364 15.3 17.7090 68982 8.7 12.8935 245 267 449 328 239 120 272 262 223 370 277 189 376 264 348 247 239 230 248 284 147 213 242 257 230 311 272 170 255 327 491 180 194 259 280 474 285 426 259 116 386 567 249 425 115 395 511 24 of 30 5.9310 0.5964 0.7082 0.6013 0.6056 1.7010 0.5862 0.5920 0.6011 0.7394 0.6756 0.5882 1.6904 0.5876 0.6770 0.5712 0.6221 0.7943 0.8759 1.2640 0.6198 0.5943 0.9627 0.5993 0.8868 0.5832 0.6761 0.6765 0.6056 0.6853 0.6129 0.7077 0.6041 0.6418 0.6392 0.6468 0.6206 0.6725 0.5941 0.6172 0.6020 0.6239 0.9992 0.6499 0.6158 0.6326 0.6716 2.2 1.1 1.9 1.4 1.7 2.8 3.0 1.8 2.2 2.0 2.4 1.5 1.7 3.6 3.3 3.1 1.8 3.0 5.3 3.4 2.2 2.4 6.8 2.8 3.3 3.4 5.3 3.4 5.2 6.2 2.9 4.2 2.9 6.5 4.1 3.7 2.2 4.5 2.6 4.0 3.3 2.5 3.7 4.9 4.4 2.9 3.0 0.3336 0.0762 0.0884 0.0763 0.0777 0.1690 0.0762 0.0755 0.0770 0.0900 0.0842 0.0756 0.1581 0.0758 0.0838 0.0752 0.0798 0.0963 0.1044 0.1329 0.0803 0.0769 0.0940 0.0762 0.0995 0.0756 0.0826 0.0820 0.0766 0.0862 0.0785 0.0816 0.0785 0.0803 0.0821 0.0835 0.0792 0.0852 0.0754 0.0787 0.0773 0.0793 0.1064 0.0762 0.0796 0.0797 0.0834 1.6 0.8 0.5 0.5 0.8 1.5 1.7 0.6 0.5 1.4 0.7 0.5 1.5 1.1 2.0 0.9 1.3 1.8 4.7 2.3 1.1 0.7 4.4 1.7 1.1 1.2 1.4 0.6 4.7 5.4 1.0 1.2 1.0 1.1 0.5 2.1 1.5 3.5 1.1 2.5 2.1 0.7 2.1 2.3 0.9 0.8 2.2 0.73 0.77 0.27 0.37 0.51 0.52 0.56 0.32 0.23 0.71 0.27 0.32 0.87 0.29 0.61 0.28 0.68 0.62 0.89 0.67 0.51 0.29 0.65 0.60 0.34 0.35 0.26 0.17 0.89 0.87 0.33 0.27 0.34 0.16 0.12 0.57 0.66 0.78 0.43 0.62 0.64 0.28 0.58 0.46 0.20 0.28 0.73 1855.9 473.6 545.9 474.3 482.3 1006.3 473.5 469.0 477.9 555.3 521.4 469.5 946.1 471.2 518.6 467.3 494.7 592.5 640.4 804.2 497.7 477.8 579.3 473.6 611.8 469.6 511.5 508.0 475.6 533.2 487.4 505.4 487.3 497.9 508.7 516.8 491.5 526.9 468.8 488.1 479.8 491.9 651.6 473.3 493.8 494.3 516.5 25.8 3.8 2.6 2.3 3.9 13.6 7.6 2.7 2.3 7.6 3.3 2.3 13.1 4.8 10.1 3.9 6.0 10.4 28.8 17.4 5.3 3.2 24.3 7.7 6.7 5.4 6.7 2.8 21.5 27.4 4.5 5.6 4.6 5.1 2.4 10.4 7.1 17.7 5.1 11.6 9.9 3.3 13.3 10.4 4.2 3.9 11.0 1965.8 475.0 543.7 478.1 480.8 1008.9 468.4 472.2 477.9 562.0 524.1 469.7 1004.9 469.4 525.0 458.8 491.2 593.6 638.7 829.7 489.7 473.6 684.7 476.8 644.6 466.5 524.4 524.7 480.8 530.0 485.4 543.4 479.8 503.4 501.8 506.5 490.2 522.2 473.5 488.1 478.5 492.3 703.4 508.4 487.2 497.7 521.7 18.9 4.1 7.9 5.2 6.3 17.8 11.3 6.9 8.3 8.7 10.0 5.8 10.9 13.5 13.6 11.4 7.1 13.3 25.1 19.5 8.4 9.1 33.8 10.8 15.8 12.8 21.7 14.1 20.1 25.5 11.1 17.6 11.1 26.0 16.3 14.6 8.7 18.2 9.9 15.3 12.8 9.8 18.6 19.7 17.0 11.3 12.3 2083.5 481.3 534.4 496.3 473.7 1014.5 443.8 487.6 478.0 589.4 535.9 470.9 1135.6 460.6 552.9 416.3 474.8 597.7 632.8 898.7 452.6 453.2 1048.5 492.1 761.6 451.2 580.9 597.7 505.5 516.1 476.2 706.0 444.0 528.4 470.5 460.4 484.1 502.0 496.0 488.0 472.2 494.2 872.7 669.5 456.3 513.8 544.5 26.0 15.1 39.3 28.0 31.5 48.1 55.5 38.5 46.8 30.8 51.2 32.4 16.7 76.4 57.2 66.3 29.5 50.5 51.7 52.9 41.4 50.8 104.3 49.9 65.7 71.0 111.0 73.3 51.8 67.4 60.2 85.8 60.8 141.7 90.6 66.5 37.2 60.9 52.4 68.3 56.9 52.9 61.5 93.4 95.5 60.6 44.9 2083.5 473.6 545.9 474.3 482.3 1014.5 473.5 469.0 477.9 555.3 521.4 469.5 1135.6 471.2 518.6 467.3 494.7 592.5 640.4 898.7 497.7 477.8 579.3 473.6 611.8 469.6 511.5 508.0 475.6 533.2 487.4 505.4 487.3 497.9 508.7 516.8 491.5 526.9 468.8 488.1 479.8 491.9 651.6 473.3 493.8 494.3 516.5 26.0 3.8 2.6 2.3 3.9 48.1 7.6 2.7 2.3 7.6 3.3 2.3 16.7 4.8 10.1 3.9 6.0 10.4 28.8 52.9 5.3 3.2 24.3 7.7 6.7 5.4 6.7 2.8 21.5 27.4 4.5 5.6 4.6 5.1 2.4 10.4 7.1 17.7 5.1 11.6 9.9 3.3 13.3 10.4 4.2 3.9 11.0 89.1 98.4 102.2 95.6 101.8 99.2 106.7 96.2 100.0 94.2 97.3 99.7 83.3 102.3 93.8 112.3 104.2 99.1 101.2 89.5 110.0 105.4 55.2 96.2 80.3 104.1 88.1 85.0 94.1 103.3 102.3 71.6 109.8 94.2 108.1 112.2 101.5 105.0 94.5 100.0 101.6 99.5 74.7 70.7 108.2 96.2 94.9 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION 1.5 0.7 1.8 1.3 1.4 2.4 2.5 1.7 2.1 1.4 2.3 1.5 0.8 3.4 2.6 3.0 1.3 2.3 2.4 2.6 1.9 2.3 5.2 2.3 3.1 3.2 5.1 3.4 2.4 3.1 2.7 4.0 2.7 6.5 4.1 3.0 1.7 2.8 2.4 3.1 2.6 2.4 3.0 4.4 4.3 2.8 2.1 U 206Pb/ 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Error 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Best Age Standard Concentration (ppm) 204Pb U/Th 207Pb* Error (%) 235U* Error (%) 238U Error (%) Correlation 238U* Error (Ma) 235U Error (Ma) 207Pb* Error (Ma) (Ma) Error (Ma) (%) TC3003 TC3003 050708BUNC‐18 050708BUNC‐19 050708BUNC‐20 050708BUNC‐21 050708BUNC‐22 050708BUNC‐23 050708BUNC‐24 050708BUNC‐25 050708BUNC‐28 050708BUNC‐29 050708BUNC‐30 050708BUNC‐32 050708BUNC‐33 050708BUNC‐34 050708BUNC‐35 050708BUNC‐36 050708BUNC‐37 050708BUNC‐39 050708BUNC‐40 050708BUNC‐41 050708BUNC‐42 050708BUNC‐43 050708BUNC‐44 050708BUNC‐45 050708BUNC‐47 050708BUNC‐48 050708BUNC‐49 050708BUNC‐50 050708BUNC‐51 050708BUNC‐52 050708BUNC‐53 050708BUNC‐55 050708BUNC‐57 050708BUNC‐58 050708BUNC‐59 050708BUNC‐60 050708BUNC‐61 050708BUNC‐62 050708BUNC‐63 050708BUNC‐64 050708BUNC‐66 050708BUNC‐67 050708BUNC‐68 050708BUNC‐69 050708BUNC‐70 050708BUNC‐71 050708BUNC‐72 050708BUNC‐73 Analysis Table A1. (continued) Isotope Ratios Apparent Ages (Ma) 371 247 358 265 552 461 118 140 669 277 310 352 255 276 278 219 416 148 364 235 655 433 464 307 181 249 372 199 259 365 80 200 223 283 327 398 470 134 171 366 141 347 459 173 330 273 269 516 23505 8385 42780 17937 31956 45312 9879 14154 45591 20115 54408 31059 14148 22923 12243 33819 73827 474 24504 19575 34785 33759 32130 21573 26796 18105 25245 14964 24681 20349 4134 49281 22074 28218 33927 20967 33729 11487 37545 19380 12669 16803 37221 12864 30768 16164 18831 18051 6.9 1.0 2.1 6.4 11.8 3.1 0.9 2.4 13.4 2.5 5.3 37.0 0.9 3.9 3.4 2.1 2.2 1.8 8.9 6.5 8.3 3.2 29.7 5.2 1.4 2.2 2.4 4.4 2.3 7.0 0.8 1.1 5.7 4.1 9.2 1.7 12.2 2.5 1.1 12.4 1.8 2.2 2.0 2.3 4.5 5.7 3.5 1.5 17.7383 17.0144 9.1449 17.7017 17.6818 16.5630 16.4654 17.2091 17.5422 17.3680 13.6416 16.9001 17.0166 17.4886 15.7829 13.1724 13.3536 15.7071 17.5488 15.6843 16.9981 13.9348 17.7250 17.5269 8.5878 16.9366 17.2198 15.3176 13.7677 17.7178 17.2900 8.1528 14.9694 16.4449 16.3294 17.4177 17.5233 17.1414 8.8016 17.3965 16.8880 16.9923 17.1283 16.8862 14.6481 16.3230 16.8751 17.1880 1.4 1.8 1.7 1.8 1.0 1.4 2.5 1.7 1.6 1.5 1.3 1.7 1.1 1.0 7.4 1.5 1.0 164.6 1.6 2.5 2.1 1.6 1.0 1.3 1.7 0.7 1.1 3.9 1.6 1.3 2.6 1.8 5.6 1.3 1.1 1.6 1.6 1.5 2.6 2.4 1.6 2.1 0.8 1.8 1.4 2.3 3.0 3.8 0.5933 0.6357 3.3879 0.6074 0.6035 0.8586 0.8888 0.7114 0.6118 0.6773 1.7016 0.6792 0.6978 0.6217 0.7020 1.9420 1.8865 0.0187 0.6190 0.8490 0.6867 1.3740 0.6059 0.6215 3.8727 0.7120 0.6932 0.8101 1.4802 0.6053 0.6174 6.2483 0.8104 0.8296 0.7782 0.6601 0.6253 0.6766 4.3080 0.6234 0.7615 0.7346 0.7086 0.7274 1.0640 0.7456 0.7187 0.6302 2.8 1.9 2.1 2.0 1.1 1.6 2.8 1.8 1.6 1.7 1.8 2.8 1.6 1.4 7.5 1.6 1.1 164.8 1.6 2.8 3.1 2.3 1.5 1.5 4.8 1.1 1.2 6.6 3.4 1.4 2.7 1.9 9.2 1.4 1.8 1.8 1.8 1.5 8.0 3.2 1.8 2.9 1.0 1.9 3.3 2.4 3.2 3.9 0.0763 0.0784 0.2247 0.0780 0.0774 0.1031 0.1061 0.0888 0.0778 0.0853 0.1684 0.0833 0.0861 0.0789 0.0804 0.1855 0.1827 0.0021 0.0788 0.0966 0.0847 0.1389 0.0779 0.0790 0.2412 0.0875 0.0866 0.0900 0.1478 0.0778 0.0774 0.3695 0.0880 0.0989 0.0922 0.0834 0.0795 0.0841 0.2750 0.0787 0.0933 0.0905 0.0880 0.0891 0.1130 0.0883 0.0880 0.0786 2.5 0.5 1.2 0.9 0.5 0.8 1.3 0.6 0.5 0.7 1.3 2.3 1.1 1.0 1.2 0.5 0.5 8.0 0.5 1.2 2.4 1.7 1.2 0.8 4.5 0.8 0.5 5.4 3.1 0.5 0.7 0.5 7.2 0.5 1.4 0.9 0.8 0.5 7.6 2.2 0.9 2.1 0.6 0.5 3.0 0.5 1.1 0.5 0.88 0.27 0.58 0.43 0.44 0.50 0.47 0.32 0.31 0.43 0.69 0.81 0.70 0.70 0.16 0.31 0.48 0.05 0.33 0.43 0.75 0.72 0.77 0.52 0.93 0.74 0.42 0.81 0.89 0.37 0.27 0.26 0.79 0.36 0.77 0.46 0.45 0.33 0.94 0.66 0.48 0.70 0.57 0.27 0.90 0.21 0.35 0.13 474.2 486.8 1306.7 484.1 480.6 632.8 650.3 548.3 483.2 527.8 1003.0 515.5 532.5 489.3 498.3 1097.1 1081.7 13.7 488.9 594.3 523.9 838.2 483.5 490.2 1393.0 540.5 535.3 555.5 888.6 482.8 480.7 2026.9 543.6 608.2 568.3 516.3 492.9 520.7 1566.1 488.1 574.9 558.7 543.9 550.1 690.4 545.3 543.5 487.6 11.4 2.3 14.2 4.0 2.3 4.8 8.2 3.0 2.3 3.6 11.7 11.3 5.6 4.8 5.6 5.0 5.2 1.1 2.5 6.8 11.8 13.2 5.5 3.7 56.0 4.3 2.6 28.7 25.3 2.3 3.3 8.7 37.7 2.9 7.4 4.2 3.7 2.5 105.3 10.1 4.7 11.0 3.0 2.6 19.8 2.6 5.8 2.4 473.0 499.6 1501.6 481.9 479.5 629.3 645.7 545.5 484.7 525.1 1009.1 526.3 537.5 490.9 540.0 1095.7 1076.3 18.8 489.2 624.1 530.8 877.9 481.0 490.8 1608.0 545.9 534.7 602.5 922.3 480.6 488.2 2011.2 602.7 613.4 584.4 514.7 493.1 524.7 1694.9 492.0 574.9 559.2 543.9 555.0 735.8 565.7 549.9 496.2 10.7 7.4 16.2 7.5 4.4 7.4 13.6 7.5 6.3 6.9 11.6 11.6 6.5 5.6 31.3 10.7 7.2 30.8 6.4 12.8 13.0 13.6 5.9 5.9 38.7 4.7 5.0 30.2 20.8 5.2 10.3 16.5 41.6 6.4 7.9 7.5 6.9 6.3 66.1 12.6 7.8 12.6 4.3 8.0 17.5 10.2 13.4 15.1 467.2 558.7 1788.6 471.7 474.3 617.0 629.8 533.8 491.8 513.7 1022.4 573.4 558.4 498.5 720.3 1092.9 1065.5 730.6 490.9 733.6 560.8 979.2 468.8 493.7 1902.3 568.7 532.5 783.5 1003.7 469.7 523.6 1995.2 831.6 632.5 647.7 507.4 494.1 542.5 1858.0 510.1 575.0 561.5 544.2 575.2 876.7 648.5 576.6 536.6 30.2 39.3 30.6 39.1 22.9 29.4 54.1 36.7 34.4 33.7 26.5 36.4 24.3 22.4 156.6 30.4 19.3 948.5 34.2 52.6 45.6 32.6 21.8 28.4 31.3 16.3 23.7 81.4 32.1 27.8 56.2 32.4 117.1 28.1 24.5 36.1 34.9 31.7 47.5 53.1 34.0 45.4 18.1 39.0 29.4 49.4 64.4 83.6 474.2 486.8 1788.6 484.1 480.6 632.8 650.3 548.3 483.2 527.8 1022.4 515.5 532.5 489.3 498.3 1092.9 1065.5 13.7 488.9 594.3 523.9 979.2 483.5 490.2 1902.3 540.5 535.3 555.5 1003.7 482.8 480.7 1995.2 543.6 608.2 568.3 516.3 492.9 520.7 1858.0 488.1 574.9 558.7 543.9 550.1 690.4 545.3 543.5 487.6 11.4 2.3 30.6 4.0 2.3 4.8 8.2 3.0 2.3 3.6 26.5 11.3 5.6 4.8 5.6 30.4 19.3 1.1 2.5 6.8 11.8 32.6 5.5 3.7 31.3 4.3 2.6 28.7 32.1 2.3 3.3 32.4 37.7 2.9 7.4 4.2 3.7 2.5 47.5 10.1 4.7 11.0 3.0 2.6 19.8 2.6 5.8 2.4 101.5 87.1 73.1 102.6 101.3 102.5 103.2 102.7 98.3 102.7 98.1 89.9 95.4 98.2 69.2 100.4 101.5 1.9 99.6 81.0 93.4 85.6 103.1 99.3 73.2 95.0 100.5 70.9 88.5 102.8 91.8 101.6 65.4 96.2 87.7 101.8 99.8 96.0 84.3 95.7 100.0 99.5 99.9 95.6 78.7 84.1 94.3 90.9 U 206Pb/ 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Error 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Best Age Standard Concentration (ppm) 204Pb U/Th 207Pb* Error (%) 235U* Error (%) 238U Error (%) Correlation 238U* Error (Ma) 235U Error (Ma) 207Pb* Error (Ma) (Ma) Error (Ma) (%) TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION 25 of 30 TC3003 050708BUNC‐74 050708BUNC‐75 050708BUNC‐76 050708BUNC‐77 050708BUNC‐78 050708BUNC‐79 050708BUNC‐80 050708BUNC‐82 050708BUNC‐83 050708BUNC‐84 050708BUNC‐85 050708BUNC‐86 050708BUNC‐87 050708BUNC‐88 050708BUNC‐89 050708BUNC‐90 050708BUNC‐91 050708BUNC‐92 050708BUNC‐93 050708BUNC‐94 050708BUNC‐95 050708BUNC‐97 050708BUNC‐98 050708BUNC‐99 050708BUNC‐100 SAMPLE 04.27.08‐AngB 042708AngB‐1 042708AngB‐2 042708AngB‐3 042708AngB‐4 042708AngB‐5 042708AngB‐7 042708AngB‐9 042708AngB‐11 042708AngB‐12 042708AngB‐13 042708AngB‐14 042708AngB‐17 042708AngB‐19 042708AngB‐20 Analysis Table A1. (continued) Isotope Ratios Apparent Ages (Ma) 17214 336 18984 22521 12054 26226 16080 25293 19248 10821 23430 29439 24180 24306 17742 23694 22773 8958 11916 14928 14994 12810 13593 66840 15225 20496 5986 9066 6376 13638 4840 12774 9692 19520 5768 10998 12830 6424 8162 395 486 192 302 218 483 321 290 200 339 378 242 349 184 187 335 313 125 423 182 252 297 427 166 182 401 122 166 125 356 112 297 206 383 144 230 236 99 185 26 of 30 6.0 2.3 2.1 1.7 3.6 1.9 3.4 2.6 2.1 4.6 5.8 2.8 1.7 3.3 11.7 1.4 1.0 7.6 6.2 4.9 7.7 10.6 2.7 1.3 10.1 1.4 9.1 15.0 6.4 10.2 2.5 1.2 19.3 1.4 4.5 1.7 1.7 1.1 2.3 16.5795 17.3517 17.6271 18.0159 17.9001 18.2966 18.1668 17.7293 17.4357 17.9697 15.5875 17.9512 17.9111 18.0274 17.5322 14.8943 16.1942 17.6298 17.8553 17.5287 17.5970 16.2117 14.5466 17.4906 17.4422 15.9145 17.7149 16.6221 17.2460 17.6640 17.6597 17.0775 16.4528 17.1228 15.6316 17.3148 16.9084 8.5669 17.2265 0.6 1.6 2.4 2.1 1.1 2.5 1.8 1.6 1.1 2.3 2.0 29.3 3.2 2.6 0.8 30.9 1.5 1.8 2.5 2.8 2.5 2.9 1.2 2.0 1.0 2.4 2.6 2.4 1.8 1.5 1.9 1.8 1.1 1.8 4.8 3.8 2.4 1.3 2.1 0.7653 0.6579 0.6529 0.5722 0.5744 0.5270 0.5169 0.5927 0.6660 0.5917 0.7536 0.6192 0.6199 0.6019 0.6216 0.0205 0.9417 0.5972 0.5877 0.5967 0.5971 0.7835 1.2359 0.6102 0.6212 0.9442 0.5862 0.7210 0.6964 0.5901 0.6064 0.7071 0.7474 0.6973 0.7380 0.6203 0.6288 5.4105 0.6797 1.8 1.7 3.0 2.4 1.4 2.7 1.8 2.1 2.1 3.3 2.3 29.3 3.5 2.6 1.7 31.1 1.6 1.9 2.7 3.5 3.2 5.3 1.3 2.5 1.2 2.7 2.7 2.5 1.8 1.6 2.0 1.9 1.3 1.8 5.2 3.8 2.5 1.5 2.2 0.0920 0.0828 0.0835 0.0748 0.0746 0.0699 0.0681 0.0762 0.0842 0.0771 0.0852 0.0806 0.0805 0.0787 0.0790 0.0022 0.1106 0.0764 0.0761 0.0759 0.0762 0.0921 0.1304 0.0774 0.0786 0.1090 0.0753 0.0869 0.0871 0.0756 0.0777 0.0876 0.0892 0.0866 0.0837 0.0779 0.0771 0.3362 0.0849 1.7 0.6 1.8 1.2 0.9 0.8 0.5 1.4 1.8 2.4 1.2 0.6 1.3 0.6 1.5 3.7 0.7 0.5 0.9 2.1 2.0 4.5 0.5 1.5 0.5 1.3 0.9 0.8 0.5 0.5 0.5 0.6 0.6 0.5 2.0 0.7 0.6 0.8 0.5 0.94 0.37 0.61 0.51 0.66 0.30 0.27 0.65 0.84 0.73 0.53 0.02 0.38 0.21 0.87 0.12 0.41 0.27 0.34 0.61 0.64 0.84 0.38 0.60 0.44 0.47 0.31 0.30 0.27 0.32 0.25 0.32 0.49 0.29 0.37 0.18 0.23 0.55 0.24 567.5 512.8 516.8 464.8 463.6 435.8 424.7 473.5 521.3 478.9 527.0 499.8 499.3 488.3 490.4 14.2 676.3 474.4 472.9 471.3 473.5 568.0 790.1 480.6 487.7 666.9 468.1 537.3 538.4 469.8 482.2 541.2 550.7 535.3 518.0 483.6 478.8 1868.2 525.4 9.2 3.2 8.9 5.4 4.2 3.4 2.1 6.3 8.9 11.2 6.2 2.9 6.4 2.6 6.9 0.5 4.4 2.3 4.2 9.7 9.2 24.5 3.7 7.0 2.4 8.0 3.8 3.9 2.6 2.4 2.3 3.2 3.3 2.8 9.7 3.3 2.6 13.5 2.6 577.1 513.3 510.3 459.4 460.9 429.8 423.1 472.6 518.3 472.0 570.3 489.4 489.8 478.4 490.9 20.6 673.8 475.5 469.4 475.1 475.4 587.5 817.0 483.7 490.6 675.1 468.4 551.2 536.6 470.9 481.3 543.0 566.7 537.1 561.2 490.1 495.3 1886.5 526.6 7.9 7.0 11.9 8.8 5.3 9.3 6.4 8.1 8.5 12.5 10.1 114.4 13.4 10.0 6.5 6.3 8.1 7.1 10.1 13.2 12.1 23.8 7.4 9.7 4.5 13.3 10.2 10.6 7.6 6.1 7.6 8.1 5.5 7.7 22.5 15.0 9.7 12.9 9.0 614.9 515.8 481.1 432.7 447.0 398.1 414.0 468.3 505.1 438.4 746.7 440.7 445.6 431.2 493.0 842.1 665.5 480.7 452.6 493.4 484.9 663.1 891.1 498.2 504.4 702.7 470.1 609.4 529.2 476.5 477.0 550.7 631.5 544.9 740.7 520.4 572.3 1906.7 531.7 13.4 35.5 52.3 46.0 23.9 57.0 39.6 35.9 25.1 50.4 41.3 665.6 71.0 57.4 17.9 658.2 32.0 39.6 56.1 60.7 54.3 61.5 25.4 44.6 22.8 50.7 57.0 51.3 38.5 33.7 42.5 39.9 23.5 38.6 102.5 83.0 52.4 22.6 46.8 567.5 512.8 516.8 464.8 463.6 435.8 424.7 473.5 521.3 478.9 527.0 499.8 499.3 488.3 490.4 14.2 676.3 474.4 472.9 471.3 473.5 568.0 790.1 480.6 487.7 666.9 468.1 537.3 538.4 469.8 482.2 541.2 550.7 535.3 518.0 483.6 478.8 1906.7 525.4 9.2 3.2 8.9 5.4 4.2 3.4 2.1 6.3 8.9 11.2 6.2 2.9 6.4 2.6 6.9 0.5 4.4 2.3 4.2 9.7 9.2 24.5 3.7 7.0 2.4 8.0 3.8 3.9 2.6 2.4 2.3 3.2 3.3 2.8 9.7 3.3 2.6 22.6 2.6 92.3 99.4 107.4 107.4 103.7 109.5 102.6 101.1 103.2 109.2 70.6 113.4 112.0 113.2 99.5 1.7 101.6 98.7 104.5 95.5 97.6 85.7 88.7 96.5 96.7 94.9 99.6 88.2 101.7 98.6 101.1 98.3 87.2 98.3 69.9 92.9 83.7 98.0 98.8 U 206Pb/ 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Error 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Best Age Standard Concentration (%) (ppm) 204Pb U/Th 207Pb* Error (%) 235U* Error (%) 238U Error (%) Correlation 238U* Error (Ma) 235U Error (Ma) 207Pb* Error (Ma) (Ma) Error (Ma) TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION 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 property of apatite that varies as a function of [eU], especially in slowly cooled rocks [Shuster et al., 2006; Flowers et al., 2009]. Because slow cooling allows more time for He diffusion, it enhances the effect of radiation damage, and thus the presence of a large spread in AHe dates that correlate to [eU] within a single sample is an indication that the sample spent and extended period of time in the HePRZ [Flowers et al., 2007]. During radiogenic decay, emitted a particles can travel up to 20 mm through the apatite crystal lattice, and some are ejected from crystal altogether. For grains <200 mm in radius, this leads to significant He loss and anomalously young ages [Farley, 2002; Ehlers and Farley, 2003]. 27 of 30 TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION [71] Because a ejection is largely dependent on the size and surface to volume ratio of the crystal [Reiners and Farley, 2001], the dimensions and morphology of an apatite grain are routinely used with the Ft correction method to calculate the amount of 4He lost by a ejection and correct the AHe age [Farley et al., 1996]. In Figure A1 we present the complete time‐temperature paths of six AHe samples (see Figure 10) modeled using HeFTy [Ketcham, 2005], the radiation trapping model of Shuster et al. [2006] (Do/a2), and the static alpha ejection correction of Farley et al. [1996]. [72] Acknowledgments. This project was made possible through generous support from NSF EAR 0710724. The KU (U‐Th)/He laboratory acknowledges support from NSF EAR 0414817. The Arizona LaserChron Center is supported by NSF EAR 0443387 and NSF EAR 0732436. We thank Peter DeCelles, George Gehrels, and Lindsay Schoenbohm for scientific inputs and Chris Krugh, Roman Kislitsyn, and Jamey Stutz for technical and field support. References Abascal, L. V. (2005), Combined thin‐skinned and thick‐skinned deformation in the central Andean foreland of northwestern Argentina, J. South Am. Earth Sci., 19, 75–81, doi:10.1016/j.jsames.2005.01.004. Acocella, V., L. Vezzoli, R. Omarini, M. Matteini, and R. Mazzuoli (2007), Kinematic variations across the Eastern Cordillera at 24°S (central Andes): Tectonic and magmatic implications, Tectonophysics, 434, 81– 92, doi:10.1016/j.tecto.2007.02.001. Allmendinger, R. W., and T. Gubbels (1996), Pure and simple shear plateau uplift, Altiplano‐Puna, Argentina and Bolivia, Tectonophysics, 259, 1–13, doi:10.1016/0040-1951(96)00024-8. Allmendinger, R. W., V. A. Ramos, T. E. Jordan, M. Palma, and B. L. Isacks (1983), Paleogeography and Andean structural geometry, northwest Argentina, Tectonics, 2, 1–16, doi:10.1029/TC002i001p00001. Arriagada, C., P. R. Cobbold, and P. Roperch (2006), Salar de Atacama basin: A record of compressional tectonics in the central Andes since the mid‐Cretaceous, Tectonics, 25, TC1008, doi:10.1029/ 2004TC001770. Bally, A. W., P. L. Gordy, and G. A. Stewart (1966), Structure, seismic data and orogenic evolution of southern Canadian Rocky Mountains, Bull. Can. Pet. Geol., 14, 337–381. Barnes, J. B., and T. A. Ehlers (2009), End member models for Andean Plateau uplift, Earth Sci. Rev., 97, 105–132, doi:10.1016/j.earscirev. 2009.08.003. Barnes, J. B., T. A. Ehlers, N. McQuarrie, P. B. O’Sullivan, and S. Tawackoli (2008), Thermochronometer record of Central Andean Plateau growth, Bolivia (19.5°S), Tectonics, 27, TC3003, doi:10.1029/2007TC002174. Biswas, S., I. Coutand, D. Grujic, C. Hager, D. Stockli, and B. Grasemann (2007), Exhumation and uplift of the Shillong Plateau and its influence on the eastern Himalayas: New constraints from apatite and zircon (U‐ Th‐[Sm])/He and apatite fission track analyses, Tectonics, 26, TC6013, doi:10.1029/2007TC002125. Bosio, P. P., J. Powell, C. del Papa, and F. Hongn (2009), Middle Eocene deformation– sedimentation in the Luracatao Valley: Tracking the beginning of the foreland basin of northwestern Argentina, J. South Am. Earth Sci., 28, 142–154, doi:10.1016/j.jsames.2009.06.002. Bywater‐Reyes, S., B. Carrapa, M. Clementz, and L. Schoenbohm (2010), The effect of late Cenozoic aridification on sedimentation in the Eastern Cordillera of NW Argentina (Angastaco basin), Geology, 38, 235–238, doi:10.1130/G30532.1. Cahill, T. A., and B. L. Isacks (1992), Seismicity and shape of the subducted Nazca Plate, J. Geophys. Res., 97, 17,503–17,529, doi:10.1029/ 92JB00493. Carrapa, B., and P. G. DeCelles (2008), Eocene exhumation and basin development in the Puna of northwestern Argentina, Tectonics, 27, TC1015, doi:10.1029/2007TC002127. Carrapa, B., D. Adelmann, G. E. Hilley, E. Mortimer, E. R. Sobel, and M. R. Strecker (2005), Oligocene uplift and development of plateau morphology in the southern central Andes, Tectonics, 24, TC4011, doi:10.1029/2004TC001762. Carrapa, B., E. Sobel, and M. R. Strecker (2006), Cenozoic orogenic growth in the central Andes: Evidence from rock provenance and apatite fission track thermochronology along the southernmost Puna Plateau margin (NW Argentina), Earth Planet. Sci. Lett., 247, 82–100, doi:10.1016/j.epsl.2006.04.010. TC3003 Carrapa, B., B. Hauer, L. Schoenbohm, M. R. Strecker, A. K. Schmitt, A. Villanueva, and A. S. Gomez (2008a), Dynamics of deformation and sedimentation in the northern Sierras Pampeanas: An integrated study of the Neogene Fiambala basin, NW Argentina, Geol. Soc. Am. Bull., 120, 1518–1543, doi:10.1130/B26111.1. Carrapa, B., L. Schoenbohm, M. Clementz, S. Bywater, and J. Quade (2008b), Oxygen isotope evidence from Cenozoic paleosol carbonates on the Puna Plateau of NW Argentina: Low or dry in the Neogene?, Eos Trans. AGU, 89(53), Fall Meet. Suppl., Abstract T42A–03. Carrapa, B., L. Schoenbhom, P. G. DeCelles, M. Clementz, and K. Hungtington (2009), Surface response to lithospheric delamination: An example from the Puna Plateau of NW Argentina, Geol. Soc. Am. Abstr. Programs, 41, 516. Carrera, N., and J. A. Muñoz (2008), Thrusting evolution in the southern Cordillera Oriental (northern Argentine Andes): Constraints from growth strata, Tectonophysics, 459, 107–122, doi:10.1016/j.tecto.2007.11.068. Carrera, N., J. A. Muñoz, F. Sàbat, R. Mon, and E. Roca (2006), The role of inversion tectonics in the structure of the Cordillera Oriental (NW Argentinean Andes), J. Struct. Geol., 28, 1921–1932, doi:10.1016/j.jsg. 2006.07.006. Coutand, I., P. R. Cobbold, M. de Urreiztieta, P. Gautier, A. Chauvin, D. Gapais, E. A. Rossello, and O. Lòpez‐Gamundí (2001), Style and history of Andean deformation, Puna Plateau, northwestern Argentina, Tectonics, 20, 210–234. Coutand, I., B. Carrapa, A. Deeken, A. K. Schmitt, E. Sobel, and M. R. Strecker (2006), Orogenic plateau formation and lateral growth of compressional basins and ranges: Insights from sandstone petrography and detrital apatite fission‐track thermochronology in the Angastaco Basin, NW Argentina, Basin Res., 18, 1–26, doi:10.1111/j.1365-2117. 2006.00283.x. Cristallini, E., A. H. Cominguez, and V. A. Ramos (1997), Deep structure of the Metan‐Guachipas region: Tectonic inversion in northwestern Argentina, J. South Am. Earth Sci., 10, 403–421, doi:10.1016/S08959811(97)00026-6. Cristallini, E. O., A. H. Cominguez, V. A. Ramos, and E. D. Mercerat (2004), Basement double1136 wedge thrusting in the northern Sierras Pampeanas of Argentina (27 degrees S); constraints from deep seismic reflection, in Thrust Tectonics and Hydrocarbon Systems, edited by K. R. McClay, AAPG Mem., 82, 65–90. Dahlstrom, C. D. A. (1970), Structural geology in the eastern margin of the Canadian Rocky Mountains, Bull. Can. Pet. Geol., 18, 332–406. Davis, D., J. Suppe, and F. A. Dahlen (1983), Mechanics of fold‐and‐thrust belts and accretionary wedges, J. Geophys. Res., 88, 1153–1172, doi:10.1029/JB088iB02p01153. DeCelles, P. G., and P. C. DeCelles (2001), Rates of shortening, propagation, underthrusting, and flexural wave migration in continental orogenic systems, Geology, 29, 135–138, doi:10.1130/0091-7613(2001) 029<0135:ROSPUA>2.0.CO;2. DeCelles, P. G., and B. K. Horton (2003), Early to middle Tertiary foreland basin development and the history of Andean crustal shortening in Bolivia, Geol. Soc. Am. Bull., 115, 58–77, doi:10.1130/0016-7606(2003) 115<0058:ETMTFB>2.0.CO;2. DeCelles, P. G., B. Carrapa, and G. E. Gehrels (2007), Detrital zircon U‐Pb ages provide provenance and chronostratigraphic information from Eocene synorogenic deposits in northwestern Argentina, Geology, 35, 323–326, doi:10.1130/G23322A.1. DeCelles, P. G., B. Carrapa, B. Horton, and D. Starck (2008), Foreland basin evolution in NW Argentina and implications for timing of Andean orogenesis, paper presented at XVII Argentinean Geological Conference, Int. Union of Geol. Sci., Jujuy, Argentina. DeCelles, P. G., M. N. Ducea, P. Kapp, and G. Zandt (2009), Cyclicity in Cordilleran orogenic systems, Nat. Geosci., 2, 251–257, doi:10.1038/ ngeo469. Deeken, A., E. R. Sobel, I. Coutand, M. Haschke, U. Riller, and M. R. Strecker (2006), Development of the southern Eastern Cordillera, NW Argentina, constrained by apatite fission track thermochronology: From early Cretaceous extension to middle Miocene shortening, Tectonics, 25, TC6003, doi:10.1029/2005TC001894. Dickinson, W. R., and G. E. Gehrels (2009), Use of U‐Pb ages of detrital zircons to infer maximum depositional ages of strata: A test against a Colorado Plateau Mesozoic database, Earth Planet. Sci. Lett., 288, 115–125, doi:10.1016/j.epsl.2009.09.013. Echavarria, R., R. Hernandez, R. W. Allmendinger, and J. H. Reynolds (2003), Sub‐Andean thrust and fold belt of northwest Argentina: Geometry and timing of the Andean evolution, AAPG Bull., 87, 965–985, doi:10.1306/01200300196. Ege, H., E. R. Sobel, E. Scheuber, and V. Jacobshagen (2007), Exhumation history of the southern Altiplano plateau (southern Bolivia) constrained 28 of 30 TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION by apatite fission track thermochronology, Tectonics, 26, TC1004, doi:10.1029/2005TC001869. Ehlers, T. A., and K. A. Farley (2003), Apatite 1169(U‐Th)/He thermochronometry: Methods and applications to problems in tectonic and surface processes, Earth Planet. Sci. Lett., 206, 1–14, doi:10.1016/ S0012-821X(02)01069-5. Ehlers, T. A., and C. J. Poulsen (2009), Large paleoclimate influence the intepretation of Andean Plateau paleoaltimetry, Earth Planet. Sci. Lett., 281, 238–248. Elger, K., O. Oncken, and J. Glodny (2005), Plateau‐style accumulation of deformation: Southern Altiplano, Tectonics, 24, TC4020, doi:10.1029/ 2004TC001675. Farley, K. A. (2002), (U‐Th)/He dating: Techniques, calibrations, and applications, in Noble Gases in Geochemistry and Cosmochemistry, edited by D. Porcelli et al., pp. 819–844, Mineral. Soc. of Am., Washington, D. C. Farley, K. A., R. A. Wolf, and L. T. Silver (1996), The effects of long alpha stopping distances on (U‐Th)/He ages, Geochim. Cosmochim. Acta, 60, 4223–4229, doi:10.1016/S0016-7037(96)00193-7. Flowers, R. M., D. L. Shuster, B. P. Wernicke, and K. A. Farley (2007), Radiation damage control on apatite (U‐Th)/He dates from the Grand Canyon region, Colorado Plateau, Geology, 35, 447–450, doi:10.1130/ G23471A.1. Flowers, R. M., R. A. Ketcham, D. L. Shuster, and K. A. Farley (2009), Apatite (U‐Th)/He thermochronometry using a radiation damage accumulation and annealing model, Geochim. Cosmochim. Acta, 73, 2347– 2365, doi:10.1016/j.gca.2009.01.015. Garzione, C. N., P. Molnar, J. C. Libarkin, and B. J. MacFadden (2006), Rapid late Miocene rise of the Bolivian Altiplano: Evidence for removal of mantle lithosphere, Earth Planet. Sci. Lett., 241, 543–556, doi:10.1016/j.epsl.2005.11.026. Gehrels, G. E., V. A. Valencia, and J. Ruiz (2008), Enhanced precision, accuracy, efficiency, and spatial resolution of U‐Pb ages by laser ablation multicollector inductively coupled plasma mass spectrometry, Geochem. Geophys. Geosyst., 9, Q03017, doi:10.1029/2007GC001805. Gillis, R. J., B. K. Horton, and M. Grove (2006), Thermochronology, geochronology, and upper crustal structure of the Cordillera Real: Implications for Cenozoic exhumation of the central Andean plateau, Tectonics, 25, TC6007, doi:10.1029/2005TC001887. Gray, M. B., and P. K. Zeitler (1997), Comparison of clastic wedge provenance in the Appalachian foreland using U/Pb ages of detrital zircons, Tectonics, 16, 151–160, doi:10.1029/96TC02911. Grier, M. E., and R. D. Dallmeyer (1990), Age of the Payogastilla Group: Implication for foreland basin development, NW Argentina, J. South Am. Earth Sci., 3, 269–278, doi:10.1016/0895-9811(90)90008-O. Grier, M. E., J. A. Salfity, and R. W. Allmendinger (1991), Andean reactivation of the Cretaceous Salta rift, northwestern Argentina, J. South Am. Earth Sci., 4, 351–372, doi:10.1016/0895-9811(91)90007-8. Gubbels, T. L., B. L. Isacks, and E. Farrar (1993), High‐level surfaces, plateau uplift, and foreland development, Bolivian central Andes, Geology, 21, 695–698, doi:10.1130/0091-7613(1993)021<0695:HLSPUA>2.3. CO;2. Hongn, F. D., and R. E. Seggiaro (2001), Hoja Geologica Cachi, 2566–III, Provincias de Salta y Catamarca, Republica Argentina, map, Bull. 248, 87 pp., Inst. de Geol. y Recursos Miner., Serv. de Geol. y Min. de Argent., Buenos Aires. Hongn, F., C. del Papa, J. Powell, I. Petrinovic, R. Mon, and V. Deraco (2007), Middle Eocene deformation and sedimentation in the Puna‐Eastern Cordillera transition (23–26°S): Control by preexisting heterogeneities on the pattern of initial Andean shortening, Geology, 35, 271–274, doi:10.1130/G23189A.1. Isacks, B. L. (1988), Uplift of the Central Andean Plateau and bending of the Bolivian Orocline, J. Geophys. Res., 93, 3211–3231, doi:10.1029/ JB093iB04p03211. Jezek, P., and H. Miller (1985), Deposition and facies distribution of turbiditic sediments of the Puncoviscana Formation (upper Precambrian‐ Lower Cambrian) within the basement of the NW Argentine Andes, Zentralbl. Geol. Palaeontol., Teil I, 9–10, 1235–1244. Jordan, T. E. (1995), Retroarc foreland and related basins, in Tectonics of Sedimentary Basins, edited by C. J. Busby and R. V. Ingersoll, pp. 331– 362, Blackwell Sci., Oxford, U. K. Jordan, T. E., and R. W. Allmendinger (1986), The Sierras Pampeanas of Argentina: A modern analogue of Rocky Mountain foreland deformation, Am. J. Sci., 286, 737–764, doi:10.2475/ajs.286.10.737. Jordan, T. E., and R. N. Alonso (1987), Cenozoic stratigraphy and basin tectonics of the Andes Mountains, 20°–28° South latitude, AAPG Bull., 71, 49–64. Jordan, T. E., and C. Mpodozis (2006), Estratigrafıa y evolucion tectonica de la cuenca Paleogena Arizaro‐Pocitos, Puna Occidental (24–25), paper TC3003 presented at XI Congreso Geologico Chileno, Dep. de Cienc. Geol., Univ. Cat. del Norte, Antofagasta, Chile. Kay, S. M., and B. L. Coira (2009), Shallowing and steepening subduction zones, continental lithospheric loss, magmatism, and crustal flow under the Central Andean Altiplano‐Puna Plateau, Mem. Geol. Soc. Am., 204, 229–259. Kay, S. M., C. Mpodozis, and B. Coira (1999), Magmatism, tectonism, and mineral deposits of the central Andes (22–33°S), in Geology and Ore Deposits of the Central Andes, edited by B. Skinner, Spec. Publ. Soc. Econ. Geol., 7, 27–59. Ketcham, R. A. (2005), Forward and inverse modeling of low‐temperature thermochronometry data, in Reviews in Mineralogy and Geochemistry, vol. 58, edited by P. W. Reiners and T. A. Ehlers, pp. 275–314, Mineral. Soc. of Am., Washington, D. C. Kley, J. (1999), Geologic and geometric constraints on a kinematic model of the Bolivian orocline, J. South Am. Earth Sci., 12, 221–235, doi:10.1016/S0895-9811(99)00015-2. Kley, J., and C. R. Monaldi (1998), Tectonic shortening and crustal thickness in the central Andes: How good is the correlation?, Geology, 26, 723–726. Kley, J., and C. R. Monaldi (2002), Tectonic inversion in the Santa Barbara System of the central Andean foreland thrust belt, northwestern Argentina, Tectonics, 21(6), 1061, doi:10.1029/2002TC902003. Kley, J., J. Müller, S. Tawackoli, V. Jacobshagen, and E. Manutsoglu (1997), Pre‐Andean and Andean age deformation in the eastern Cordillera of southern Bolivia, J. South Am. Earth Sci., 10, 1–19, doi:10.1016/ S0895-9811(97)00001-1. Kley, J., C. R. Monaldi, and J. A. Salfity (1999), Along‐strike segmentation of the Andean foreland: Causes and consequences, Tectonophysics, 301, 75–94, doi:10.1016/S0040-1951(98)90223-2. Kraemer, B., D. Adelmann, M. Alten, W. Schnurr, K. Erpenstein, E. Kiefer, P. van den Bogaard, and K. Görler (1999), Incorporation of the Paleogene foreland into the Neogene Puna Plateau: The Salar de Antofalla area, NW Argentina, J. South Am. Earth Sci., 12, 157–182, doi:10.1016/S08959811(99)00012-7. Ludwig, K. R. (2008), Isoplot 3.6, Spec. Publ., 4, 77 pp., Berkeley Geochronol. Cent., Berkeley, Calif. Marquillas, R. A., C. del Papa, and I. F. Sabino (2005), Sedimentary aspects and paleoenvironmental evolution of a rift basin: Salta Group (Cretaceous‐Paleogene), northwestern Argentina, Int. J. Earth Sci., 94(1), 94–113, doi:10.1007/s00531-004-0443-2. Marrett, R. A., R. W. Allmendinger, R. N. Alonso, and R. E. Drake (1994), Late Cenozoic tectonic evolution of the Puna Plateau and adjacent foreland, northwestern Argentine Andes, J. South Am. Earth Sci., 7, 179–207, doi:10.1016/0895-9811(94)90007-8. McQuarrie, N. (2002), The kinematic history of the central Andean fold‐ thrust belt, Bolivia: Implications for building a high plateau, Geol. Soc. Am. Bull., 114, 950–963, doi:10.1130/0016-7606(2002)114<0950: TKHOTC>2.0.CO;2. McQuarrie, N., B. K. Horton, G. Zandt, S. Beck, and P. G. DeCelles (2005), Lithospheric evolution of the Andean fold‐thrust belt, Bolivia, and the origin of the Central Andean Plateau, Tectonophysics, 399, 15–37, doi:10.1016/j.tecto.2004.12.013. McQuarrie, N., J. B. Barnes, and T. A. Ehlers (2008), Geometric, kinematic, and erosional history of the Central Andean Plateau, Bolivia (15–17°S), Tectonics, 27, TC3007, doi:10.1029/2006TC002054. Mon, R., and G. Drozdzewski (1999), Cinturones doble vergentes en los Andes del norte argentine: Hipótesis sobre su origen, Asoc. Geol. Argent. Rev., 54(1), 3–8. Mon, R., and J. A. Salfity (1995), Tectonic evolution of the Andes of northern Argentina, in Petroleum Basins of South America, edited by A. J. Tankard et al., AAPG Mem., 62, 269–283. Mortimer, E., B. Carrapa, C. Isabelle, L. Schoenbohm, E. R. Sobel, J. S. Gomez, and M. R. Strecker (2007), Fragmentation of a foreland basin in response to out‐of‐sequence basement uplifts and structural reactivation: El Cajón‐Campo del Arenal basin, NW Argentina, Geol. Soc. Am. Bull., 119, 637–653, doi:10.1130/B25884.1. Mpodozis, C., C. Arriagada, M. Basso, P. Roperch, P. Cobbold, and M. Reich (2005), Late Mesozoic to Paleogene stratigraphy of the Salar de Atacama Basin, Antofagasta, Northern Chile: Implications for the tectonics evolution of the central Andes, Tectonophysics, 399, 125–154, doi:10.1016/j.tecto.2004.12.019. Norabuena, E., L. Leffler‐Griffin, A. Mao, T. Dixon, S. Stein, I. S. Sacks, L. Ocola, and M. Ellis (1998), Space geodetic observations of Nazca‐ South America convergence across the central Andes, Science, 279, 358–362, doi:10.1126/science.279.5349.358. Oncken, O., D. Hindle, J. Kley, K. Elger, P. Victor, and K. Schemmann (2006), Deformation of the central Andean upper plate system—Facts, 29 of 30 TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION fiction, and constraints for plateau models, in The Andes: Active Subduction Orogeny, edited by O. Oncken et al., pp. 3–27, Springer, Berlin. Ramos, V. (2002), The Pampeanas flat slab of the central Andes, J. South Am. Earth Sci., 15, 59–78, doi:10.1016/S0895-9811(02)00006-8. Reiners, P. W., and M. K. Brandon (2006), Using thermochronology to understand orogenic erosion, Annu. Rev. Earth Planet. Sci., 34, 419–466, doi:10.1146/annurev.earth.34.031405.125202. Reiners, P. W., and K. A. Farley (2001), Influence of crystal size on apatite (U‐Th)/He thermochronology: An example from the Bighorn Mountains, Wyoming, Earth Planet. Sci. Lett., 188, 413–420, doi:10.1016/S0012821X(01)00341-7. Reynolds, J., C. Galli, R. Hernandez, B. Idleman, J. Kotila, R. Hilliard, and C. Naeser (2000), Middle Miocene tectonic development of the Transition Zone, Salta Province, northwest Argentina: Magnetic stratigraphy from the Metan Subgroup, Sierra de Gonzalez, Geol. Soc. Am. Bull., 112, 1736–1751, doi:10.1130/0016-7606(2000)112<1736:MMTDOT>2.0. CO;2. Riller, U., I. Petrinovic, J. Ramelow, M. R. Strecker, and O. Oncken (2001), Late Cenozoic tectonism, collapse caldera and plateau formation in the central Andes, Earth Planet. Sci. Lett., 188, 299–311, doi:10.1016/ S0012-821X(01)00333-8. Ruppel, C., and K. V. Hodges (1994), Pressure‐temperature‐time paths from two‐dimensional thermal models: Prograde, retrograde, and inverted metamorphism, Tectonics, 13, 17–44. Salfity, J. A., and R. A. Marquillas (1994), Tectonic and sedimentary evolution of the Cretaceous‐Eocene Salta Group Basin, Argentina, in Cretaceous Tectonics of the Andes, edited by J. A. Salfity, pp. 266–315, Vieweg, Brunswick, Germany. Salfity, J. A., E. F. Gallardo, J. E. Sastre, and J. Esteban (2004), El lago cuaternario de Angastaco, Valle Calchaqui, Salta, Asoc. Geol. Argent. Rev., 59(2), 312–316. Schmidt, C. J., R. A. Astini, C. H. Costa, C. E. Gardini, and P. E. Kraemer (1995), Cretaceous rifting, alluvial fan sedimentation, and Neogene inversion, southern Sierras Pampeanas, in Petroleum Basins of South America, edited by A. J. Tankard et al., pp. 341–358, Am. Assoc. Pet. Geol., Tulsa, Okla. Schoenbohm, L., and M. Strecker (2009), Normal faulting along the southern margin of the Puna Plateau, northwest Argentina, Tectonics, 28, TC5008, doi:10.1029/2008TC002341. Shuster, D. L., and K. A. Farley (2009), The influence of artificial radiation damage and thermal annealing on helium diffusion kinetics in apatite, Geochim. Cosmochim. Acta, 73, 183–196, doi:10.1016/j.gca.2008. 10.013. Shuster, D. L., R. M. Flowers, and K. A. Farley (2006), The influence of natural radiation damage on helium diffusion kinetics in apatite, Earth Planet. Sci. Lett., 249, 148–161, doi:10.1016/j.epsl.2006.07.028. Sobel, E. R., G. E. Hilley, and M. R. Strecker (2003), Formation of internally drained contractional basins by aridity‐limited bedrock incision, J. Geophys. Res., 108(B7), 2344, doi:10.1029/2002JB001883. Stacey, J., and J. Kramers (1975), Approximation of terrestrial lead isotope evolution by a two‐stage model, Earth Planet. Sci. Lett., 26, 207–221, doi:10.1016/0012-821X(75)90088-6. Starck, D., and L. M. Anzótegui (2001) The late Miocene climatic change: Persistence of a climatic signal through the orogenic stratigraphic record 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- TC3003 sented at XIII Congreso Geológico Argentino y III Congreso de Exploración de Hidrocarburos, Actas I, pp. 433–452, Asoc. Geol. Argent., Buenos Aires. Stockli, D. F., K. A. Farley, and T. A. Dumitru (2000), Calibration of the (U‐Th)/He thermochronometer on an exhumed fault block, White Mountains, Calif. Geol., 28, 983–986. Strecker, M. R., P. Cerveny, L. Arthur, and D. Malizia (1989), Late Cenozoic tectonism and landscape development in the foreland of the Andes: Northern Sierras Pampeanas (26°–28°S), Argentina, Tectonics, 8, 517–534, doi:10.1029/TC008i003p00517. Strecker, M. R., R. N. Alonso, B. Bookhagen, B. Carrapa, G. E. Hilley, E. R. Sobel, and M. H. Trauth (2007), Tectonics and climate of the southern central Andes, Annu. Rev. Earth Planet. Sci., 35, 747–787, doi:10.1146/annurev.earth.35.031306.140158. Strecker, M. R., R. Alonso, B. Bookhagen, B. Carrapa, I. Coutand, M. P. Hain, G. E. Hilley, E. Mortimer, L. Schoenbohm, and E. R. Sobel (2009), Does the topographic distribution of the central Andean Puna Plateau result from climatic or geodynamic processes?, Geology, 37, 643–646, doi:10.1130/G25545A.1. Uba, C. E., J. Kley, M. R. Strecker, and A. K. Schmitt (2009), Unsteady evolution of the Bolivian sub‐Andean thrust belt: The role of enhanced erosion and clastic wedge progradation, Earth Planet. Sci. Lett., 281, 134–146, doi:10.1016/j.epsl.2009.02.010. Uliana, M. A., K. T. Biddle, and J. Cerdan (1989), Mesozoic extension and the formation of Argentine sedimentary basins, in Extensional Tectonics and Stratigraphy of the North Atlantic Margins, edited by A. J. Tankard and H. R. Balkwill, AAPG Mem., 46, 599–614. Voss, R. (2002), Cenozoic stratigraphy of the southern Salar de Antofalla region, northwestern Argentina, Rev. Geol. Chile, 29(2), 151–165, doi:10.4067/S0716-2082002000200002. Willner, A. P., and H. Miller (1985), Structural division and evolution of the lower Paleozoic basement in the NW Argentine Andes, Zentralbl. Geol. Palaeontol., Teil I, 9–10, 1245–1255. Willner, A. P., U. S. Lottner, and H. Miller (1987), Early Paleozoic structural development in the NW Argentine basement of the Andes and its implication for geodynamic reconstructions, in Gondwana Six: Structure, Tectonics and Geophysics, Geophys. Monogr. Ser., vol. 40, edited by G. D. Mckenzie, pp. 229–239, AGU, Washington, D. C. Wolf, R. A., K. A. Farley, and L. T. Silver (1996), Helium diffusion and low‐temperature thermochronometry of apatite, Geochim. Cosmochim. Acta, 60, 4231–4240, doi:10.1016/S0016-7037(96)00192-5. Wolf, R. A., K. A. Farley, and D. M. Kass (1998), Modeling of the temperature sensitivity of the apatite (U‐Th)/He thermochronometer, Chem. Geol., 148, 105–114, doi:10.1016/S0009-2541(98)00024-2. Zeitler, P. K., A. L. Herczig, I. McDougall, and M. Honda (1987), U‐Th‐He dating of apatite: A potential thermochronometer, Geochim. Cosmochim. Acta, 51, 2865–2868, doi:10.1016/0016-7037(87)90164-5. B. Carrapa, Department of Geosciences, University of Arizona, 1040 E. 4th St., Tucson, AZ 85721, USA. D. F. Stockli, Department of Geology, Kansas University, 1475 Jayhawk Blvd., Rm. 120, Lawrence, KS 66045, USA. J. D. Trimble, Department of Geology and Geophysics, University of Wyoming, 1000 E. University Ave., Laramie, WY 82071, USA. 30 of 30