Dynamics of deformation and sedimentation in the northern Sierras Pampeanas: An integrated study of the Neogene Fiambalá basin, NW Argentina Barbara Carrapa† Department of Geology and Geophysics, University of Wyoming, Laramie, Wyoming 82071, USA Jörn Hauer* Institut für Geowissenschaften, Universität Potsdam, 14476 Golm, Germany Lindsay Schoenbohm School of Earth Sciences, Ohio State University, Columbus, Ohio 43210, USA Manfred R. Strecker Institut für Geowissenschaften, Universität Potsdam, 14476 Golm, Germany Axel K. Schmitt Department of Earth and Space Sciences, University of California, Los Angeles, California 90095-1567, USA Arturo Villanueva José Sosa Gomez Instituto de Mineralogía y Geología, Universidad Nacional de Tucumán, Miguel Lillo 209, San Miguel de Tucumán, Argentina ABSTRACT The thick-skinned Sierras Pampeanas morphotectonic domain of western and northwestern Argentina (27°S–33°S) is characterized by reverse-fault–bounded basement blocks that delimit internally deformed, Neogene sedimentary basins. Forelandbasin evolution in this part of the Andes is still not very well understood. For example, challenging questions exist as to how thickskinned deformation develops, if there are distinct spatiotemporal trends in deformation and exhumation, how such deformation styles influence sedimentation patterns, and whether or not broken foreland basins are related to regional plate-tectonic processes, such as flat-slab subduction. The Fiambalá basin of the northwestern Sierras Pampeanas is the largest of several intermontane basins in the transition to the southern margin of the Puna Plateau. This basin preserves a thick continental Neo† E-mail: bcarrapa@uwyo.edu *Present address: Department of Geosciences, University of Montana, Missoula, Montana 59812, USA. gene sequence that provides information on the dynamics of thick-skinned deformation and resulting sedimentation. The Fiambalá basin contains ~4 km of fluvial-alluvial sedimentary rocks that comprise the Tamberia, Guanchin, and Punaschotter Formations. U-Pb geochronology of ashes intercalated within the Fiambalá stratigraphic sequence demonstrates that these sedimentary rocks are late Miocene to Pliocene (8.2 ± 0.3 Ma to 3.05 ± 0.4 Ma) in age. Sedimentology and provenance data indicate that the source of the Tamberia Formation was located to the west of the modern western basin-bounding range. The Guanchin and Punaschotter Formations record input from local sources, including the modern basin-bounding range to the west and the southern Puna Plateau to the north, suggesting reorganization of the catchment area at ca. 5.5 Ma. The coarsening-upward trends recorded by the fluvial Tamberia and Guanchin Formations indicate enhanced tectonics and relief during sedimentation. The Punaschotter conglomerates record alluvial-fan sedimentation and local sources. Fault kinematic data document a contractional regime, characterized by E-W and NE-SW shortening, active throughout the middle-late Miocene and Pliocene. Furthermore, a comparison between the Fiambalá basin and similar sedimentary basins in the Sierras Pampeanas (e.g., Bermejo foreland basin) and the Eastern Cordillera leads us to propose that the study area originally constituted an integral part of a continuous and more extensive foreland-basin system (thin-skinned) for much of its early history. Our data suggest coeval intrabasin deformation along strike from the Bermejo region northward to the Eastern Cordillera. The coeval change at ca. 6 Ma from a regional to more compartmentalized (thick-skinned) tectono-sedimentary environment in the regions adjacent to the Eastern Cordillera, the southern Puna margin, and other sectors within the Sierras Pampeanas domain may thus reflect a regional tectonic process related to flat subduction. Our data, combined with existing sedimentological and petrological evidence, imply that the passage from steep to flat subduction occurred synchronously from ~30°S to ~26°S. Keywords: Sierras Pampeanas, foreland, sedimentation, thick-skinned deformation, thinskinned deformation, Andes. GSA Bulletin; November/December 2008; v. 120; no. 11/12; p. 1518–1543; doi: 10.1130/B26111.1; 13 figures; 2 tables; Data Repository item 2008143. 1518 For permission to copy, contact editing@geosociety.org © 2008 Geological Society of America Dynamics of deformation and sedimentation in the northern Sierras Pampeanas INTRODUCTION The Sierras Pampeanas morphostructural domain of western and northwestern Argentina constitutes a broken foreland with basement ranges and intervening sedimentary basins (e.g., Jordan et al., 1983). The Sierras Pampeanas are an integral part of the Cenozoic Andes orogen and are located between ~27°S and 33°S (e.g., González Bonorino, 1950; Allemendinger, 1986; Jordan and Allmendinger, 1986), a region that broadly coincides with a subhorizontal segment of subduction of the oceanic Nazca plate (Jordan et al., 1983; Ramos et al., 2002). This domain is composed of crystalline, ~N-S–oriented basement blocks reaching elevations in excess of 5 km that are bounded by reverse faults (Caminos, 1979). Sedimentary basins in the intervening low sectors are either fault bounded on their western and eastern margins or, in the case of asymmetric range uplift, bounded on one side by exhumed basement erosion surfaces that dip underneath the basins (Jordan and Allmendinger, 1986; Strecker et al., 1989; Mortimer et al., 2007). The structural uplifts and isolated, faultbounded basins in the foreland are akin to the Late Cretaceous to Eocene Laramide province of North America (Dickinson and Snyder, 1978; Jordan et al., 1983). Investigation into clastic sedimentary rocks deposited in intermontane basins within the Sierras Pampeanas provides fundamental information regarding the pattern and timing of source deformation and exhumation, basin dynamics, and sedimentary environments in evolving foreland-basin systems. North of 27°S, the Sierras Pampeanas domain merges into the Puna Plateau, which is part of the larger Central Andean Plateau and the Eastern Cordillera (e.g., Mon and Salfity, 1995). Similar to the Sierras Pampeanas, both domains are characterized mainly by reverse-fault–bounded blocks and intervening intermontane basins. The Puna-Altiplano region, also known as the Central Andean Plateau, is characterized by average elevations over 3.7 km, and it extends along strike for 1500 km (e.g., Isacks, 1988). Former foreland-basin deposits are preserved within the plateau interior and at its eastern margin (i.e., Eastern Cordillera) and have been used to constrain the pattern and timing of contractional deformation, erosion, and sedimentation (Jordan and Alonso, 1987; Viramonte et al., 1994; Vandervoort et al., 1995; Reynolds et al., 2000; Marrett and Strecker, 2000; DeCelles and Horton, 2003; Uba et al., 2005; Jordan and Mpodozis, 2006). Based on sedimentological, structural, and geophysical data, different authors have proposed that crustal thickening, eastward propagation of deformation, and initiation and development of a foreland-basin system began during the mid-Cretaceous and continued throughout the Tertiary (e.g., Horton et al., 2001; McQuarrie and DeCelles, 2001; DeCelles and Horton, 2003; Carrapa et al., 2005; Arriagada et al., 2006; Carrapa and DeCelles, 2008). In the transitional region between the plateau to the north and the Sierras Pampeanas domain to the south, the structural and sedimentary evolution remains ambiguous. A crucial issue, which we address with this study, is the structural and sedimentary behavior of broken forelands and their relationships with large-scale plate-tectonic processes such as flat-slab subduction. Our data also contribute to a better understanding of the transition from unbroken foreland (thin-skinned) to broken foreland (thick-skinned) styles of deformation and related sedimentation. Sedimentological evidence suggests that a foreland-basin system in the Bermejo region (30°S) commenced ca. 20 Ma but was later disrupted in the late Miocene by uplift of intrabasin crystalline ranges (Jordan et al., 2001). A oncecontiguous foreland basin likely existed east of the Eastern Cordillera of northwest Argentina as well at ~26°S–27°S, where pre-Miocene foreland-basin sedimentary rocks have been recognized (Starck and Vergani, 1996; Mortimer et al., 2007). A comparable setting existed at ~24°30′S in the Quebrada del Toro region of the Eastern Cordillera between 8 and 6 Ma (Hilley and Strecker, 2005) and prior to 3.5 Ma in the Quebrada de Humahuaca at 23°30′S (Walther et al., 1998; Strecker et al., 2007). All of these locations record the transition to a broken foreland and ensuing intermontane basin sedimentation characterized by Miocene-Pliocene strata that are sedimentologically similar to those observed in the Fiambalá basin and are interpreted to have been deposited in an intermontane, thickskinned setting (Hilley and Strecker, 2005; Carrapa et al., 2006a, 2006b; Coutand et al., 2006; Strecker et al., 2007). It is not clear, however, if a continuous (thin-skinned) foreland-basin system ever existed in the Fiambalá region. It is also unknown how the Fiambalá basin relates to the structurally and sedimentologically similar basins farther north in the Eastern Cordillera and northernmost Sierras Pampeanas (Horton, 1998; DeCelles and Horton, 2003; Coutand et al., 2006; Mortimer et al., 2007). Here, we present new geochronological, sedimentological, and structural data from a Miocene-Pliocene clastic sequence deposited in the Fiambalá basin in order to: (1) constrain the timing of sediment deposition in this part of the Andes; and (2) resolve the sedimentary and tectonic evolution of the Fiambalá basin within the Sierras Pampeanas morphotectonic domain. Ultimately, our new data set relates the dynamics of deformation and sedimentation in the Sierras Pampeanas region to the history of the Eastern Cordillera foreland-basin system. Structural and sedimentological data were collected mainly along two transects, referred to as Guanchin River and Northern transects (Fig. 1). GEOLOGICAL SETTING The study area is located in the southern Central Andes of northwest Argentina between 27°45′S and 67°45′S at the transition between the southern margin of the Puna Plateau and the reverse-fault–bounded, thick-skinned Sierras Pampeanas morphotectonic domain. Along strike to the south, at ~30°S, the Miocene to Holocene stratigraphic sequence preserved in the Bermejo basin has been interpreted as retroarc foreland-basin deposits characterized by three different tectono-sedimentary intervals: between 20 Ma and 7.3 Ma, a broad, wedgelike foreland basin developed to the east of the Western and Central Precordillera thrust belt; between 7.3 Ma and 6.5 Ma, a symmetrical foreland basin formed; and between 6.5 Ma and the present day, the area was characterized by thick-skinned deformation and segmentation of the contiguous foreland basin into small, reverse-fault–bounded basins typical of the present-day Sierras Pampeanas domain (Jordan et al., 2001). The Sierras Pampeanas domain is thought to be associated with diachronous, thickskinned deformation, which has been attributed to southeastward migration of flat-slab subduction (Ramos et al., 2002). Westward migration of deformation and exhumation has also been proposed (Coughlin et al., 1998). Diachronous deformation in the Eastern Cordillera, to the northeast of the Fiambalá basin, has resulted from propagation of the deformation front within a zone of structural inheritance related to a formerly extensional tectonic regime (e.g., Allmendinger et al., 1983; Allmendinger, 1986; Mortimer et al., 2007). In the neighboring El Cajón basin, this is recorded by both thermochronological and seismic-reflection data. Apatite fission-track (AFT) detrital ages document the appearance in the sedimentary section of a distinct grain-age population that is typical of the eastern basin-bounding range and that indicates uplift and erosion of the Sierra de Quilmes (Fig. 1) at 6 Ma (Mortimer et al., 2007). Seismic-reflection data show that loading and basin flexure were caused by reactivation of out-of-sequence structures bounding ranges to the west (Chango Real; Fig. 1) and east (Sierra de Aconquija; Fig. 1). Intrabasin deformation occurred at 6 Ma, resulting in exhumation of the Sierra de Quilmes range (Fig. 1) (Mortimer et al., 2007). In the Fiambalá basin, there is no clear evidence of Tertiary rocks older then ca. 8 Ma. Geological Society of America Bulletin, November/December 2008 1519 Carrapa et al. 68°0'0"W 67°0'0"W 66°0'0"W 14°S 200 km 0 st 25°0'0"S Ea er lle di Sier ras Sub and ina s Bermejo Basin 62°W Famatina Range Ran ang ange g ge A eal study area NPc SP Figure 3 6870 0m Figure 2 FN El Cajón Basin AG Sie rra Sierras Sie as Pampeanas Corral Quemado Basin 68°0'0"W Sierra d e Quilm es Ac on qu ija Cumbree s Calch aquíes leras Sierras Pampeanas Ch an go R 74°W CN Santa Barbara System Fronta l Co and Pre rdillera cordille ra Princip al and Coasta l Cordil Puna Fiambalá Fiambal Basin 28°0'0"S Angastaco Basin ra 26°0'0"S or 27°0'0"S C 26°S Axis Nazca Plate n ch en Tr Altiplano 2m 232 normal faults 0 Kilometers 0 67°0'0"W 25 50 75 25 Thrusts and reverse faults lineaments 100 B 66°0'0"W Pliocene (Punaschotter Formation) and Quaternary C Fiambalá stratigraphy Miocene (Guanchin and Tamberia Formations) Sed. rates (mm/yr) a b 4 km Punaschotter Fm. 3.6 ± 0.8 Ma (ZFT) 5.5 ± 0.1 Ma (U/Pb)* ~ 0.4 3.7 ± 0.1 Ma (U/Pb)* Quaternary volcanic rocks Quaternary evaporites Miocene-Pliocene volcanic rocks (andesite-dacite) Guanchin Fm. Triassic-Jurassic gabbro ~ 0.6 3 km late Miocene-Pliocene distal facies: a) Guanchin Formation, b) Punaschotter Formation Permian-Triassic granite 2 km Permian-Triassic red sandstone, conglomerate, marl, and related volcanic rocks (Paganzo Group) 8.2 ± 0.3 Ma (U/Pb) Cambrian-Ordovician volcanic (dacites) and metamorphic rocks 1 km Tamberia Fm. Ordovician phyllite, schist, and minor metavolcanic 5.7 ± 0.8 Ma (ZFT) 5.9 ± 1.2 Ma (ZFT) ~ 9 Ma (pmg) mrl fs ms ~. 2.5 Ordovician granodiorite and minor gabbro Cambrian phyllite, schist, gneiss, and metavolcanic rocks Paleozoic granite gr cgl Figure 1. (A) Digital elevation model (DEM) of the study area and morphotectonic domains (shown on inset map) (modified after Jordan et al., 1983; Isacks, 1988). White line corresponds to the area of the Puna Plateau characterized by internal drainage. (B) Geological map of the Fiambalá basin (location shown by white box in A and locations of the mapped areas and investigated transects (modified after Martinez, 1995). FN—Filo Negro apatite fission-track (AFT) sample; SP—Sierra de las Planchadas AFT sample; AG—Alto Grande AFT sample; CN— Cerro Negro AFT vertical profile (Carrapa et al., 2006). NPc—location at the northern end of the Fiambalá basin where extra investigations have been conducted (refer to text). (C) Simplified stratigraphy of the Fiambalá basin; ZFT (zircon fission-track) ages are after Tabutt (1986); paleomagnetic (pmg) data are from (Reynolds, 1987). (U/Pb)* ages are from dated ashes from this study (see Fig. 4). The best age for the Guanchin Formation (029) and one single mean age for the Punaschotter Formation are reported. Mrl—marls; fs—fine sand; gr—granule; cgl—conglomerate; ms—medium sand. Sedimentation rates were calculated using the ca. 9 Ma constraint from paleomagnetostratigraphy (Reynolds, 1987) for the base of the Tamberia Formation and our new U-Pb ages for the rest of the stratigraphy. 1520 Geological Society of America Bulletin, November/December 2008 Dynamics of deformation and sedimentation in the northern Sierras Pampeanas However, apatite fission-track data from basinbounding ranges at 28°S document a thick pile of sediments overlying these ranges at ca. 20 Ma (Coughlin et al., 1998). Furthermore, thermal modeling of apatite fission-track detrital populations suggests that the sediment sources for the Fiambalá basin strata were once buried, reaching temperatures up to ~90 °C (i.e., over 4 km of sediments considering a paleogeothermal gradient of 20 °C/km) until ca. 20 Ma, and they were subsequently removed (Carrapa et al., 2006). Such reworked sediments must have been deposited in areas to the east of the Fiambalá basin since evidence of reworking is not apparent in the Fiambalá strata. Thus, the sedimentary record observed today in the Fiambalá basin represents only a part of the complete sedimentary history in this region of the Andes; sedimentation may have started much earlier than the oldest sedimentary rocks preserved in the basin today. GEOLOGY OF THE FIAMBALÁ BASIN AND ITS BOUNDING RANGES ized by migmatite, schist, granite, and orthogneiss of Cambrian age (Turner, 1967). Apatite fission-track cooling ages from a vertical profile from the northern area range (Cerro Negro: CR in Fig. 1B) are between 23.1 ± 1.2 and 14.8 ± 0.7 Ma and document early to middle Miocene source cooling and exhumation (Carrapa et al., 2006b). Detrital apatite fission-track thermochronology and provenance data from the late Miocene–Pliocene Fiambalá stratigraphic section also suggest that the southern Puna Plateau margin, including the northern and eastern basin-bounding ranges (Cerro Negro and Alto Grande: CN and AG in Fig. 1B), acquired a similar (or greater) relief by 6 Ma, and that this was most likely associated with late-stage plateau uplift (Carrapa et al., 2006b). A single sample of granite from the eastern basin-bounding range (Alto Grande: AG in Fig. 1B) has an apatite fission-track age of 13.1 ± 1.7 Ma (Carrapa et al., 2006b), which may represent cooling during exhumation and removal of the sediments that once covered this range. Tertiary Clastic Sequence The Fiambalá basin contains a late Cenozoic continental clastic sequence bounded to the west by the northern termination of the Famatina Range (Fig. 1A). To the north and east, the basin is bordered by basement ranges belonging to the southeastern margin of the Puna Plateau. Western Basin-Bounding Ranges Immediately to the west of the Fiambalá basin, the Sierra de las Planchadas (part of the Famatina Range) is mainly composed of Precambrian granite, aplite, and diorite; Ordovician-Devonian siltstone-mudstone; Permian dacitic volcanic rocks, clastic sedimentary strata, and basaltic conglomerates (La Cuesta Formation); and Tertiary andesitic volcanic rocks (Fig. 1) (Turner, 1967; Martinez, 1995). Previous studies have shown that the Filo Negro (FN; Fig. 1B) and the Sierra de las Planchadas (SP, Fig. 1B) Ranges, to the west of the Fiambalá basin, are characterized by apatite fissiontrack cooling ages of 43.9 ± 4.4 and 34.6 ± 4.2 Ma, respectively, suggesting that deformation and associated exhumation of these ranges had already started by the end of the Eocene (Carrapa et al., 2006b). This is also supported by apatite fission-track data obtained south of the study area (Coughlin et al., 1998). Northern and Eastern Basin-Bounding Ranges Basement rocks located to the north and northeast of the Fiambalá basin are character- The Tertiary basin fill is divided into the Tamberia and Guanchin Formations; the unconformably overlying conglomerates are referred to as the Punaschotter Formation, following the terminology introduced by Penck (1920) (Fig. 1C). Whereas the Guanchin and Punaschotter Formations are exposed in the north as well as to the south of the basin, the Tamberia Formation crops out only along and south of the Guanchin River (Fig. 2A). Distal facies of the Punaschotter Formation crop out along the Guanchin River transect (Fig. 2). Prior to this study, little was known about the stratigraphic age of these sediments. Reynolds (1987) proposed an age of ca. 9 Ma for the base of the Tamberia Formation based on magnetostratigraphy (Fig. 1C). One ash layer in the Tamberia Formation was dated at 5.9 ± 1.2 Ma (zircon fission-track age) by Tabutt (1986). Two ash layers in the Guanchin Formation were dated at 5.3 ± 0.7 Ma and 3.6 ± 0.8 Ma, respectively, by Tabutt (1986) (Fig. 1C); however, these ashes were collected from a highly deformed stratigraphic section (Fig. 2A). It must be noted that most of the ashes in the study area display some degree of reworking. Therefore, a careful reassessment of the stratigraphy of these sedimentary rocks is necessary. where detailed mapping and sedimentological investigations were conducted during this study. An additional ash from the lower portion of the Guanchin Formation and one from the distal Punaschotter Formation were collected along the Northern transect (Fig. 3) to correlate the different mapped areas and stratigraphy. Results are presented in Figure 4 and Table 1. The stratigraphic position of the collected samples was carefully selected on the basis of detailed mapping across the entire basin (Figs. 2 and 3). Mineral separation was conducted at the University of Potsdam following established procedures. Zircon analysis was performed using the University of California–Los Angeles (UCLA) Cameca IMS 1270 ion microprobe with a mass-filtered, ~15 nÅ 16O– beam focused to a 25–30-μm-diameter spot. For further details on zircon U-Pb analysis, refer to Appendix 1.1 U-Pb Results Sample 006 is an ash collected from the upper portion of the Tamberia Formation (member B, discussed in the following). The age obtained from eight grain analyses is 8.20 ± 0.33 Ma, demonstrating that the base of the Tamberia Formation is likely older than 8.2 Ma. This is consistent with magnetostratigraphic data, which suggest an age of ca. 9 Ma for the base of the Tamberia Formation (Reynolds, 1987). Two ashes (029; conglomeratic facies) and (055; sandstone facies) were collected in the Guanchin Formation along the Guanchin River transect (Fig. 2A). Seven out of ten grains in sample 029 give a mean age of 5.51 ± 0.15 Ma. The youngest grain out of five zircons from sample 055 has an age of 5.23 ± 0.30 Ma. The range of ages of other grains in sample 055, from late Oligocene to late Miocene, suggests considerable reworking; a lack of material prevented additional analyses. The youngest zircon age is, however, consistent with the age from sample 029 and sample J26/04 collected farther north (Fig. 3A). Ash J26/04, from the lower part of the Guanchin Formation, was collected along the Northern transect (area indicated in Fig. 3A), and it yielded an age of 5.90 ± 0.20 Ma. Results from samples 029 and 055 corroborate a late Miocene age for these sedimentary rocks. In the following, we use the better constrained age of sample 029 for our geological interpretations. Three ashes (001, 074, and 003 from top to bottom) along a ~20 m vertical section were collected from the top of the Punaschotter U-Pb GEOCHRONOLOGY We present new U-Pb geochronological data from eight ashes, six of which were collected along the Guanchin River transect (Fig. 2A), 1 GSA Data Repository Item 2008143, U-Pb geochronology, is available at www.geosociety.org/ pubs/ft2008.htm. Requests may also be sent to editing@geosociety.org. Geological Society of America Bulletin, November/December 2008 1521 1522 0 Geological Society of America Bulletin, November/December 2008 27°45'S 27°40'S 27°35'S A A' 2.5 40 25 22 006 Pc 1-2 60 Pc 9-10 L2 L3 50 36 Pc 7-8 35 Tb 10 20 L1 45 Ta 25 029 Pc6 F3 F1 Pc13 L5 10 L8 13 17-18 003074L9 001 Pc5 Pc 05 15 km 67°50'W Pc 11-12 L4 60 65 75 75 L7 G n i ch n ua r ve Ri 25 Pc15 area investigated by Tabutt (1986) and Reynolds (1987) Pc14 L6 15 08 F2 Pc3-4 40 67°45'W 25 25 Pc16 30 55 35 20 20 Basement rocks (Carboniferous-Permian) Tamberia Formation (facies a:Ta; facies b:Tb) Guanchin Formation Punaschotter Formation Undifferentiated 10 20 Pc20 Pc19 25 5 A' B' location of cross section e.g. 029: ash sample for U-Pb analysis reverse fault syncline main river fractures in pebbles-cobbles F anticline detailed logs L Pc pebble counting locations B' 67°40'W Figure 2. (A) Geological map of the Guanchin River transect; for location refer to Figure 1B. (Continued on following page.) 70 70 F4 5 67°55'W Carrapa et al. Geological Society of America Bulletin, November/December 2008 -2000 0 -1000 0 1000 2000 3000 A’ (1) Tb P 5 (2) P: Punaschotter Formation G: Guanchin Formation Tb: Tamberia Formation (facies b) Ta: Tamberia Formation (facies a) 1 10 Ta (3) Figure 8 2 15 Distance (km) G P 3 20 Tb Figure 10 Figure 9 ? 25 P ? G 30 ? B’ Figure 2 (continued). (B) Structural profile along the Guanchin River transect. 1, 2 and 3 are locations along the profile where kinematic analysis was undertaken. Fault planes are plotted on the stereoplot; the small arrows on the cyclographic traces indicate the movement of the hanging wall based on slickenside measurements. Kinematic results are presented as “fault plane solution” (cf. Allmendinger, 2001) indicating pressure-tension (P-T) quadrants. The gray quadrants indicate extensional fields from which shortening direction is derived. (1) Thrust contact between the Carboniferous-Permian red beds and the Punaschotter Formation. (2) Thrust contact between the Tamberia and upper Guanchin Formations. (3) Thrust within the Guanchin Formation. See text for more explanations. Elevation (m) B Dynamics of deformation and sedimentation in the northern Sierras Pampeanas 1523 Carrapa et al. 67°54' W 27°18' S A 67°52' W A' Pc 6j 66 PS1 67°50' W Pc 5j Pc 4j 67°48' W 67°46' W Pc 3j 70 11 18 J28/04 Pc 1j Pc 2j 10 60 48 14 29 18 55 16 L2 27°20' S 10 Pc 12j L3 Palo Blanco L4 17 15 55 PS2 ? ? 54 12 Pc 11j J26/04 16 Pc J4/04 10j 42 27°20' S L1 B' 50 ? Pc 7j 27°22' S 60 24 77 22 70 74 18 9j 24 12 16 24 70 80 do olora Rio C 80 82 Pc 20 20 27°22' S 45 Pc 8j 85 27°24' W 27°24' S 85 80 74 0 1 2 km 67°54' W 67°52' W 67°50' W 67°48' W 67°46' W Holocene (undifferentiated sediments) Carboniferous-Permian strike & dip Early to late Pliocene (Punaschotter Fm.) Ordovician-Carboniferous overturned layer Late Miocene to early Pliocene (Guanchin Fm.) A' B' thrust fault location of cross section e.g. J26/04: ash sample for U-Pb analysis NW SE P 1000 2000 P Tb -1000 -2000 Ta -3000 P: Punaschotter Formation G: Guanchin Formation Tb: Tamberia Formation (facies b) Ta: Tamberia Formation (facies a) -4000 -5000 -6000 -7000 0 2 P 1000 G 0 -8000 SE 3000 2000 Elevation (m) NW 3000 Elevation (m) B anticline 4 6 8 10 Distance (km) P G 0 Tb -1000 -2000 Ta -3000 -4000 -5000 -6000 décollement 12 14 -7000 -8000 0 2 4 8 6 10 Distance (km) 12 14 Figure 3. (A) Geologic map of the Northern transect; for location refer to Figure 1B. (B) Balanced cross section along the Northern transect showing both the box fold (left) and blind thrust (right) scenarios; for location refer to Figure 3A. 1524 Geological Society of America Bulletin, November/December 2008 Dynamics of deformation and sedimentation in the northern Sierras Pampeanas A n ( ) A ( ) n n n n n n F n Figure 4. Zircon U-Pb results; 207Pb/206Pb versus 238U/206Pb and 206Pb/238U model ages (histograms and probability density curves). Plot shows fixed-intercept regression (207Pb/206Pb = 0.83) and disequilibrium-corrected concordia curve between 3 and 20 Ma. Weighted average ages for individual samples are stated with 2σ uncertainties. MSWD—mean square of weighted deviates. Geological Society of America Bulletin, November/December 2008 1525 Carrapa et al. TABLE 1. ZIRCON U-Pb RESULTS Sample-grain-spot 006-g1-s1 006-g2-s1 006-g5-s1 006-g1-s2 006-g4-s1 006-g8-s1 006-g9-s1 006-g14-s1 006-g18-s1 006-g19-s1 Pb*/238U (x10–3) 3.059 76.76 4.004 1.268 12.74 1.300 1.922 4.748 1.336 8.627 206 Pb*/238U (1 s.e.) (x10–3) 0.131 3.81 0.198 0.0666 0.464 0.0466 0.0919 0.186 0.0368 0.279 207 Pb*/235U (x10–3) 235.5 618.9 107.9 15.25 1387 11.55 74.67 407.4 14.02 860.9 207 Pb*/235U (1 s.e.) (x10–3) 11 31.2 10.4 1.44 53.1 0.882 7.03 15.2 1.03 29.4 Correlation of concordia ellipses 0.80 0.90 0.78 0.29 0.89 0.22 0.81 0.52 0.48 0.80 Disequilibrium corrected, 207Pb corrected† Pb/238U ± Radiogenic U age (Ma) Age (Ma) % (ppm) 6.88 1.04 34.5 417 476 23 99.8 256 20.9 1.4 80.9 215 7.80 0.44 94.8 431 4.12 4.22 4.9 188 8.26 0.30 97.7 403 8.72 0.71 69.9 298 8.11 1.47 26.3 236 8.35 0.24 96.2 695 7.53 2.44 13.4 341 206 Th (ppm) 353 59 164 626 239 323 316 333 613 114 055-g6-s1 055-g5-s1 055-g4-s1 055-g3-s1 055-g2-s1 4.539 4.282 5.081 0.8194 1.562 0.207 0.191 0.234 0.0231 0.0617 293.6 95.51 413.9 7.623 36.93 13.1 6.04 24.4 0.459 2.66 0.69 0.62 0.84 0.31 0.32 13.5 23.6 10.0 5.23 8.54 1.5 1.3 2.0 0.15 0.42 45.9 85.3 30.4 97.3 84.0 174 157 523 818 186 79 91 186 273 110 029-g14-s1 029-g13-s1 029-g11-s1 029-g10-s1 029-g9-s1 029-g8-s1 029-g6-s1 029-g5-s1 029-g4-s1 0.8567 5.027 0.8630 0.8623 0.8588 0.9248 0.8907 85.93 0.9154 0.0251 0.225 0.0389 0.0303 0.0305 0.0283 0.0294 3.17 0.0267 6.058 45.97 9.175 9.332 9.11 15.55 8.373 710.1 13.02 0.43 3.55 1.06 0.78 1.2 1.2 0.601 31.9 0.555 0.39 0.53 0.49 0.65 0.14 0.63 0.28 0.80 0.48 5.59 31.6 5.42 5.42 5.41 5.42 5.68 530 5.57 0.16 1.5 0.26 0.20 0.21 0.19 0.19 19 0.17 99.3 97.5 96.0 95.9 96.1 90.3 97.2 99.7 92.7 1101 168 336 669 284 2088 775 38 1514 215 107 249 296 149 3707 179 23 141 074-g1-s1 074-g2-s1 074-g6-s1 074-g7-s1 074-g9-s1 074-g12-s1 074-g13-s1 074-g14-s1 074-g15-s1 074-g17-s1 074-g19-s1 1.009 0.6217 0.5889 0.6118 0.592 0.6827 0.7122 0.6245 0.5698 3.881 0.5815 0.0355 0.0214 0.017 0.0294 0.0203 0.028 0.0378 0.0268 0.0153 0.205 0.0195 9.599 5.553 4.746 9.592 8.801 21.38 18.67 10.09 4.541 384.3 4.954 0.946 0.681 0.349 1.78 0.977 1.39 2.36 0.875 0.327 22.8 0.595 0.20 0.08 0.29 –0.11 –0.11 0.36 0.37 0.01 0.24 0.89 0.03 6.40 4.00 3.82 3.68 3.58 3.46 3.82 3.76 3.71 3.61 3.76 0.23 0.14 0.11 0.21 0.14 0.20 0.28 0.18 0.10 1.84 0.13 97.1 97.6 98.4 91.4 92.1 76.9 81.6 90.9 98.5 14.1 98.0 376 886 1931 150 523 352 272 418 1553 158 870 157 399 1421 107 556 277 217 133 569 109 427 001-g5-s1 001-g6-s1 001-g7-s1 001-g8-s1 001-g9-s1 001-g13-s1 001-g14-s1 001-g17-s1 001-g18-s1 001-g19-s1 1.510 0.5858 1.937 0.5739 4.896 0.6144 1.245 0.5671 0.6224 0.5984 0.0405 0.02 0.0673 0.0177 0.8950 0.0218 0.0356 0.0293 0.0317 0.0233 105.4 4.814 162.6 6.073 507.4 8.786 81.92 7.734 7.879 9.668 3.91 0.556 7.55 0.468 103 0.723 2.19 1.04 1.07 1.01 0.59 0.06 0.54 0.44 0.88 0.47 0.85 0.39 –0.13 0.29 4.09 3.79 3.57 3.64 3.17 3.74 3.69 3.48 3.85 3.58 0.34 0.13 0.61 0.12 8.16 0.15 0.26 0.20 0.21 0.16 41.2 98.3 28.1 96.1 9.8 92.6 44.9 93.2 94.2 90.9 1400 579 336 690 1246 606 1221 191 207 666 796 424 360 429 908 680 562 176 190 565 003-g1-s1 003-g2-s1 003-g3-s1 003-g4-s1 003-g5-s1 003-g6-s1 003-g7-s1 003-g9-s1 003-g10-s1 003-g14-s1 0.6558 0.6062 3.176 0.6571 0.5898 0.6680 0.6263 0.7904 0.5831 0.6579 0.0284 0.0235 0.101 0.0201 0.0186 0.0307 0.0202 0.0781 0.0152 0.0449 13.89 9.734 294.4 10.35 4.894 13.07 9.836 22.96 5.216 12.59 1.7 0.875 11.7 0.647 0.617 1.13 0.684 3.12 0.358 2.08 0.16 0.34 0.77 0.37 0.28 0.42 0.51 0.61 0.48 –0.10 3.72 3.62 4.16 3.96 3.81 3.86 3.76 4.10 3.73 3.82 0.21 0.16 0.92 0.13 0.12 0.21 0.14 0.53 0.10 0.31 86.3 91.0 19.9 91.3 98.2 87.8 91.3 79.0 97.6 88.2 287 479 247 983 1065 299 679 63 1323 142 241 590 226 411 937 243 606 47 1623 98 1.567 1.398 79.29 43.84 0.8928 0.7881 93.19 181.7 26.34 76.51 1.126 0.5845 1.029 0.099 0.112 1.95 0.879 0.0474 0.0364 1.88 6.45 0.819 2.88 0.128 0.026 0.106 0.84 0.85 0.63 0.83 0.78 0.52 0.86 0.77 0.89 0.78 0.60 0.67 0.84 2.97 3.25 479 274 2.87 3.16 571 1034 5.09 472 3.00 2.83 3.17 0.92 0.93 12 5 0.41 0.29 11 36 7.50 17 0.98 0.21 0.94 29.1 35.3 97.2 99.1 49.5 61.2 99.3 95.8 3.0 99.2 40.7 74.6 46.6 307 341 139 779 416 752 376 61 332 458 142 717 298 617 360 47 368 1018 1151 74 29 450 256 243 1713 180 J26-04-g1-s1 1.384 0.0447 59.88 3.38 0.84 5.83 0.34 65.8 J26-04-g2-s1 1.099 0.0321 25.96 1.87 0.63 5.97 0.23 84.0 J26-04-g3-s1 1.184 0.0337 34.42 2.78 0.71 6.11 0.27 78.9 J26-04-g4-s1 1.201 0.0349 39.03 2.93 0.71 5.92 0.28 75.8 J26-04-g5-s1 1.056 0.0288 24.67 1.66 0.73 5.81 0.21 84.2 J26-04-g6-s1 1.17 0.0253 31.13 1.34 0.55 6.21 0.18 81.2 J26-04-g7-s1 1.778 0.0942 114.2 9.88 0.70 5.40 0.82 46.3 J26-04-g8-s1 1.494 0.0755 71.13 6.58 0.85 6.03 0.61 61.8 J26-04-g9-s1 37.06 0.743 286.3 6.51 0.86 233.1 4.6 99.3 J26-04-g10-s1 1.349 0.0561 58.31 4.83 0.67 5.80 0.45 65.8 Note: 207Pb was corrected using common Pb compositions: 206Pb/204Pb = 18.86; 207Pb/204Pb = 15.62 (Sañudo-Wilhelmy and Flegal, 1994). *Radiogenic. †230 Th disequilibrium correction factor was calculated from (Th/U)zircon/(Th/U)melt, where (Th/U)melt = 3 (average from Siebel et al., 2001). 911 2488 1217 1056 2291 1420 342 612 1292 493 3498 5522 672 1535 2038 1056 182 428 794 434 J4-04-g1-s1 J4-04-g2-s1 J4-04-g3-s1 J4-04-g4-s1 J4-04-g5-s1 J4-04-g6-s1 J4-04-g7-s1 J4-04-g8-s1 J4-04-g9-s1 J4-04-g10-s1 J4-04-g11-s1 J4-04-g12-s1 J4-04-g13-s1 1526 206 129.8 106.5 853 356.2 54.3 37.96 830.6 2644 2924 661 79.14 19.71 65.84 11.8 10.4 33.1 8.66 4.92 2.94 20.2 143 95.1 25.6 9.51 2.15 11.4 Geological Society of America Bulletin, November/December 2008 Dynamics of deformation and sedimentation in the northern Sierras Pampeanas Formation, along the Guanchin River transect, at the highest elevation reached by those sedimentary rocks in the basin (Fig. 2A). In all three samples, more than 10 grains yielded overlapping Pliocene ages; from top to bottom, sample 001 gave an age of 3.69 ± 0.06 Ma, sample 074 gave an age of 3.74 ± 0.05 Ma, and sample 003 gave an age of 3.77 ± 0.05 Ma. Only one late Miocene grain was dated, and based on the lack of detrital contamination, reworking of these ashes appears to have been minor. Ash J04/04 was collected from distal facies of the Punaschotter Formation (Fig. 3); eight grains yielded an age of 3.05 ± 0.44 Ma. These results, therefore, date the deposition of the Punaschotter Formation in the Fiambalá basin to the early-middle Pliocene. Tabutt (1986) suggested an age of 3.6 ± 0.6 Ma for the Guanchin Formation on the basis of zircon fission-track dating of one ash collected along the highly deformed portion of the Guanchin River transect (Fig. 2A). This result implies that the overlying Punaschotter conglomerates must be younger than ca. 3.6 Ma. However, our new data document that the onset of the Punaschotter deposition must have been older than 3.77 ± 0.05 Ma, as indicated by sample 003 from the top of the Punaschotter conglomerates along the Guanchin River transect. These new ages thus refine the stratigraphic record of the basin fill and provide the opportunity to assess the onset of conglomerate deposition in adjacent intermontane basins. SEDIMENTOLOGY The following descriptions are based on detailed facies analysis and measured sections, conducted by centimeter-scale tape and compass measurements, along the Guanchin River and Northern transects (Figs. 5, 6, and 7). Tamberia Formation The Tamberia Formation crops out only in the central (Guanchin River transect) and southern parts of the basin (Figs. 2 and 3), where it consists of over 2000 m of red sandstone, siltstone, and clast-supported, well-organized, and wellsorted conglomerate. On the basis of distinct lithofacies, we have divided the Tamberia Formation into two members (members A and B). The lower member A (Figs. 5A and 5B) consists of ~1000 m of reddish medium-grained sandstone bodies of meter-scale thickness, intercalated with centimeter- to meter-scale mudstones with large mud cracks (up to 1 m scale); the sandstone to mudstone ratio is >>1. These units generally lack sedimentary structures, have a massive to tabular character, and show only local parallel and trough cross-stratification. Up-section, this member becomes coarser and grades into well-sorted granular sandstone and fine- to coarse-grained, horizontally stratified, generally normally graded conglomerates that constitute the upper member B. This member reaches a thickness of ~1000 m and has an overall massive and tabular character, with local crude horizontal bedding and imbricated clasts, and a general lack of deep scour surfaces (Fig. 5C). The overall external geometry of the sedimentary bodies of the Tamberia Formation is tabular; the lateral extent of the sedimentary bodies constituting this formation is on the order of several tens of kilometers. The member A strata show a lack of features, such as lateral accretion foresets, prominent levees, oxbow lake deposits, and crevasse splay deposits, which would indicate meandering channels and anastomosing fluvial systems. For this reason, and because of the massive and laterally extensive character of the sand bodies, we suggest that the lower Tamberia strata were deposited in low-sinuosity, sandy, braided channels (Smith, 1970; Allen, 1983; Miall and Turner-Peterson, 1989; Hein and Walker, 1977; Smith and Perez-Arlucea, 1994; Miall, 1996; Makaske, 2001; Lunt and Bridge, 2004; Wooldridge and Hickin, 2005). Fine-grained deposits are often intercalated among the sandstone beds, but they lack paleosol development and plant remains, suggesting river bank instability and poorly confined flood flows (Nemec and Steel, 1984; Miall, 1996). Furthermore, the presence of mud cracks in the mud layers suggests an episodically dry environment. The horizontally stratified, imbricated, and locally trough cross-stratified conglomerates of member B, which lack sharp erosional contacts, are interpreted to be the result of bed-load deposits of broad channels on the floodplain (Nemec and Steel, 1984; Miall, 1996). The limited trough cross-stratified sandstones are interpreted as channel deposits (Nemec and Steel, 1984; Miall, 1996); the associated siltstone layers were derived from waning flow conditions following channel deposition and abandonment. The general tabular and coarse character of these sedimentary bodies suggests deposition through poorly confined flows in a high-energy river representing the proximal part of a large braided fluvial system. We interpret the coarsening-upward trend from member A to member B to be the result of progradation of the depositional system and possible enhancement of tectonic activity driving source exhumation, erosion, and sedimentation. The provenance data discussed later document erosion of source rocks directly west of the Fiambalá basin, supporting this interpretation. Guanchin Formation The Guanchin Formation consists predominantly of sandstone and conglomerate. We estimate a minimum thickness of ~1500 m along the Northern transect (Fig. 3A). The contact with the stratigraphically lower Tamberia Formation along the Guanchin River transect is a high-angle reverse fault (Fig. 2A); the contact is not exposed along the Northern transect. The Guanchin Formation is better exposed along the Northern transect (Fig. 3), which is essentially undeformed, whereas the Guanchin River transect is deformed and cut by several structures (Figs. 2A and 2B). The following facies descriptions are based primarily on observations and logging along the Northern transect, complemented by additional observations and limited logging along the Guanchin River transect (Figs. 6A and 6B). The Guanchin Formation consists of medium- to coarse-grained lenticular gray to pink, centimeter- to meter-scale sandstones, with intercalations of fine- and medium-grained conglomerates and mudstones. A coarseningand thickening-upward trend is observed along both transects. In the sandstone bodies, planar to trough cross-stratification, up to 2 m high, is common, but massive sandstone bodies are also present. The sandstone bodies generally show erosional bases. Dewatering structures occur locally. Centimeter- to meter-scale conglomeratic lenses are common and frequently show fining-upward trends and express sharp erosional basal contacts. Massive- to planar-laminated red mudstone layers, which range in thickness from 0.5 cm up to 40 cm, are typical of these deposits. In some cases, the mudstones are reworked and deposited as muddy intraclasts (diameter up to 25 cm; Fig. 6C) within sandstones. The mud layers must have consolidated prior to erosion and reworking, suggesting a periodically dry environment. Poorly sorted and moderately rounded, gray to pink conglomerate occurs toward the top of the formation. The conglomerate beds have a massive character and local horizontal stratification and erosional basal contacts. On the basis of its lithofacies association (Table 2) and the absence of typical features associated with meandering channels, the Guanchin Formation is interpreted to have been deposited in an ephemeral braided fluvial environment. We infer the planar and trough cross-stratified sandstones to be representative of channel deposits. The structureless, meter-scale sandstone bodies are interpreted to be the result of rapid deposition from suspension during floods. The conglomerate bodies with sharp erosional contacts are considered to represent channel fills resulting from bed-load Geological Society of America Bulletin, November/December 2008 1527 Carrapa et al. A Log 3-Tb (facies B) Tamberia Fm. (Guanchin transect) Log 2-Tb (facies B) Tamberia Fm. (Guanchin transect) Log 1-Ta (facies A) Tamberia Fm. (Guanchin transect) +~1200 m (estimated cover) Sr Gh +~750 m (estimated cover) Sh St 50 m 50 m Sh Gc Pc8 Pc9 +~300 m (estimated cover) Gc Sh/St-Gt Gc Sh Sr St Sh 25 m 25 m Gh Sm 25 m 25 m St/Gt Gh Pc10 Sr Sh Sh Pc7 Mrl FS MS Gr Cgl(~5cm) Cgl(~10cm) Mud cracks Conglomerate Ripples Trough cross-stratification Sandstone Intraclasts Ignimbrite (1) Sandstone with granules (2) with conglomeratic lens(1) and channel(2) Pc: pebble counting locations e.g. Sp: facies after Miall (1978), refer to Table 2. Figure 5. (A) Sedimentological logs of the Tamberia Formation along the Guanchin River transect; refer to Table 2 for facies explanation and to Figure 2A for location. Mrl—marlstone; fs—fine sandstone; ms—middle sandstone; gr—granule; cgl—conglomerate. (Continued on following page.) 1528 Geological Society of America Bulletin, November/December 2008 Dynamics of deformation and sedimentation in the northern Sierras Pampeanas PHOTO 1 Tabular, massive sandstones member A Tamberia Fm. PHOTO 2 Massive sandstones with mud cracks member A Tamberia Fm. 1m person for scale B PHOTO 1 Horizontally laminated and stratified, clast-supported conglomerates (Gh) member B Tamberia Fm. PHOTO 2 Imbricated, clast-supported conglomerates (Gi) member B Tamberia Fm. imbrications 1m 1m C Figure 5 (continued). (B) Typical facies A of the Tamberia Formation: massive and tabular sandstones (photo 1) with occasional large mud cracks (photo 2). Refer to Table 2 for facies explanation. (C) Typical facies of the Tamberia member B Formation: horizontally laminated and stratified (photo 1), occasionally imbricated (photo 2), clast-supported conglomerates. Refer to Table 2 for facies explanation. Geological Society of America Bulletin, November/December 2008 1529 Carrapa et al. A Mud cracks Conglomerate Ripples Trough cross-stratification Sandstone Intraclasts Ignimbrite (1) (2) Sandstone with granules and conglomeratic channel(1) and lens(2) Log 5 Guanchin Fm. (Guanchin transect) +~140 m (estimated cover) 15-10-03/09#* Log 4 Guanchin Fm. (Guanchin transect) Pc: pebble countings e.g. Sp: facies after Miall (1978); refer to Table 2. 50 m Log 6 Guanchin Fm. (Guanchin transect) +~150 m (estimated cover) Pc13 Pc12 25 m 25 m 25 m Ash 055 Pc14 Pc11 Mrl. FS MS Gr Cgl(~5cm) Cgl(~10cm) Figure 6. (A) Sedimentological logs of the Guanchin Formation along the Guanchin River transect; refer to Table 2 for facies explanation and to Figure 2A for location. Mrl—marlstone; fs—fine sandstone; ms—middle sandstone; gr—granule; cgl—conglomerate. (Continued on following two pages.) 1530 Geological Society of America Bulletin, November/December 2008 Dynamics of deformation and sedimentation in the northern Sierras Pampeanas B Guanchin Fm. (Northern transect) 158.63 m (2) (1) Sh Gc 150 m Sh, Gh 150 m 150.26 m Sh Fl Sh Fl Gh 125 m 125 m Fl Sp Fl Sh cobbles & boulders water escape structures intraclasts (mudstone) sandstone with conglomeratic lens(1) and channel(2) geometrically estimated distance from L3 top to L4 basis: 383 m Gc 125 m 122.15 m Sh Sh, Gh fining upward coarsening upward Fl 100 m Sh Sh Gc Sh Gc pc: pebble countings e.g. Gh: main facies after Miall (1978) Sh Sh Gh Gc Fl Sh Sh Figure 6 (continued). (B) Sedimentological logs of the Guanchin Formation along the Northern transect; refer to Table 2 for facies explanation. Refer to Figure 3A for location. (Continued on following page.) Gc Sh Fl Sh 100 m 100 m Gh Gh Sh Gh Gp Sh Fl Sh Gc Sh Sh 220 m Gh Gh Sh Sh Sh Sh Fl Gh Sh trough cross-stratification occasional boulders (>50cm diameter) Fl 75 m pebbly sandstone with granules ash layer/ignimbrite Gc 100 m planar cross-stratification Quaternary cover Gh Sh parallel lamination or bedding Sh Sh 125 m mudstone sandstone Sh Gh, Sh Fl 75 m 75 m Fl Gp Gh Fl Fl Gh Sh, Sm 75 m Sh, Gh 200 m Gh Gh Fl Sh Sh Sh Gh Sh Fl 50 m Sh Fl 50 m 50 m Gc Gh Gc Sp Fl Sh Gh 25 m Gh Sh Sh, Sm Sh, Fl Gp Fl Sh Sp Sh Gh Gt Sh Fl Sh Gh St Fl 25 m Gh 175 m Sh Sh Gh Sh Sh, Sm Sh, Gh Sh, Gh Fl Fl Gc Gh Fl Gh Sh 25 m Sh Sh J26/04 Sh, Gh 50 m Sh, Sm Gc 25 m Fl Sp Sh 150 m Fl Sh Sh Sh Sp 0m Mrl FS MS Gr Cgl (~5 cm) Cgl (~10 cm) L1 Bottom Gh pc 11J Sh Gh pc 12J Fl Sh Sp 0m Mrl FS MS Gr Cgl (~5 cm) Cgl (~10 cm) distance: 558 m L2 Gh Sh Sh 0m Mrl FS MS Gr Cgl (~5 cm) Cgl (~10 cm) distance: 879 m L3 0m J6/04# Fl Gh Fl Gh Mrl FS MS Gr Cgl (~5 cm) Cgl (~10 cm) Mrl FS MS Gr Cgl (~5 cm) Cgl (~10 cm) distance: 1973 m Sh, Fl L4 Top Geological Society of America Bulletin, November/December 2008 1531 Carrapa et al. C 30 cm PHOTO 2 Trough cross-stratified sandstones of the Guanchin Fm. 1m PHOTO 1 Trough cross-stratified conglomerates (Gt) and trough cross-stratified sandstones (St) of the Guanchin Fm. Figure 6 (continued). (C) Typical facies of the Guanchin Formation: trough cross-bedded conglomerates (Gt) (photo 1), trough cross sandstones (St), and reworked mud clasts (photo 3). Refer to Table 2 for facies explanation. 30 cm PHOTO 3 reworked mud clasts in the Guanchin Fm. 1532 Geological Society of America Bulletin, November/December 2008 Dynamics of deformation and sedimentation in the northern Sierras Pampeanas A Location in the basin Northern transect Location in the basin Guanchin transect West center East (distal) West East (distal) (pc6) Pc18-P 001: 3.69 ± 0.12 Ma 074: 3.74 ± 0.10 Ma 003: 3.77 ± 0.10 Ma mrl n=50 fs ms mcg cgl Pc20-P n=49 Punaschotter Formation Pc15-P Pc4j n=48 Pc4-G Pc16-G Pc12-G 1km Pc3-G Guanchin Formation Pc5j Pc19-P Pc17-P Pc6-G Pc10j n = 100 (PS100) J4/04: 3.05 ± 0.44 Ma Pc3j PS2 mrl fs ms mcg Pc2j cgl Pc1j n = 10 (PS31_A) Pc9j Pc7j n = 15 (PS31_B) 055: 5.23 ± 0.30 Ma Pc11-G PS1 1km Pc5-P 0.5 km Punaschotter Formation Pc6jjj n = 12 (PS2_4) Pc8j mrl fs ms mcg n = 6 (PS1_4) cgl Pc13-G 2 km n=11 member B member A Pc10-T 006: 8.20 ± 0.33 Ma Pc2-T Pc8-T Pc1-T Pc7-T 1 km Tamberia Formation Pc9-T 1 km Guanchin Formation Pc14-G Pc11j Pc12j J26-04: 5.91 ± 0.22 Ma cgl mrl fs ms mcg Typical source lithologies present in the source and relative abundance in % cgl West of Guanchin profile West of Northern profile Granite Pelitic intraclast Aplite Red sandstone Dacitic volcanic and metasediment Andesitic volcanic White sandstone Mud crack Paleosol Plant remain mrl fs ms mcg Qz. and Quartzite Schist-Gneiss/Phyllite Northeastern (southern Puna) and eastern sources 50% Diorite e.g. 055, J26/04: U-Pb age of zircons from ashes (Figures 2, 3, 4) Figure 7 (continued on following page.) Geological Society of America Bulletin, November/December 2008 1533 Carrapa et al. B Typical source lithologies 100% Typical clast lithologies along the Guanchin River transect Typical clast lithologies along the Northern transect Granite Aplite Dacitic volcanic and metasediment Andesitic volcanic Pelitic intraclast Red sandstone 80% White sandstone Qz. and Quartzite Schist-Gneiss/Phyllite 60% Diorite 40% Facies codes Gm Gc Gh/i Gt/p Sm St Sp Average Punaschotter Fm. Average Guanchin Fm. Average Punaschotter Fm. Average Guanchin Fm. TABLE 2. FACIES TABLE FOLLOWING THE SCHEME OF MIALL ET AL. (1978) Lithofacies Sedimentary structures Interpretation Conglomerate, matrix supported Conglomerate, clast supported Massive, faint gradation Conglomerate, clast supported Conglomerate, clast supported Horizontal lamination, occasional imbrication Trough and planar cross-beds; average foreset thickness of 50 cm to 1 m Massive or faint lamination Sandstone, fine to coarse Sandstone, fine to very coarse, may be pebbly Massive, faint horizontal lamination, imbrication Trough cross-beds, 10–50 cm average scale Sandstone, fine to very coarse, may be pebbly Sandstone, fine to very coarse, may be pebbly Sandstone, very fine to coarse Sandstone, fine to medium, well sorted Sandstone, siltstone, mudstone Planar cross-beds Fsm/m Siltstone, mudstone Massive, desiccation cracks P Paleosol carbonate Pedogenic features: nodules, filaments Sh/l Sr Seod Fl 1534 Average Tamberia Fm. Source to the northeast (Puna margin) and east Source west of Northern transect Source west of Guanchin transect 20% Figure 7 (continued). (A) Clast count data collected along the Guanchin River and Northern transects; for location refer to Figures 2 and 3. Pc—clast count data; for location see Figure 2A. Mrl—marlstone; fs—fine sandstone; ms—middle sandstone; mcg—granule; cgl—conglomerate; e.g. 055 indicates sample code for ashes indicated in Figures 2B, 3B, and 4; PS—Punaschotter paleocurrent (pc) data indicated in Figure 3B. (B) Cumulative histograms for the clast-count data collected along different transects and from different sediment source lithologies. The relative abundance of source lithologies represents a volumetric estimation of those lithologies in the source area. The source lithologies are shown in rough stratigraphic order. High-strength (cohesive) debris flow High-strength or lowstrength (noncohesive) debris flow Longitudinal gravel bars, lag deposits Minor channel fills, transverse bed forms Distal debris flow Subaqueous 3D dunes, sustained unidirectional currents Subaqueous 2D dunes Horizontal laminations/low-angle (<15°) cross-beds Ripple cross-lamination Scour fills Large-scale cross-lamination (>25°) Fine laminations, very small ripples Eolian dunes Ripples (lower flow regime) Overbank, abandoned channel or waning flood deposits Overbank, abandoned or drape deposits Soil with chemical precipitation Geological Society of America Bulletin, November/December 2008 Dynamics of deformation and sedimentation in the northern Sierras Pampeanas deposits in small channels on the floodplain. Mudstone bodies are interpreted as pond-fill in shallow floodplain areas after flooding events or crevasse splays. The horizontally stratified conglomerates occurring at the top of this formation are likely the result of varying discharge in an alluvial-plain environment (Flint, 1985; DeCelles, 1991). Overall, we interpret the Guanchin Formation to be the result of individual flood events on an alluvial plain (Nemec and Steel, 1984; Miall, 1996). The coarsening- and thickening-upward trends expressed by the Guanchin rocks could be the result of a second event of progradation of the sedimentary system coupled with a further pulse of source tectonic activity that may have been related to eastward propagation of deformation, enhanced source exhumation, and resulting reduction of the distance between source and basin. Punaschotter Formation The Punaschotter Formation consists of medium- to coarse-grained conglomerates (Figs. 7A and 7B). The boundary with the underlying Guanchin Formation is both transitional and unconformable in places. The thickness of this unit is ~600 m. These conglomerates are mainly massive and disorganized; parallel layering and slight normal gradation occur locally. The conglomerates are mainly clast-supported and generally poorly sorted. The clasts are wellto subrounded and range in size from 5 cm up to 50 cm. Beds and clast assemblages are disorganized, and in some places, imbrications were observed and measured (Fig. 7A). Ash layers with thicknesses up to 2 m are commonly intercalated. We interpret these conglomerates to be representative of hyperconcentrated flows. This facies association was likely the result of highdensity stream flows (Todd, 1989), which are typically associated with proximal alluvial-fan environments (DeCelles et al., 1991). PROVENANCE Clast counts from the Tamberia, Guanchin, and Punaschotter conglomerates help constrain sediment source lithologies and spatiotemporal changes in the sediment source. New data from the Guanchin and Punaschotter Formations have been collected and integrated with existing data (Carrapa et al., 2006) in light of the new stratigraphic constraints. Clast-count analyses were performed at 20 localities along the Guanchin River transect and at 12 locations along the Northern transect (Figs. 7A and 7B). Pebbles of different lithologies were counted every 5–10 cm (depending on granulometry) within a 50 × 50 cm grid. The grid was shifted parallel to bedding until at least 100 clasts were counted at each locality. Paleocurrent directions were recorded in imbricated conglomerates by measuring 10–100 imbrications per location. These data were then restored to horizontal using the bedding plane and are reported as rose diagrams in Figure 7. Clast-Count Data The conglomerates of the Tamberia Formation, member B, are mainly composed of dacite, red sandstone, and minor andesite, granite, and quartzite clasts (Fig. 7A). These lithologies represent the Permian-Triassic red sandstone, Cambrian-Ordovician volcanics (Fig. 1B), and Paleozoic granites and granodiorites, all present in western bounding ranges (Fig. 7B). Dacite clasts are more abundant in the northwestern part of the basin than directly to the west; the upward increase in dacite clasts suggests a reorganization of the drainage system toward the northwest (Figs. 1B and 7B). Conglomerates of the Guanchin Formation are compositionally similar to Tamberia Formation conglomerates (member B), except that they also contain schist and phyllite clasts. Both conglomerates investigated along the Guanchin transect and those along the Northern transect are from more basinward (eastern) locations compared to the Tamberia locations, and they likely record more distal contributions (or a more extensive source). However, most of the conglomerates investigated along the Northern transect are proximal with respect to the western basin-bounding ranges, and therefore they are likely to record more local sources. Along the Guanchin River transect, the conglomerates record a strong contribution from andesite and granite lithologies and minor red sandstone, dacite, quartzite, aplite, and pelite lithologies; the latter document reworking from underlying mudstone layers within the same formation (Fig. 6C; photo 3). The up-section increase in granitic clasts may reflect a contribution from a reorganization of the sediment-source region to include granites present both to the north and to the east of the basin. Also, schists and phyllites, typical of northern and eastern sources, suggest a reorganization of the sediment source toward the north and northeast. This reflects exhumation and erosion of the Puna Plateau margin, which is also corroborated by apatite fission-track data (Carrapa et al., 2006). Two clast counts from the Northern transect record a major input from granite, phyllite, and limited red sandstone, biotite schist, and other minor components. The lower volcanic contribution in the Northern transect compared to the Guanchin River transect reflects the different abundance of such lithologies in the original sediment source in the late Miocene and highlights its local provenance signal. The composition of conglomerates of the Punaschotter Formation along both transects reflects a wider contribution of sources with respect to those recorded in the Tamberia Formation conglomerates, but it is similar to those in the Guanchin Formation. In particular, the Punaschotter conglomerates of the Guanchin River transect record a strong contribution from schist, phyllite, and gneiss, which is typical of northern and eastern sources (Puna-margin related). Distal facies, which crop out basinward (east), record a wider spectrum of lithologies, suggesting a larger drainage including western, northern, and eastern sources. This is also supported by our paleocurrent data (Fig. 7A). In order to evaluate the relative effects of unroofing versus reorganization of the source area, we compiled a clast-count diagram that shows the typical lithologies of different sediment sources (Fig. 7B). However, in a regional reconstruction such as this, some spatial information is lost. In the case of simple unroofing due to tectonic exhumation along a single structure, lithologies typical of shallower crustal levels should decrease and eventually disappear up-section, while lithologies typical of deeper crustal levels should appear at a certain stratigraphic level and increase up-section (Graham et al., 1986). This trend is not observed in Figure 7B, suggesting that a reorganization of the source to involve different and more abundant specific lithologies (i.e., schist, phyllite, and gneiss) was superimposed on simple initial tectonic unroofing. Taken together, the conglomerates along both transects document first the erosion of western bounding ranges and then a reorganization of the source area toward the north and northeast at the time of deposition of the upper Tamberia and Guanchin Formations, between ca. 8.2 Ma and ca. 5.2–5.5 Ma. This trend continued to the time of deposition of the Punaschotter Formation conglomerates, which record a contribution typical of western, northern, and eastern sources. STRUCTURAL DEFORMATION Structural investigations were conducted along the two mapped transects (Figs. 2 and 3); detailed structural analyses were carried out mainly along the Guanchin River transect where structures are better exposed (e.g., Figs. 8, 9, and 10). Fold axes and fault-plane lineations were used to reconstruct the kinematics of deformation. The kinematic analysis was Geological Society of America Bulletin, November/December 2008 1535 Carrapa et al. E W Ta mberia Ta mberia Tamberia (anticline) Punaschotter Guanchin 1m Tamberia (syncline) NE SW Ta mberia 1m Figure 8. Open folds in the Tamberia Formation; for locations, refer to Figures 2A and 2B. performed using the program FaultKinWin 1.1 (Allmendinger, 2002) following the method outlined in Marrett and Allmendinger (1990). This method allows the contouring of two pressuretemperature (P-T) axis distributions; these two populations are projected as P-T quadrants, similar to earthquake focal mechanisms (Fig. 2B). Furthermore, ubiquitous transgranular pebble and cobble fractures developed at contact points between neighboring clasts have been used as a proxy for the direction of tension and to infer the direction of compression (e.g., Eisbacher and Hopkins, 1977). The directions of transgranular fracture planes in clasts are represented in stereonets in Figure 11. These data were used ultimately to distinguish different phases of deformation. Next, we describe deformation along the Guanchin River transect, the Northern transect, and a separate location (NPc in Figs. 1B and 11) at the northern end of the Fiambalá basin. W Guanchin River Transect person for scale The westernmost reverse fault along the Guanchin River transect strikes NNW, dips 76ºSW, and places Permian basement rocks of the Paganzo Group (Figs. 1 and 2B) over the Tamberia Formation (conglomeratic facies of member B). Lineations plunge 71ºS, indicating dominantly dip-slip displacement and NE-SW– directed shortening (Fig. 2B). Along strike to the north, the same structure carries Ordovician metavolcanics and Carboniferous granitic rocks over Punaschotter conglomerates (Fig. 12). 1536 Figure 9. Tight fold in the Tamberia Formation; for location, refer to Figures 2A and 2B. Immediately east of the main bounding structure, the Tamberia Formation is overturned, and to the east, it is tightly folded into a series of anticline-syncline pairs (Fig. 2). Folds in the Tamberia Formation indicate ~E-W to NE-SW shortening (Figs. 8 and 9). The Tamberia For- mation is unconformably overlain by the Punaschotter Formation, which is folded into a gentle, slightly north-plunging syncline that has a N-S axis (Fig. 2B). To the east, a second reverse fault, which strikes N-S (Figs. 2A and 2B), places the Geological Society of America Bulletin, November/December 2008 Dynamics of deformation and sedimentation in the northern Sierras Pampeanas W Guanchin Fm. E 1m Figure 10. Example of open folds (syncline-anticline pair) in the Guanchin Formation; for location, refer to Figures 2A and 2B. Transgranular fractures in pebbles and cobbles Guanchin River transect F4: upper Punaschotter Northernmost Fiambalá Basin NPc: Punaschotter undifferentiated F3: Punaschotter n = 50 n = 78 F2: upper Guanchin Fm. n = 85 F1: lower Guanchin Fm. n = 83 n = 85 Figure 11. Transgranular pebble-cobble fractures constraining the orientation of tension from which the direction of shortening may be inferred. For location of the measurements, refer to Figures 1 and 2A. Geological Society of America Bulletin, November/December 2008 1537 Carrapa et al. person for scale Figure 12. Granitic basement carried (310/35°) over the Punaschotter Formation; view is to the NNE along the Northern transect. For location, refer to Figure 1. deformed Tamberia Formation (member A) to the west over the Guanchin Formation to the east. Minor faults associated with this structure were measured (Fig. 2B) and indicate NE-SW shortening. Ash sample 029 has an age of 5.51 ± 0.15 Ma and was collected just under this fault contact, indicating that deformation took place sometime after ca. 5.5 Ma. To the east of this fault, strata of both the Guanchin and Punaschotter Formations dip gently (~10º) to the east, and there appears to be a transitional contact, rather than an unconformity, between the two formations. Farther east, a thrust within the Guanchin Formation strikes NW-SE and dips shallowly (24º) to the west (Fig. 2B). Kinematic analysis of fault striae indicates ENE-WSW shortening (Fig. 2B). East of this thrust, the Guanchin Formation is folded into an open fold, indicating E-W to NE-SW shortening (Fig. 10). Farther east, the dip increases near the GuanchinTamberia contact. The Tamberia strata near this location appear to be tightly folded (Fig. 9). Although no structures are exposed in the eastern part of the section, two are suggested by the mapped relationships. The first would place strata of the Tamberia Formation on those of the Guanchin Formation, but it is covered by Punaschotter conglomerates. The second, to the east of the section, would account for open folds and back-tilting in the easternmost Guanchin rocks. Both are indicated with dashed lines in the cross 1538 section (Fig. 2B). The Punaschotter Formation in this region is deformed into a gentle syncline. It is in clear unconformity with the Tamberia Formation to the west, and while the eastern contact with the Guanchin Formation is not exposed, the dips are approximately the same in both units (~30ºW), indicating a possible conformable contact. Northern Transect and Northernmost Fiambalá Basin A similar style of deformation is observed along the Northern transect, though the amount of deformation within the basin is less (Figs. 3A and 3B). The major basin-bounding fault is present, but there are more structural complications in the basement hanging-wall rocks to the west. The Tamberia Formation is not exposed, and the Guanchin and Punaschotter Formations are folded but apparently not broken by faults along this transect (Fig. 3B). Along the west side of the basin, basement rocks are carried over late Miocene to Pliocene sedimentary rocks of the Fiambalá basin along the basin-bounding fault, which here is a NNWstriking reverse fault (Fig. 12). These basement rocks are internally deformed by several minor faults, the strikes of which vary from NNW to NNE and both east and west vergence. Fault zones within the granite commonly include fault breccias and gouge of strongly fractured and cataclasized granite, which varies in color from white to green to violet. Whereas most structures show evidence for thrust-sense displacement, striae on one basement fault indicate strike-slip (dip 6° toward 349° on a NNW-SSE– striking fault plane). Adjacent to the basin-bounding fault, the Punaschotter conglomerates are complexly, tightly folded, and in places overturned (Fig. 3A). To the east, the Guanchin and the Punaschotter Formations are folded into a broad, asymmetric anticline with a NNE-SSW–striking fold axis. Here, the Punaschotter-Guanchin contact is transitional. The eastern limb of the fold dips ~20°ESE, whereas the western limb is steeper, with dip angles around 65°WNW. This fold can either be interpreted as a west-vergent box fold or as a fault-propagation fold above a westdirected blind thrust (Fig. 3B and 3C). Due to the lack of subsurface data (e.g., seismic-reflection profiles), this question cannot be resolved. However, fault-propagation folds above westvergent blind faults in the Bermejo Valley to the south (Zapata and Allmendinger, 1997) exist in the region and suggest that similar structures could also be present in the Fiambalá basin. If the structure is a box fold, the décollement would be located within the basement at a depth of 10 km (Fig. 3B). A west-vergent propagating blind thrust below the fold could have a deeper décollement depending on the dip angle of the fault (Fig. 3C). Similar results for décollement depth (10–20 km), based on seismic-reflection profiles, have been reported in the Bermejo Valley (Zapata and Allmendinger, 1997; Allmendinger and Zapata, 2000; Jordan et al., 2001). A minimum horizontal shortening of 18% is calculated along the Northern transect. This value also agrees with estimates for the Bermejo basin (Allmendinger and Zapata, 2000). The orientation of fold axes generally indicates E-W contraction. Minor folds in the Punaschotter Formation along the Rio Colorado have fold axes that strike NW-SE, reflecting NE-SW shortening. The basin-bounding fault, which can be traced north from the Guanchin River and Northern transects, is a reverse fault that dips 75ºW in the northernmost part of the basin. This is steeper than along the Northern transect but similar to the dip along the Guanchin River transect. Only a few hundred meters of the top of the Guanchin Formation are exposed in this area, deposited directly on bedrock. In most places, a series of thick ignimbrite units is deposited directly on basement. The ignimbrites likely originated from the Cerro Blanco Caldera complex (Viramonte et al., 1994; Coira and Kay, 2004) on the Puna Plateau to the north. Although the ignimbrites are not dated here, elsewhere, Cerro Geological Society of America Bulletin, November/December 2008 Dynamics of deformation and sedimentation in the northern Sierras Pampeanas Blanco units have yielded 40Ar/39Ar ages of 0.2 ± 0.1 Ma (Siebel et al., 2001) and 0.55 ± 0.1 Ma (Seggiaro et al., 2008). The ignimbrite units are interbedded toward their top with Punaschotter Formation or facies-equivalent conglomerates. The ignimbrites and conglomerates are gently folded and are cut by two east-vergent reverse faults, which document NE-SW–directed shortening (Schoenbohm and Strecker, 2005). Younger N-S extension is accommodated along an E- to NE-striking normal fault to the north of the Fiambalá basin, which likely corresponds to a younger phase of deformation (late Pliocene to present) documented in many locations across the Central Andean Plateau (e.g., Allmendinger, 1986; Allmendinger et al., 1989; Marrett et al., 1994). Punaschotter Contact The base of the Punaschotter Formation is unconformable with the Tamberia Formation in every exposure (e.g., western and eastern ends of the Guanchin transect), indicating that structural deformation and erosion predate deposition of the Punaschotter conglomerates. The relationship between the Punaschotter and Guanchin Formations, however, is more complex. In the central part of the Guanchin transect (Fig. 2), and in the Northern transect (Fig. 3), the Punaschotter-Guanchin contact is transitional, indicating little or no deformation. To the east, we were not able to observe the contact directly, but dips are similar in the Guanchin Formation and overlying Punaschotter conglomerates (30ºW). The Punaschotter conglomerates that are transitional with the Guanchin Formation, along the Guanchin transect, are younger than 5.2 Ma and older than 3.7 Ma (Figs. 2 and 4). There are no stratigraphic constraints, however, for the unconformably overlying Punaschotter conglomerates that crop out just east of the main basin-bounding fault in the Guanchin transect (Fig. 2B). These relationships probably reflect a combination of propagation of deformation into the basin, and variation in the age of the base of the Punaschotter Formation. Deformation occurred both before and during deposition of the Punaschotter conglomerates. STRUCTURAL EVOLUTION OF THE FIAMBALÁ BASIN Faults and blind faults within the basin accommodate varying amounts of displacement and apparently terminate along strike without carrying on into adjacent regions. The style of deformation, however, is consistent in all study areas. Most structural indicators, such as fault and fold-axis orientations and striae on major and minor fault planes, document E-W to NE-SW contraction. Transgranular fractures in clasts document NE-SW and NW-SE tension (Fig. 11), which can be reconciled with fault kinematic indicators. The nature of the Punaschotter contact, overthrusting of Punaschotter conglomerate by the basin-bounding fault, and overthrusting of Guanchin ash-bearing strata indicate deformation from at least ca. 5.5 Ma to 3.7 Ma. Warped and offset fluvial terraces suggest protracted deformation to the present day. A regional kinematic shift in Pliocene-Quaternary time (Allmendinger, 1986; Allmendinger et al., 1989; Marrett et al., 1994) is documented to the far north of the Fiambalá basin, where a normal fault offsets basement and ignimbrite and accommodates N-S extension (Schoenbohm and Strecker, 2005). Changes in the tectonic stress field have been inferred from the southern Puna Plateau margin in the past 2–4 m.y. as a consequence of plateau uplift, lateral extrusion of crust from the plateau to the south (Allmendinger, 1986), late Cenozoic reorganization of South American plate motion (Marrett and Strecker, 2000), and gravitational extensional spreading (Schoenbohm and Strecker, 2005). No evidence of faulting along the east side of the Fiambalá basin has been detected. Thus, the Fiambalá basin seems to have been formed through progressive loading along reverse faults farther to the west (e.g., Chilean Precordillera; Fig. 1) at the time of Tamberia deposition, just west of the Fiambalá basin during Guanchin and Punaschotter deposition, and back-tilting of the Sierra de Fiambalá range along a fault(s) along its eastern edge. This is consistent with the structural geometry in other Puna-margin basins where the main active bounding structure loading the basin is located to the west—as deformation propagates eastward, it encounters inherited zones of crustal weakness, and it locks, enhancing both loading from the west and back-tilting from eastern and intrabasin structures (e.g., Mortimer et al., 2007). SEDIMENTARY EVOLUTION AT 28°S The sedimentary record preserved in the Fiambalá basin suggests that sedimentation and deformation occurred in a distal to proximal braided fluvial system and in alluvial fans within a regional to broken foreland-basin setting. The oldest sedimentary rocks preserved in the basin are ca. 9 Ma (Reynolds, 1987). However, the basin history may be significantly longer, as suggested by the presence of poorly documented Miocene (Fig. 1) and pre-Miocene sedimentary rocks preserved today to the west of the basin (Turner, 1967). Our provenance and structural data suggest that the basin received sediments and subsided in response to progressive tectonic activity and loading. At ca. 6 Ma, a reorganization of the paleodrainage system occurred to include northern and eastern source areas. The Punaschotter conglomerates record coarse alluvial-fan sedimentation characterized by a local drainage pattern, with sources both to the west and northeast. We interpret the reorganization of paleodrainage at ca. 6 Ma to have been due to the combination of the Puna Plateau growing southward and uplift of ranges east and north of the basin (Alto Grande; AD in Fig. 1B). The systematic coarsening upward of the investigated strata is consistent with progradation of the sedimentary system in response to eastward migration of deformation. The characteristic unconfined to poorly confined facies of the Neogene Fiambalá basin sequence suggest dry climate that prevented vegetation and enhanced bank instability and avulsion. We speculate that the Tamberia Formation strata represent the last stage of a more extensive foreland basin, as supported by the facies association that documents deposition in a large river system with source in ranges located farther to the west with respect to the Sierra de las Planchadas. Correlation with coeval strata along strike in the Bermejo basin supports this interpretation. Furthermore, sedimentation rates (Fig. 1C) appear to have been much higher during Tamberia deposition (>1 mm/yr) than during Guanchin and Punaschotter deposition (<1 mm/yr), which is consistent with a large drainage and regional foreland at the time of Tamberia deposition. The two cycles of coarsening (first in the Tamberia, then in the Guanchin Formation) are interpreted to reflect propagation of the sedimentary system related to growth and emplacement of two distinct thrust sheets. Numerous faults have been documented within and to the west of the Sierra de las Planchadas (Fig. 1B). However, it was beyond the scope of this study to further investigate the timing of emplacement of these thrust sheets. Collectively, our data suggest that the Fiambalá basin might have once formed a continuous, unbroken foreland with other Neogene basins within the present-day Sierras Pampeanas region. In the study area, this basin was disrupted at ca. 6 Ma (Fig. 13). Basin disruption was coeval with paleodrainage reorganization and establishment of the high relief along the southern Puna margin as indicated by apatite fission-track data (Carrapa et al., 2006b). Tertiary sedimentary rocks have been reported west of the study area (Turner, 1967). These sedimentary rocks document an older history of the Fiambalá basin that could date to Eocene-Oligocene time and would be in agreement with recent thermochronological data, which suggest that the present-day Geological Society of America Bulletin, November/December 2008 1539 Geological Society of America Bulletin, November/December 2008 SF 10-5 Ma (AFT m.a.) (12) 11-9 Ma (AFT m.a.) (12) 15-10 (AFT c.a.) (12) 10-5 Ma (AFT m.a.) (12) reverse fault activation and exhumation at ca.6 Ma (16) 50 100 150 200 250 Km 64°W 25-20 Ma exhumation age after burial inferred from AFT modeling pre-Miocene range exhumation (AFT age) Indirect evidence of foreland basin deposits covering present-day topographic highs: 15-20 Ma exhumation age after burial inferred from AFT modeling (m.a.) CQ 65°W 38-29 Ma(13) Sierra Pampeanas Bj SM SU Fb EC A (16) 66°W Ea s t e r n Co rdille ra SVZ: Sierra de Valle Fertil SF: Sierra Famatina SM: Sierra La Maz SU: Sierra Umango Bj: Bermejo foreland Basin Fb: Fiambalá Basin CQ: Corral Quemado EC: El Cajón A: Angastaco AFT c.a.: apatite fission-track cooling age AFT m.a.: inferred age of final exhumation from AFT modeling e.g., (1) see references in the text 0 creation of high relief by 6 Ma (14) ca. 25 Ma(15) Puna Plateau 67°W 4km 0 km 13.5Ma(2) 1 km 11Ma(2) 9.8Ma(2) 9Ma(2) 8Ma(2) 3 km 7Ma(2) 20 Ma possible older sedimentary rocks not preserved today (12, 14) 9 Ma(3) 1km 2km 8.2 Ma(3) 3km 4km 5.5 Ma(3) 3.7 Ma(3) Fiambalá (Fb) Basin Conglomerate Fluvial sandstone Eolian sandstone Floodplain mudstone 10.7 Ma(6) 5.7 Ma(7) Basins A EC CQ Fb S Bj N time 10 5 0 (8) (8) Paleocurrent direction average provenance (e.g. from NE to SW in this example) inferred from typical clast count composition regional deformation (continuous foreland basin) local, intrabasin deformation (broken foreland) e.g. (1); reference for chronostratigraphic ages and paleoflow indicators; refer to figure caption for details 1 km 9 Ma(4) 2 km 7.14 Ma(5) 3.66 Ma(5) 3 km 1 km El Cajon (EC) Basin Corral Quemado (CQ) Basin less constrained possible well constrained Spatio-temporal distribution for intrabasin thick-skinned deformation and sedimentation 17.5 Ma(1) 16 Ma(1) 1 km 14.5 Ma(1) 2 km 13.5 Ma(1) 13 Ma(1) 3 km 12.3 Ma(1) 12 Ma(1) 11 Ma(1) 5 km 4 km 5Ma(2) 5 km 3.7Ma(2) 2.5Ma(2) 6 km Bermejo (Bj) Basin thick-skinned deformation (TKD) thin-skinned deformation (TND) 1 km 13.4 Ma(9) 3 km 9.1 Ma(10) 5 km 6 km (11) (12) Angastaco (A) Basin Figure 13. Satellite image of the southern Central Andes and location of the Neogene basins discussed in the text. The stratigraphy of the Bermejo foreland basin is after Jordan et al. (2001), the Fiambalá basin stratigraphy is from the present work, Corral Quemado basin stratigraphy is after Muruaga (2001), El Cajon basin stratigraphy is after Mortimer et al. (2007), and Angastaco basin stratigraphy is after Coutand et al. (2006) and Carrapa et al. (2006a). Local intrabasin deformation in the Fiambalá, Corral Quemado, El Cajon, and Angastaco basins is inferred from activation of high-angle faults, basin-bounding ranges (both to the west and east), change in provenance, paleocurrent indicators, and the presence of unconformities. (1) Jordan et al. (2001) and Reynolds et al. (2001); (2) Johnson et al. (1986), Fernandez (1996), Jordan et al. (2001); (3) present work; (4) Sasso (1997); (5) Latorre et al. (1997); (6) Strecker (1987) and Strecker et al. (1989); (7) Bossi et al. (1997); (8) Mortimer et al. (2007); (9) Coutand et al. (2006); (10) Marshall et al. (1983); (11) Carrapa et al. (2006a); (12) Coughlin et al. (1998); (13) Coutand et al. (2001); (14) Carrapa et al. (2006b); (15) Carrapa et al. (2005); and (16) Deeken et al. (2006). 32°S 31°S 30°S 29°S 28°S 27°S 26°S 25°S Precordil lera 68°W TKD TND 69°W TKD 24°S 70°W F SV Precordillera TKD 1540 TKD 23°S Carrapa et al. Dynamics of deformation and sedimentation in the northern Sierras Pampeanas basin-bounding ranges were once covered by sedimentary units (>4 km thick) that are no longer preserved (Carrapa et al., 2006). Indeed, in addition to data presented in this study, thermochronological data from Sierras Pampeanas blocks, and related sedimentary rocks, suggest that these ranges were once covered by forelandbasin sediments (Coughlin et al., 1998). Regional isopach mapping (Fielding and Jordan, 1988; Re, 1995; Re et al., 2003) indicates that Sierras Pampeanas basement rocks, farther south, were buried beneath ~11 km of combined Upper Paleozoic to Neogene strata. In the northernmost Sierras Pampeanas Cumbres Calchaquíes and Sierra Aconquija blocks (Fig. 1), up to 1600 m of Cenozoic sediments were eroded after ca. 6 Ma (Sobel and Strecker, 2003). A scenario in which a continuous foreland basin is disrupted by thick-skinned deformation is also supported by independent stratigraphic evidence from the Corral Quemado basin (east of the Fiambalá basin; Fig. 1A). Here, Upper Miocene (ca. 12[?]–9 Ma) (Sasso, 1997) sedimentary rocks of the Hualfin, Las Arcas, and Chiquímil Formations are characterized by channels and high-energy flow facies that document axial flow and are derived from the Puna margin located to the northwest (e.g., Muruaga, 2001). The Andalhuala Formation (ca. 7 Ma; LaTorre et al., 1997) is characterized by more proximal fluvial facies (Muruaga, 2001). The Pliocene Corral Quemado Formation (3.53 ± 0.04 Ma; Butler et al., 1984) is composed of coarse alluvial-fan conglomerates that originated from proximal local sources (Fig. 13). These Pliocene rocks are overthrust by basinbounding basement rocks (Muruaga, 2001), documenting ongoing deformation after ca. 3.5 Ma. Overall, the Corral Quemado stratigraphy records a situation where proximal, intrabasinbounding ranges started to provide sediments to the basins sometime between 7 Ma and 3.5 Ma. Northeast of the Corral Quemado basin, Tertiary sedimentary basins, such as the El Cajon and Angastaco basins, record similar sedimentary-tectonic features, suggesting that sedimentation began in a continuous foreland-basin system that was later disrupted by thick-skinned deformation at ca. 6 Ma (Grier and Dallmeyer, 1990; Mortimer et al., 2007; Coutand et al., 2006) (Fig. 13). In particular, in the Angastaco basin (Eastern Cordillera), the oldest record of foreland-basin deposition, is represented by the Quebrada de los Colorados Formation (ca. 37 Ma; Carrapa et al., 2006; Hongn et al., 2007). This formation was deposited in a foredeep (Starck and Vergani, 1996) or “wedge-top”-like environment (Hongn et al., 2007), which was part of a contiguous and large foreland basin that extended along the southern Central Andes from Bolivia to Argentina (DeCelles and Horton, 2003; Carrapa et al., 2006; Carrapa and DeCelles, 2008). The El Cajón basin at ~27°S contains mainly middle Miocene to Pliocene sedimentary rocks (ca. 11 to <5.7 Ma; Strecker et al., 1989; Bossi et al., 2000, 2001). The tectonosedimentary record of the El Cajón basin documents a disruption of a foreland-basin system at ca. 6 Ma, when the basin was incorporated into the deformation front and subsequently disrupted (Mortimer et al., 2007) (Fig. 13). Similar settings exist in the present-day Quebrada del Toro and Quebrada de Humahuaca at ~24°30′S and 23°30′S, respectively (e.g., Walther et al., 1998; Reynolds et al., 2001; Hilley and Strecker, 2005; Strecker et al., 2007). In all of these locations, the intermontane basin stage, associated with tectonic compartmentalization of the formerly contiguous foreland, is recorded by deposition of coarse and locally sourced facies, similar to the ones recognized in the Guanchin and Punaschotter Formations in the Fiambalá basin. An extensive foreland-basin system has also been described south of the study area at 30°S (Zapata, 1998; Jordan et al., 2001). The Bermejo foreland basin was formed at ca. 20 Ma in response to shortening, loading, and fault tilting of western bounding ranges that constitutes the Eastern Precordillera (Zapata, 1998). The Bermejo basin is considered to have passed from a continuous to a broken foreland-basin setting with crystalline basement uplifts at ca. 6.5 Ma (Fig. 13; Jordan et al., 2001). This transition coincided with deformation and uplift of crystalline basement blocks within the basin in late Miocene and Pliocene time. In summary, there is evidence for both eastward propagation of deformation in the Fiambalá basin (two coarsening-upward sequences with appropriate, though unstudied, faults to the west), and disruption of a continuous and larger foreland basin at ca. 6 Ma. This basin was of regional importance and stretched from the Bermejo basin in the transition between the Sierras Pampeanas and the Precordillera in the south to the Eastern Cordillera in the north. Our new data from the Fiambalá basin bridge the gap between these two regions. The continuous foreland basin was disrupted throughout its extent at around 6 Ma by thick-skinned deformation through the uplift of crystalline basement blocks. The fact that very little to no direct record of a continuous foreland basin (>8–9 Ma) is preserved in the Fiambalá basin may suggest that exhumation was greater in the study area than to the south and to the north, where such a record still exists. This could be a result of the location of the basin at the southernmost end of the Puna Plateau, the growth of which might have enhanced exhumation of early foreland-basin strata. CONCLUSIONS 1. The investigated sedimentary rocks of the Fiambalá basin were deposited between 8.2 and 3.05 Ma. Deposition of the lower Tamberia Formation (8.2 Ma) occurred in the distal part of an arid ephemeral fluvial system that originated from an exhuming region located farther to the west with respect to the modern westernmost basin-bounding range. Sedimentary rocks of the upper Tamberia Formation were deposited in the proximal part of the same ephemeral fluvial system and record more proximal sediment sources. Overall, the Tamberia strata document deposition in a regional foreland-basin system. The sedimentary rocks of the Guanchin Formation (ca. 5.5 Ma) document sedimentation on an alluvial plain that had sediment sources located to the west and to the north-northeast in the southern Puna margin. Both the Tamberia and Guanchin Formations indicate coarseningand thickening-upward trends resulting from the progressive deformation and exhumation of east-verging reverse faults and subsequent decrease in distance between source and basin due to eastward fault propagation and sedimentary progradation. A reorganization of the drainage system, to include northern and northeastern sources (southern Puna margin), is recorded at ca. 5.5 Ma. The Punaschotter Formation consists of alluvial-fan conglomerates; by the time of deposition of these rocks (ca. <3.7 Ma), the basin was overfilled and received detritus from western, northern, and northeastern sources. The Guanchin and Punaschotter strata document deposition in a broken (thick-skinned) forelandbasin system characterized by local sources. 2. Structural analyses of folds, faults, and fault striae document E-W to NE-SW contraction during basin formation. Deformation occurred within the basin from at least 5.5 Ma to 3.7 Ma. The regionally recognized shift in kinematics from E-W or NE-SW shortening to N-S extension can be identified only in the far northern part of the Fiambalá basin, along the southern margin of the Puna Plateau. 3. Comparisons between the Fiambalá basin, the Bermejo foreland basin to the south, and basins in the transition between the Sierras Pampeanas and the Eastern Cordillera to the north suggest a similar structural-sedimentological evolution. In all three regions, thick-skinned, intrabasin deformation started at ca. 6 Ma. If the transition from thin- to thick-skinned deformation was indeed related to flat-slab subduction (e.g., Ramos et al., 2002; Siame et al., 2005), then our study suggests that flat-slab subduction is recorded quasi simultaneously along strike, from 26°S to 32°S, at ca. 6 Ma, rather than in a time-transgressive manner. Existing Geological Society of America Bulletin, November/December 2008 1541 Carrapa et al. thermochronological data suggest that older rocks, possibly deposited in a regional foreland basin, once covered the Sierras Pampeanas basement and began to be stripped off at ca. 20 Ma. We propose that the Fiambalá basin was once part of a larger, continuous (thin-skinned) foreland-basin system including areas to the north in the Eastern Cordillera and to the south in the Sierras Pampeanas. If true, such a regional foreland-basin system existed despite the differences in the geological details of these areas. The present-day Fiambalá basin is thus a snapshot of a longer basin history, related to a larger, contiguous older foreland basin that may have extended along the central southern Andes from north to south over a length of 1000 km. ACKNOWLEDGMENTS Marc Hendrix, Ken Ridgway, Jim Schmitt, Cari Johnson, and Karl Karlstrom are kindly acknowledged for their careful and constructive reviews. We thank Chiara Barbieri, Gregorio Araoz, and Christian Reinoso for field assistance, and Ricardo Alonso for fruitful scientific discussions. Estelle Mortimer provided invaluable scientific input. We are grateful to the Alexander von Humboldt Foundation, Deutsche Forschungsgemeinschaft (DFG) (project CA 481/5-1 and a Leibniz grant to M. Strecker) and the U.S. National Science Foundation (NSF; Tectonics; project EAR-0635630) for funding this research. REFERENCES CITED Allen, J.R.L., 1983, Studies in fluviatile sedimentation: Bars, bar-complexes and sandstone sheets (low sinuosity braided streams) in the Brownstones (L. Devonian), Welsh Border: Sedimentary Geology, v. 33, p. 237–293, doi: 10.1016/0037-0738(83)90076-3. Allmendinger, R.W., 1986, Tectonic development, southeastern border of the Puna Plateau, northwestern Argentine Andes: Geological Society of America Bulletin, v. 97, p. 1070–1082, doi: 10.1130/0016-7606(1986)97<1070 :TDSBOT>2.0.CO;2. Allmendinger, R.W., 2002, FaultKin program: Http://www. geo.cornell.edu/geology/faculty/RWA/programs.html. Allmendinger, R.W., and Zapata, T.R., 2000, The footwall ramp of the Subandean décollement, northernmost Argentina, from extended correlation of seismic reflection data: Tectonophysics, v. 321, no. 1, p. 37–55, doi: 10.1016/S0040-1951(00)00077-9. Allmendinger, R.W., Ramos, V.A., Jordan, T.E., Palma, M., and Isacks, B., 1983, Paleogeography and Andean structural geometry, northwest Argentina: Tectonics, v. 2, p. 1–16, doi: 10.1029/TC002i001p00001. Allmendinger, R. W., Strecker, M., Eremchuk, J. E., and Francis, P., 1989, Neotectonic deformation of the southern Puna plateau, NW Argentina: Journal of South American Earth Sciences, v. 2, p. 111–130. Arriagada, C., Cobbold, P.R., and Roperch, P., 2006, Salar de Atacama basin: A record of compressional tectonics in the central Andes since the mid-Cretaceous: Tectonics, v. 25, doi: 10.1029/2004TC001770. Bossi, G.E., Muruaga, C., Georgieff, S.M., Ahumada, A.L., Ibanez, L.M., and Vides, M.E., 1997, The Santa Maria–Hualfin Neogene basin of the Pampean Ranges: An example of mixed tectonic evolution, in Congresso Latino Americano de Sedimentologia, Porlamar, Venezuela, p. 97–104. Bossi, G.E., Vides, M.E., Ahumada, A.L., Georgieff, S.M., Muruaga, C.M., and Ibanez, L.M., 2000, Analisis de las paleocorrentes y de la varianza de los componentes a tres niveles, Neogeno del Valle del Cajon, Catamarca, 1542 Argentina: Asociatión Argentina de Sedimentología, v. 7, no. 1–2, p. 23–47. Bossi, G.E., Georgieff, S.M., Gavriloff, J.C., Ibanez, L.M., and Muruaga, C., 2001, Cenozoic evolution of the intramontane Santa Maria basin, Pampean Ranges, northwestern Argentina: Journal of South American Earth Sciences, v. 14, p. 725–734, doi: 10.1016/S0895-9811(01)00058-X. Butler, R.F., Marshall, L.G., Drake, R.E., and Curtis, G.H., 1984, Magnetic polarity stratigraphy and K-Ar dating of late Miocene and early Pliocene continental deposits, Catamarca Province, NW Argentina: The Journal of Geology, v. 92, p. 623–636. Caminos, R., 1979, Sierras Pampeanas Noroccidentales, Salta, Tucumán, Catamarca, La Rioja y San Juan; II Simposio de Geología Regional Argentina 1: Córdoba, Academia Nacional de Ciencias, p. 225–292. Carrapa, B., and DeCelles, P.G., 2008, Eocene exhumation and basin development in the Puna of northwestern Argentina: Tectonics, v. 27, TC1015, doi: 10.1029/2007TC002127, 2008. Carrapa, B., Adelmann, D., Hilley, G.E., Mortimer, E., Sobel, E., and Strecker, M.R., 2005, Oligocene range uplift and development of plateau morphology in the southern central Andes: Tectonics, v. 24, TC4011, doi: 10.1029/2004TC001762. Carrapa, B., DeCelles, P.G., Gerhels, G., Mortimer, E., and Strecker, M.R., 2006a, Early Tertiary exhumation, erosion, and sedimentation in the Central Andes: NW Argentina: Eos (Transactions, American Geophysical Union), v. 87, no. 52, abstract AGUFM.T31E.08. Carrapa, B., Sobel, E., and Strecker, M.R., 2006b, 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 and Planetary Science Letters, v. 247, p. 82–100, doi: 10.1016/j.epsl.2006.04.010. Coira, B., and Kay, S.M., 2004, Central Andean plateau ignimbrites in the Puna Back-Arc Cenozoic Volcanic Province (PBCV): Pucon, Chile, International Association of Volcanology and Chemistry of the Earth’s Interior General Assembly. Coughlin, T.J., O’Sullivan, P.B., Kohn, B.P., and Rodney, J.H., 1998, Apatite fission-track thermochronology of the Sierras Pampeanas, central western Argentina: Implications for the mechanism of plateau uplift in the Andes: Geology, v. 26, no. 11, p. 999–1002, doi: 10.1130/00917613(1998)026<0999:AFTTOT>2.3.CO;2. Coutand, I., Cobbold, P.R., de Urreiztieta, M., Gautier, P., Chauvin, A., Gapais, D., Rossello, E.A., and LòpezGamundí, O., 2001, Style and history of Andean deformation, Puna Plateau, northwestern Argentina: Tectonics, v. 20, p. 210–234. Coutand, I., Carrapa, B., Deeken, A., Schmitt, A.K., Sobel, E., and Strecker, M.R., 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 Research, v. 18, p. 1–26. DeCelles, P.G., and Horton, B.K., 2003, Early to middle Tertiary foreland basin development and the history of Andean crustal shortening in Bolivia: Geological Society of America Bulletin, v. 115, no. 1, p. 58–77, doi: 10.1 130/0016-7606(2003)115<0058:ETMTFB>2.0.CO;2. DeCelles, P.D., 1991, Controls on synorogenic alluvial-fan architecture, Beartooth conglomerate (Paleocene), Wyoming and Montana: Sedimentology, v. 38, p. 567–590, doi: 10.1111/j.1365-3091.1991.tb01009.x. DeCelles, P.G., Gray, M.B., Ridgway, K.D., Cole, R.B., Srivastava, P., Pequera, N., and Pivnik, D.A., 1991, Kinematic history of foreland uplift from Paleocene synorogenic conglomerate, Beartooth Range, Wyoming and Montana: Geological Society of America Bulletin, v. 103, p. 1458–1475, doi: 10.1130/0016-7606(1991)10 3<1458:KHOAFU>2.3.CO;2. Deeken, A., Sobel, E.R., Coutand, I., Haschke, M., Riller, U., and Strecker, M.R., 2006, Construction of the southern Eastern Cordillera, NW Argentina: From Early Cretaceous extension to middle Miocene shortening, constrained by AFT-thermochronometry: Tectonics, doi: 10.1029/2005TC001894, v. 25, p. 1–21. Dickinson, W.R., and Snyder, W.S., 1978, Plate tectonics of the Laramide orogeny, in Matthews, V., III, ed., Lar- amide Folding Associated with Basement Block Faulting in the Western United States: Geological Society of America Memoir 151, p. 355–366. Eisbacher, G.H., and Hopkins, S.L., 1977, Mid-Cenozoic paleogeomorphology and tectonic setting of the St. Elias Mountains, Yukon Territory: Geological Survey Canada Paper 77, no. 1B, p. 319–335. Fernandez, A.E., 1996, Seismic Analysis, Paleoclimatology and Fluvial Architecture of the Bermejo Basin, Central Andes, Western Argentina [Ph.D. thesis]: Ithaca, New York, Cornell University, 279 p. Fielding, E.J., and Jordan, T.E., 1988, Active deformation of the boundary between the Precordillera and Sierras Pampeanas, in Perry, W.J., and Schmidt, C.J., eds., Argentina and comparison with ancient Rocky Mountain deformation: Geological Society of American Memoir 171, p. 143-163. Flint, S., 1985, Alluvial fan and playa sedimentation in an Andean arid closed basin: The Pacencia Group, Antofagasta Province, Chile: Journal of the Geological Society of London, v. 142, p. 533–546, doi: 10.1144/ gsjgs.142.3.0533. González Bonorino, F., 1950, Algunos problemas geológico de las Sierras Pampeanas: Revista de la Asociación Geológica Argentina, v. 5, p. 81–110. Graham, S.A., Tolson, R.B., DeCelles, P.G., Ingersoll, R.V., Bargar, E., Caldwell, M., Cavazza, W., Edwards, D.P., Follo, M.F., Handschy, J.F., Lemke, L., Moxon, I., Rice, R., Smith, G.A., and White, J., 1986, Provenance modelling as a technique for analysing source terrane evolution and controls on foreland sedimentation, in Allen, P.A., and Homewood, P., eds, Foreland basins: International Association of Sedimentologists Special Publication 8, p. 425–436. Grier, M.E., and Dallmeyer, R.D., 1990, Age of the Payogastilla Group: Implications for foreland basin development, NW Argentina: Journal of South American Earth Sciences, v. 4, p. 269–278, doi: 10.1016/0895-9811(90)90008-O. Hein, F., and Walker, R.G., 1977, Bar evolution and development of stratification in the gravelly, braided Kicking Horse River, B.C.: Canadian Journal of Earth Sciences, v. 14, p. 562–570. Hilley, G.E., and Strecker, M.R., 2005, Processes of oscillatory basin filling and excavation in a tectonically active orogen: Quebrada del Toro Basin, NW Argentina: Geological Society of America Bulletin, v. 117, no. 7–8, p. 887–901, doi: 10.1130/B25602.1. Hongn, F., del Papa, C., Powell, J., Petrinovic, I., Mon, R., and Deraco, V., 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, v. 35, p. 271–274, doi: 10.1130/G23189A.1. Horton, B.K., 1998, Sediment accumulation on top of the Andean orogenic wedge: Oligocene to late Miocene basins of the Eastern Cordillera, southern Bolivia: Geological Society of America Bulletin, v. 110, no. 9, p. 1174–1192, doi: 10.1130/0016-7606(1998)110<1174: SAOTOT>2.3.CO;2. Horton, B.K., Hampton, B.A., and Waadners, G.L., 2001, Paleogene synorogenic sedimentation in the Altiplano Plateau and implications for initial mountain building in the Central Andes: Geological Society of America Bulletin, v. 113, p. 1387–1400, doi: 10.1130/0016-7606(20 01)113<1387:PSSITA>2.0.CO;2. Isacks, B., 1988, Uplift of the central Andean Plateau and bending of the Bolivian orocline: Journal of Geophysical Research, v. 93, p. 3211–3231, doi: 10.1029/ JB093iB04p03211. Johnson, N.M., Jordan, T.E., Johnsson, P.A., and Naeser, C.W., 1986, Magnetic polarity stratigraphy, age and tectonic setting of fluvial sediments in the eastern Andean foreland basin, San Juan Province, Argentina, in Allen, P.A., Homewood, P., eds., Foreland Basins: Ithaca, New York, International Association of Sedimentologists Special Publication 8, p. 63–75. Jordan, T.E., and Allmendinger, R.W., 1986, The Sierras Pampeanas of Argentina: A modern analogue of Laramide deformation: American Journal of Science, v. 286, p. 737–764. Jordan, T.E., and Alonso, R.N., 1987, Cenozoic Stratigraphy and Basin tectonics of the Andes Mountains, 20o-28o Geological Society of America Bulletin, November/December 2008 Dynamics of deformation and sedimentation in the northern Sierras Pampeanas South Latitude: American Association of Petroleum Geologists Bulletin, v. 71, p. 49–64. Jordan, T.E., and Mpodozis, C., 2006, Estratigrafía y evolución tectónica de la cuenca Paleógena Arizaro–Pocitos, Puna Occidental (24°–25°), in XI Congreso Geologico Chileno: Antofagasta, p. 57–60. Jordan, T.E., Isacks, B.L., Allmendinger, R.W., Brewer, J.A., Ramos, V.A., and Ando, C.J., 1983, Andean tectonics related to geometry of subducting Nazca Plate: Geological Society of America Bulletin, v. 94, p. 341-361. Jordan, T.E., Schlunegger, F., and Cardozo, N., 2001, Unsteady and spatially variable evolution of the Neogene Andean Bermejo foreland basin, Argentina: Journal of South American Earth Sciences, v. 14, p. 775–798, doi: 10.1016/S0895-9811(01)00072-4. Latorre, C., Quade, J., and McIntosh, W.C., 1997, The expansion of C-4 grasses and global change in the late Miocene: STable isotope evidence from the Americas: Earth and Planetary Science Letters, v. 146, p. 83–96, doi: 10.1016/S0012-821X(96)00231-2. Lunt, I.A., and Bridge, J.S., 2004, Evolution and deposits of a gravelly braid bar, Sagavanirktok River, Alaska: Sedimentology, v. 51, p. 415–432, doi: 10.1111/j.13653091.2004.00628.x. Makaske, B., 2001, Anastomosing rivers: A review of their classification, origin and sedimentary products: EarthScience Reviews, v. 53, no. 3–4, p. 149–196, doi: 10.1016/S0012-8252(00)00038-6. Marrett, R., and Allmendinger, R.W., 1990, Kinematic analysis of fault-slip data: Journal of Structural Geology, v. 12, p. 973–986. Marrett, R., and Strecker, M.R., 2000, Response of intracontinental deformation in the central Andes to late Cenozoic reorganization of South American plate motions: Tectonics, v. 19, no. 3, p. 452–467, doi: 10.1029/1999TC001102. Marrett, R., Allmendinger, R.W., Alonso, A.N., and Drake, R.E., 1994, Late Cenozoic tectonic evolution of the Puna plateau and adjacent foreland, northwestern Argentine Andes: Journal of South American Earth Sciences, v. 7, p. 179–207. Marshall, L.G., Hoffstetter, R., and Pascual, R., 1983, Mammals and stratigraphy: Geochronology of the continental mammal-bearing Tertiary of South America: Paleovertebrata, v. 93, p. 1–93. Martinez, L. del V., Nullo, F.E., Caminos, R.L., Panza, J.L., and Chipulina, M.A., 1995, Mapa Geologica de la Provincia de Catamarca: República Argentina. Secretaría de Minería: Buenos Aires, Dirección Nacional del Servicio Geológico, scale 1:500,000. McQuarrie, N., and DeCelles, P., 2001, Geometry and structural evolution of the central Andean backthrust belt, Bolivia: Tectonics, v. 20, p. 669–692, doi: 10.1029/2000TC001232. Miall, A.D., 1978, Lithofacies types and vertical profile models in braided river deposits: A summary, in Miall, A.D., ed., Fluvial Sedimentology: Canadian Society of Petroleum Geologists Memoir 5, p. 597–604. Miall, A.D., 1996, The Geology of Fluvial Deposits: New York, Springer-Verlag, 582 p. Miall, A.D., and Turner-Peterson, C.E., 1989, Variations in fluvial style in the Westwater Canyon Member, Morrison Formation (Jurassic), San Juan Basin, Colorado Plateau: Sedimentary Geology, v. 63, p. 21–60. Mon, R., and Salfity, J.A., 1995. Tectonic evolution of the Andes of Northern Argentina, in Tankard, A.J., Suárez, R., Welsink, H.J., eds., Petroleum basins of South America. Memoir, 62: American Association of Petroleum Geologists, p. 269–283. Mortimer, E., Coutand, I., Schoenbohm, L., Carrapa, B., Sosa Gomez, J., Sobel, E., and Strecker, M.R., 2007, Fragmentation of a foreland basin in response to out-ofsequence basement uplifts and structural reactivation: El Cajon–Campo del Arenal basin: Geological Society of America Bulletin, v. 119, p. 637–653, doi: 10.1130/ B25884.1. Muruaga, C., 2001, Estratigrafía y desarrollo tectosedimentario de sedimentos terciarios en los alrededores de la Sierra de Hualfín, borde suroriental de la Puna, Catamarca, Argentina: Revista de la Asociación Argentina de Sedimentología, v. 8, no. 1, p. 27–50. Nemec, W., and Steel, R.J., 1984, Alluvial and coastal conglomerates: Their significant features and some comments on gravelly mass-flow deposits, in Koster, E.H., and Steel, J.R., Sedimentology of Gravels and Conglomerates: Canadian Society of Petroleum Geologists Memoir 10, p. 1–31. Penk, W., 1920, Der Südrand der Puna de Atacama (NWArgentinien): Ein Beitrag zur Kenntnis des Andinen Gebirgstypus und der Frage der Gebirgsbildung: Der Abhandlungen der Sachsischen Akademie der Wissenchaften, v. 1, p. 3-420. Ramos, V.A., Cristallini, E.O., and Pérez, D.J., 2002, The Pampean flat-slab of the Central Andes: Journal of South American Earth Sciences, v. 15, p. 59–78, doi: 10.1016/ S0895-9811(02)00006-8. Re, G.H., 1995, Evolucion tectonosedimentaria del depocentro de la Cuenca del Antepais Andino (27° a 33°) y relacion con el cambio en el angulo de subduccion de la Placa de Nazca, in IX Congreso Latinoamericano de Geologa: Caracas, Venezuela, Ministerio de Energia y Minas, CD-ROM. Re, G.H., Jordan, T.E., and Kelley, S., 2003, Cronología y paleogeografía del Terciario de la cuenca intermontana de Iglesia septentrional, Andes de San Juan, Argentina: Revista Asociación Geológica Argentina, v. 58, p. 31–48. Reynolds, J.H., 1987, Chronology of Neogene Tectonics in Western Argentina (27°–33°S) Based on the Magnetic Polarity Stratigraphy of Foreland Basin Sediments [Ph.D. thesis]: Ithaca, New York, Cornell University, 353 p. Reynolds, J.H., Galli, C.I., Hernández, R.M., Idleman, B.D., Kotila, J.M., Hilliard, R.V., and Naeser, C.W., 2000, Middle Miocene tectonic development of the transition zone, Salta Province, northwest Argentina: Magnetic stratigraphy from the Metán Subgroup, Sierra de González: Geological Society of America Bulletin, v. 112, p. 1736–1751, doi: 10.1130/0016-7606(2000)11 2<1736:MMTDOT>2.0.CO;2. Reynolds, J.H., Hernandez, R.M., Galli, C.I., and Idleman, J.K., 2001, Magnetostratigraphy of the Quebrada La Porcelana section, Sierra de Ramos, Salta Province, Argentina: Age limits for the Neogene Oran Group and uplift of the southern Sierras Subandinas: Journal of South American Earth Sciences, v. 14, p. 681–692, doi: 10.1016/S0895-9811(01)00069-4. Sañudo-Wilhelmy, S.A., and Flegal, A.R., 1994, Temporal variations in lead concentrations and isotopic composition in the Southern California Bight: Geochimica and Cosmochimica Acta, v. 58, p. 3315–3320. Sasso, A., 1997, Geological Evolution and Metallogenic Relationships of the Farallon Negro Volcanic Complex, NW Argentina [Ph.D. thesis]: Kingston, Queens University, 842 p. Schoenbohm, L., and Strecker, M., 2005, Extension in the Puna-Altiplano Plateau since 1–2 Ma: Lithospheric delamination, lower crustal flow or gravitational collapse?: Geological Society of America Abstracts with Programs, v. 37, no. 7, p. 272. Seggiaro, R.E., Becchio, R., Schnurr, W., Adelmann, D., and Erpenstein, K., 2008, Hoja Geológica 2569-IV, Antofalla: Buenos Aires, Servicio Geológico Minero Argentino, scale 1:250,000 (in press). Siame, L.L., Bellier, O., Sébrier, M., and Araujo, M., 2005, Deformation partitioning in flat subduction setting: Case of the Andean foreland of western Argentina (28°S–33°S): Tectonics, v. 24, p. TC5003, doi: 10.1029/2005TC001787. Siebel, W., Schnurr, W., Hahne, K., Kraemer, B., Trumbull, R.B., van den Bogaard, P., and Emmermann, R., 2001, Geochemistry and isotope systematics of small- to medium-volume Neogene-Quaternary ignimbrites in the southern Central Andes: Evidence for derivation from andesitic magma sources: Chemical Geology, v. 171, p. 213–237, doi: 10.1016/S0009-2541(00)00249-7. Smith, J.D., 1970, Stability of a sand bed subjected to a shear flow of low Froude number: Journal of Geophysical Research, v. 75, p. 5928–5940, doi: 10.1029/ JC075i030p05928. Smith, N.D., and Perez-Arlucea, M., 1994, Fine-grained splay deposition in the avulsion belt of the lower Saskatchewan River, Canada: Journal of Sedimentary Research, v. B64, p. 159–168. Sobel, E., and Strecker, M.R., 2003, Uplift, exhumation and precipitation: Tectonic and climatic control of Late Cenozoic landscape evolution in the northern Sierras Pampeanas, Argentina: Basin Research, v. 15, p. 431-451. Starck, D., and Vergani, G., 1996, Desarrollo tecto-sedimentario del Cenozoico en el sur de la Provincia de SaltaArgentina, in Congreso Geológico Argentino: Buenos Aires, XII Congreso Geológico Argentine y III Congreso de Exploración de Hidrobcarbuos, Actas I, p. 433–452. Strecker, M.R., 1987, Late Cenozoic Landscape Development, the Santa Maria Valley, Northwest Argentina [Ph.D. thesis]: Ithaca, New York, Cornell University, 261 p. Strecker, M.R., Cerveny, P., Bloom, A.L., and Malizia, D., 1989, Late Cenozoic tectonism and landscape development in the foreland of the Andes: Northern Sierras Pampeanas (26°–28°), Argentina: Tectonics, v. 8, no. 3, p. 517–534, doi: 10.1029/TC008i003p00517. Strecker, M.R., Alonso, R.N., Bookhagen, B., Carrapa, B., Hilley, G.E., Sobel, E.R., and Trauth, M.H., 2007, Tectonics and climate of the southern Central Andes: Annual Review of Earth and Planetary Sciences, v. 35, p. 747–787, doi: 10.1146/annurev.earth.35.031306.140158. Tabutt, K.T., 1986, Fission Track Chronology of Foreland Basins in the Eastern Andes: Magmatic and Tectonic Implications [M.S. thesis]: Ithaca, New York, Cornell University, 100 p. Todd, S.P., 1989, Stream-driven, high density gravelly traction carpets: Possible deposits in the Trabeg Conglomerate Formation, SW Ireland, and some theoretical considerations of their origin: Sedimentology, v. 36, p. 513–530, doi: 10.1111/j.1365-3091.1989.tb02083.x. Turner, G., 1967, Descripcion geologica de la Hoja 13b, Chaschuil, Provincias de Catamarca y la Rioja: Dirección Nacional de Geología y Minería, v. 106, p. 1–78. Uba, C.E., Heubeck, C., and Hulka, C., 2005, Facies analysis and basin architecture of the Neogene Subandean synorogenic wedge, southern Bolivia: Sedimentary Geology, v. 180, p. 91–123, doi: 10.1016/j.sedgeo.2005.06.013. Vandervoort, D.S., Jordan, T.E., Zeitler, P.K., and Alonso, R.N., 1995, Chronology of internal drainage development and uplift, southern Puna Plateau, Argentine Central Andes: Geology, v. 23, p. 145–148, doi: 10.1130/009 1-7613(1995)023<0145:COIDDA>2.3.CO;2. Viramonte, J.G., Reynolds, J.H., Del Papa, C., and Disalvo, A., 1994, The Corte Blanco garnetiferous tuff: A distinctive late Miocene marker bed in northwestern Argentina applied to magnetic polarity stratigraphy in the Rıo Yácones, Salta Province: Earth and Planetary Science Letters, v. 121, p. 519–531, doi: 10.1016/0012-821X(94)90088-4. Walther, A.M., Orgeira, M.J., Reguero, M.A., Verzi, D.H., Vilas, J.F.A., Alonso, R., Gallado, E., Kelly, S., and Jordan, T., 1998, Estudio paléomagnetico, paleontológico y radiométrico de la Formación Uquía (Plio- Pleistoceno) en Esquina Blanca (Jujuy): Buenos Aires, Actas X Congreso Latinoamericano Geologico, v. 1, p. 77. Wooldridge, C.L., and Hickin, E.J., 2005, Radar architecture and evolution of channel bars in wandering gravel–bed rivers: Fraser and Squamish Rivers, British Columbia, Canada: Journal of Sedimentary Research, v. 75, p. 844–860, doi: 10.2110/jsr.2005.066. Zapata, T.R., 1998, Crustal structure of the Andean thrust front at 30°S latitude from shallow and deep seismic profiles, Argentina: Journal of South American Earth Sciences, v. 11, p. 131–151, doi: 10.1016/ S0895-9811(98)00006-6. Zapata, T.R., and Allmendinger, R.W., 1997, Evolución de la deformación del frente de corrimientos de la Precordillera, Provincia de San Juan: Revista Asociación Geológica Argentina, v. 52, no. 2, p. 115–131. MANUSCRIPT RECEIVED 22 SEPTEMBER 2006 REVISED MANUSCRIPT RECEIVED 12 JANUARY 2008 MANUSCRIPT ACCEPTED 12 FEBRUARY 2008 Printed in the USA Geological Society of America Bulletin, November/December 2008 1543