<<

Downloaded from gsabulletin.gsapubs.org on July 6, 2010

Oligocene–Miocene basin evolution in the northern , : Implications for evolution of the central Andean backthrust belt and high plateau

Bryan P. Murray1,†, Brian K. Horton2 , Ramiro Matos3, and Matthew T. Heizler4 1Department of Earth Science, University of California, Santa Barbara, Webb Hall, Santa Barbara, California 93106-9630, USA 2Institute for Geophysics and Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin, Austin, Texas 78712, USA 3Instituto de Investigaciones Geológicas y Medio Ambiente, Universidad Mayor de San Andrés, Casilla Postal 4787, , Bolivia 4New Mexico Bureau of Geology and Mineral Resources, Socorro, New Mexico 87801, USA

ABSTRACT sedimentation and fold-thrust deformation patterns in the central (e.g., Horton et al., in the frontal (west-southwestern) zone of the 2001; McQuarrie, 2002; DeCelles and Horton, The upper to lower central Andean backthrust belt was concen- 2003; McQuarrie et al., 2005; McQuarrie et al., and Aranjuez formations are exposed trated during late Oligocene–early Miocene 2008). Other studies have proposed that plateau in north-northwest–trending outcrop belts of time. These age results are consistent with uplift resulted from deeper processes such as the central Andean backthrust belt situated previous studies of east-derived sedimenta- thickening of thermally weakened lower crust within the central Andean plateau along the tion in the Altiplano and indicate regional by ductile horizontal shortening and magmatic boundary between the northern Altiplano uplift of the Eastern Cordillera at this time. addition (e.g., Isacks, 1988; Allmendinger et al., and the Eastern Cordillera of Bolivia. Sedi- Upsection trends in provenance data further 1997) or from rapid removal of a dense lower mentary lithofacies analyses indicate that suggest a progressively greater contribu- lithospheric root (e.g., Garzione et al., 2006; these coarse-grained siliciclastic formations tion from younger strata, possibly Ghosh et al., 2006; Garzione et al., 2008; Hoke were deposited primarily in alluvial fan to due to activation of new thrust faults dur- and Garzione, 2008). braided fl uvial environments. An upsection ing west-southwestward propagation of the The timing of crustal shortening in the change from principally fi ne-grained sand- backthrust belt toward the Altiplano. Such a central Andes remains relatively poorly con- stone to cobble conglomerate is consistent progression of late Oligocene–early Miocene strained. Previous workers have suggested with increased proximity to the sediment shortening along the Altiplano–Eastern Cor- initial deformation and possible early plateau source with time. Paleocurrent analyses re- dillera boundary likely refl ects signifi cant uplift between 30 and 25 Ma on the basis of a veal that fl ow was predominantly directed crustal thickening, potential isostatic uplift, late Oligocene to early Miocene infl ux of syn- toward the west-southwest away from and initial topographic expression of the east- orogenic sediment to the Altiplano and Andean Cordillera Real, the elevated core of the East- ern margin of the central Andean plateau. foreland basin (Isacks, 1988; Sempere et al., ern Cordillera. Provenance data from con- 1990; Gubbels et al., 1993; Allmendinger glomerate clast compositions and sandstone INTRODUCTION et al., 1997). In contrast, paleoelevation studies petrofacies suggest derivation from recycled utilizing stable oxygen isotope data from Neo- quartz-rich metasedimentary and sedimen- Located in the central portion of the gene basin fi ll have suggested rapid plateau up- tary rocks from the Paleozoic section in the ~8000-km-long Andean orogenic belt, the Alti- lift in late Miocene time (Garzione et al., 2006; Eastern Cordillera. The paleofl ow orienta- plano is one of the largest continental plateaus Ghosh et al., 2006; Garzione et al., 2008; Hoke tions, sediment provenance, and increased on Earth. The Andes are the topographic ex- and Garzione, 2008). Other studies focusing proximity of the sediment source suggest that pression of Cenozoic contractional stain associ- on thermochronological data (Benjamin et al., deposition of the Peñas and Aranjuez for- ated with along the western margin 1987; McBride et al., 1987; Farrar et al., 1988; mations was related to surface uplift of the of South America (Fig. 1; Isacks, 1988), yet Masek et al., 1994; Barnes et al., 2006; Gillis Eastern Cordillera relative to the Altiplano. the mechanisms for uplifting the central Andes et al., 2006; Ege et al., 2007; Barnes et al., Growth strata observed in the Aranjuez For- and Altiplano are not well understood. The high 2008; McQuarrie et al., 2008) and relict fore- mation further indicate that shortening was elevation of the central Andean plateau is iso- land basin deposits (Kennan et al., 1995; Lamb synchronous with deposition, probably in a statically supported by up to 70-km-thick crust and Hoke, 1997; Horton et al., 2001; DeCelles hinterland basin. New 40Ar/39Ar ages from (e.g., Beck et al., 1996). Some have suggested and Horton, 2003) have proposed initial Andean a lowermost exposed igneous unit and inter- that continuous eastward propagation of a fold- deformation as early as Late Cretaceous to bedded ash-fall tuff beds in the Aranjuez thrust belt, including a west-directed backthrust middle time. and Peñas formations show that synorogenic belt and east-directed basement-involved mega- Despite their differences, these varied thrust sheets, was the cause of crustal thicken- inter pretations for the mechanisms and tim- †E-mail: [email protected] ing, deformation, and observed topographic ing of shortening and plateau uplift share a

GSA Bulletin; September/October 2010; v. 122; no. 9/10; p. 1443–1462; doi: 10.1130/B30129.1; 14 fi gures; 4 tables; Data Repository item 2010105.

For permission to copy, contact [email protected] 1443 © 2010 Geological Society of America Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Murray et al.

Cordillera or Eastern Cordillera. The stratigra- phy of the Altiplano can be summarized as fol- Beni lows: (1) a 250–900-m-thick lower succession Plain 16° S of to mid- carbonate (El Molino Formation) and east-derived, distal- Fig. 2 fl uvial siltstone and sandstone (Santa Lucia

Subandean Formation); (2) a 3000–6500-m-thick middle Western Cordillera Eastern CordillCordillera succession of upper Eocene to Oligocene west- AltiplanoAltiplano derived fl uvial sandstone (Potoco Formation); (3) a 1000–4000-m-thick upper succession of ZZone Chaco one Plain poorly dated, partially Potoco-age–equivalent 20° S

e (Oligocene to middle Miocene), east-derived ra fl uvial and alluvial fan conglomerate (Coniri Formation) and numerous upper Oligocene to volcaniclastic units (Swanson et al., 1987; Sempere et al., 1990; Kennan et al., 1995; Lamb and Hoke, 1997; Horton et al., 2001; Horton et al., 2002; DeCelles and Hor- 22° S ton, 2003). Previous studies have related the 70° W 65° W 60° W observed stratigraphy to an eastward migration Figure 1. Map showing regional topography of the central Andes. of the Andean foreland basin system, with the Major physiographic divisions are separated by dashed lines (after El Molino and Santa Lucia formations depos- Isacks, 1988; McQuarrie et al., 2005). Darker colors (orange, red, ited in a backbulge depozone, the main body gray, black) defi ne regions of higher topography (>3.5 km). Box of the Potoco Formation recording foredeep shows location of Figure 2. deposition, and the upper Potoco, Coniri, and younger volcaniclastic formations representing intermontane (hinterland) basin deposition in the Altiplano (DeCelles and Giles, 1996; Hor- common basis in the consideration of sediment the Altiplano adjacent to the Eastern Cordillera ton et al., 2001; Horton et al., 2002; DeCelles accu mulation patterns within and adjacent to in order to determine the relationship of these and Horton, 2003). the central Andean plateau. For this reason, the deposits to upper-crustal deformation and topo- In the northeastern Altiplano, the Peñas and spatial and temporal trends in the confi guration graphic development along the margins of the Aranjuez formations form discontinuous out- of individual sedimentary basins, including the central Andean plateau. crops of coarse-grained siliciclastic deposits age, provenance, burial history, and relation- that trend north-northwest, roughly parallel to ships to upper-crustal structures, are important GEOLOGIC SETTING tectonic strike in the adjacent Eastern Cordillera to any models of deformation and uplift in the (Fig. 2). The basal contacts of these formations central Andes. The central Andes in Bolivia can be sub- are in angular unconformity above Devonian The present study of the middle Cenozoic divided into fi ve physiographically distinct strata. The Peñas Formation is up to ~1100 m Peñas and Aranjuez formations, siliciclastic provinces (from west to east): the Western Cor- thick and is best exposed near the town of Peñas deposits exposed in the northeastern Altiplano dillera, the Altiplano, the Eastern Cor dillera, (Fig. 3A) and at Cerro Santa Ana, north of the of Bolivia, helps constrain the history of ini- the Subandean Zone, and the Beni-Chaco Plain La Paz and metropolitan areas (Fig. 3B). tial hinter land basin development in the central (Fig. 1; Isacks, 1988). The Altiplano is an in- Numerous smaller outcrops of the Peñas For- Andean plateau and relative surface uplift along ternally drained, ~200-km-wide region of low mation with limited exposure are also present the plateau margin. Although previous stud- topographic relief located in the hinterland of in the region northwest of La Paz (Fig. 2). The ies have focused on synorogenic sedimentary the central Andean fold-thrust belt with an aver- Aranjuez Formation is >340 m thick and is ex- deposits exposed in the Altiplano and Eastern age elevation of 3.8 km. Bounding the Altiplano posed south of La Paz near Zona Sur, Pedregal Cordillera (e.g., Sempere et al., 1990; Kennan are the Eastern Cordillera, which represents the (Fig. 3C), Jankho Khaua, and east of Rio La et al., 1995; Lamb and Hoke, 1997; Horton, uplifted core of the Andean fold-thrust belt, and Paz (Fig. 2). The Peñas and Aranjuez forma- 1998; Horton et al., 2002; DeCelles and Hor- the Western Cordillera, the location of the active tions have been previously mapped as separate ton, 2003; Horton, 2005), there has been limited magmatic arc. Both of these systems rise to peak units based solely on outcrop location; however, research on the basin-fi ll exposures in closest elevations >6 km. Located farther east is the based on their proximity, similar litho facies, proximity to the Cordillera Real, the range with Subandean Zone, the frontal region of the active and presumed age, they are considered to be the highest peaks and greatest concentration of fold-thrust belt, and the Beni-Chaco Plain, the stratigraphic equivalents (e.g., Lohmann, 1970; thermochronological data in the Eastern Cor- modern retroarc foreland basin (Isacks, 1988; GEOBOL, 1995b) and also have been asso- dillera (Benjamin et al., 1987; McBride et al., Jordan, 1995; Horton and DeCelles, 1997; ciated with the and Salla formations 1987; Farrar et al., 1988; Gillis et al., 2006). Lamb and Hoke, 1997). located >100 km to the south-southeast (e.g., This study presents new results from an inves- The Altiplano basin contains a thick succes- Gillis et al., 2004, 2006). tigation of the sedimentology, provenance, and sion (up to 12 km) of regionally extensive Upper Potential sediment sources for the northern 40Ar/39Ar chronology of the Peñas and Aran- Cretaceous through Cenozoic sedimentary and Altiplano basin include principally Paleo- juez formations along the northeastern edge of volcanic rocks derived from either the Western zoic sedimentary rocks exposed in the central

1444 Geological Society of America Bulletin, September/October 2010 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Basin evolution in the northern Altiplano, Bolivia

Eastern Cordillera

Fig. 3A

′ Tres Rios 16°15 S Lago Titicaca

Peñas Fm. Fault

Fig. 3B Figure 2. Regional geologic map Altiplano of the La Paz (Bolivia) region El Alto along the Eastern Cordillera– La Paz Altiplano boundary showing Fig. 3C the generalized geologic units and major faults (simplifi ed Aranjuez Fm. from SERGEOMIN, 1997). Jankho Boxes show locations of Fig- Khaua ure 3. Asterisk denotes location Rio La Paz Fig. 8B of photograph in Figure 8B.

16°45′S N Coniri Fault

Coniri Fm.

20 km

68°30′W 68°00′W Miocene-Quaternary sedimentary rocks Upper Triassic granite

Upper Oligocene sedimentary rocks Paleozoic metasedimentary and sedimentary rocks

Andean backthrust belt of the easternmost Alti- DeCelles and Horton, 2003). A continuous Vila Vila Formation (700–1700 m thick), a plano and Eastern Cordillera and principally succession of Paleozoic marine deposits in- thickly bedded red to purple sandstone with Cenozoic igneous rocks of the Western Cordi- cludes: (1) the Ordovician (~2300-m- minor shale; (6) the Devonian Belén Formation llera. The Eastern Cordillera is a 200-km-wide thick) and Amutara (~800- to 3300-m-thick) (1000–1400 m thick), a unit of grayish-green range that creates a drainage divide between formations, low-grade metasedimentary units to brown shale and minor sandstone; (7) the the Altiplano and the river networks that fl ow composed of dark-gray to white quartzite, Devonian Sica Sica (400- to 600-m-thick) and eastward toward the South American . slate, and phyllite; (2) the Ordovician–Silurian Colpacucho (500- to 900-m-thick) formations, The Eastern Cordillera at 16° S–18° S is com- Cancañiri Formation (0–150 m thick), a poorly white to light-brown sandstone with inter- posed primarily of Ordovician–Carboniferous sorted, dark-gray to green diamictite with gran- bedded siltstone; and (8) the Carboniferous sedimentary and low-grade metasedimentary ite clasts and sandy mudstone; (3) the Silurian Copa cabana Formation (15–300 m thick), a rocks with local exposures of Cretaceous– Uncia Formation (850–1200 m thick), a unit of unit of limestone, sandstone, and siltstone that Cenozoic strata, Permo-Triassic and Neogene black to dark-brown shale and minor siltstone; occurs locally (GEOBOL, 1995c; Sempere, granitoids, and Jurassic–Cretaceous hypabys- (4) the Silurian Catavi Formation (500–800 m 1995; González et al., 1996; Suárez and Diaz, sal and volcanic fl ows (Gubbels et al., 1993; thick), a thinly interbedded, gray to green 1996; SERGEOMIN, 1997; McQuarrie and SERGEOMIN, 1997; Horton et al., 2002; sandstone and siltstone unit; (5) the Devonian DeCelles, 2001).

Geological Society of America Bulletin, September/October 2010 1445 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Murray et al. B Qa ed from Tps W ′ Db Reddish-purple sandstone Tan sandstone Quartzite Gray siltstone Qa 68°18 60 ed from GEOBOL, ed from 25.29 ± 0.04 Ma number of clasts 46 43 Ar age Tps 80 39 70 n = 30 Db Ar/ 40 Paleocurrent direction (trough cross-stratification, n = number of measured limbs) 41 12 Tpc Db CONGLOMERATE CLAST COMPOSITION CLAST CONGLOMERATE Qa 35 Tpc Tps 55 n = 64 Qa Tpc Location of measured section (Figure 4) 56 Belén Fm. Sica Fm. Aranjuez Fm. Peñas Fm. sandstone Muela del Diablo diorite Alluvium La Paz Fm. Peñas Fm. conglomerate 8 Tps 60 Tpc 10 n = 59 A Tv Db Ta Ds Tlp Qa 40 Tps Tpc W ′ 39 Tpc 1 km 29 68°22 Qa

N C

S 16°21 S 16°25 ′ ′ Ta D W ′ 22 AA Pedregal 68°03 Tv 81 Qa Qa Ta W 82 Tps 65 60 Tlp Ds 88 68°29 ′ 85 52 35 Zona Sur 44 B Tpc 56 25 Qa Ta Tpc Tps Ta Ta W 60 ′ Db Ds 65 68°05 Qa 60 Ta A Qa Ds 14.11 ± 0.07 Ma 14.11 C 55 Ta Peñas 66 70 Db Ta Tlp 25.31 ± 0.06 Ma Ds Ta 30 75 35 Tps Ds W 1 km Ds 1 km 68°31 ′ 30 Tlp

N N

S 16°33

S 16°34 ′

′ S 16°16 S 16°14 ′ ′ 1995a, 1995c). (B) Cerro Santa Ana region, showing structural relationships, conglomerate clast-count results, and paleocurrent directions for the Peñas Formation (modifi for directions and paleocurrent conglomerate clast-count results, showing structural relationships, Ana region, Santa 1995a, 1995c). (B) Cerro GEOBOL, 1995b). ed from (modifi Aranjuez Formation sections in the and location of measured showing structural relationships GEOBOL, 1995b). (C) South La Paz region, Figure 3. Geologic maps of study sites. (A) Peñas region, showing structural relationships and location of measured sections in the Peñas Formation (modifi sections in and location of measured showing structural relationships 3. Geologic maps of study sites. (A) Peñas region, Figure

1446 Geological Society of America Bulletin, September/October 2010 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Basin evolution in the northern Altiplano, Bolivia

Regionally, the Eastern Cordillera is a biver- Western Cordillera suggests that this region 1996). Upward fi ning and coarsening trends in gent fold-thrust belt with major décollement may have been the frontal zone of the Andean individual sandstone beds may have resulted horizons in the Coroico, Cancañiri, Uncia, and fold-thrust belt during Eocene time, prior to from lateral migration and abandonment of Belén formations (McQuarrie and DeCelles, defor mation in the Eastern Cordillera and Sub- channels and bars or from changes in streamfl ow 2001; McQuarrie, 2002). The eastern portion andean Zone (Horton et al., 2002; DeCelles and velocities during deposition (e.g., Miall, 1977; of the Eastern Cordillera consists primarily of Horton, 2003). Uba et al., 2005). Mudstones are interpreted to east-northeast–directed thrust faults that expose represent overbank deposition in adjacent fl ood- Silurian–Devonian strata (McQuarrie, 2002; SEDIMENTOLOGY plain settings that were later subjected to incision McQuarrie et al., 2008). The westernmost East- by sandy channel systems (Miall, 1977). ern Cordillera is a region of west-southwest– Four stratigraphic sections measured in Lithofacies Association 2: Cross-Stratifi ed directed thrust faults, referred to as the central the Peñas and Aranjuez formations provide the Sandstone with Interbedded Conglomerate Andean backthrust belt or Huarina fold-thrust basis for the following descriptions of sedi- belt (e.g., Roeder, 1988; Sempere et al., 1990; mentary lithofacies of individual bedsets and This lithofacies association occurs within McQuarrie and DeCelles, 2001). The central interpretations of depositional processes (Fig. 4 the intermediate ~500-m-thick section of the Andean backthrust belt extends from ~15° S and Table 1). The ten sedimentary lithofacies Peñas Formation and is composed primarily to 22° S and consists of Ordovician–Devonian are grouped into a distinctive collection of four of medium- to coarse-grained sandstones with strata with limited exposures of Carboniferous sedimentary lithofacies associations that are in- trace granules to cobbles (Table 2). Individual and Cretaceous rocks. Locally, frontal thrusts terpreted to refl ect processes in different deposi- sandstone beds vary in thickness from medium of the central Andean backthrust belt, includ- tional environments (Table 2). In the following to very thickly bedded (~0.1–1.5 m) and extend ing the Huarina and Coniri faults (Fig. 2), place interpretations, the terms distal, intermediate, or laterally for tens to a few hundred meters (Fig. Paleozoic strata over mid-Cenozoic Altiplano proximal refer to the relative proximity of the 5B). Amalgamated channel-fi ll structures with basin fi ll, and deform and uplift Cenozoic basin to the sediment source region. low-angle cross stratifi cation are common in strata such as the Peñas and Aranjuez forma- this facies association, suggesting multistory tions unconformably deposited on the hang- channels, possibly in a distributary system (Fig. Peñas Formation ing wall of these faults (Sempere et al., 1990; 5C). Interbedded with the sandstones are clast- SERGEOMIN, 1997). Based on the presence supported, subrounded to well-rounded granule of upper Oligocene to lower Miocene growth Two stratigraphic sections were measured in to cobble conglomerates that typically defi ne strata that are overlapped by slightly deformed the ~1100-m-thick Peñas Formation (Figs. 3A, internally stratifi ed lenses that are not laterally younger strata, most upper-crustal shortening 4A, and 4B). The base of the Peñas Formation extensive. Mudstone is limited in this facies in the Eastern Cordillera apparently ceased by sits in angular unconformity above the under- asso ciation and occurs as discontinuous lenses. ~20–10 Ma when deformation was transferred lying Devonian Belén Formation. The Peñas Deposition of this facies association is in- eastward into the Subandean Zone (Roeder, Formation constitutes a progradational coarsen- terpreted to have occurred in the intermediate 1988; Sempere et al., 1990; Gubbels et al., ing upward package that can be subdivided into zone of a streamfl ow-dominated alluvial fan 1993; Lamb and Hoke, 1997; McQuarrie and three different facies associations attributable to system, with most deposition taking place in DeCelles, 2001; Horton, 2005). depositional processes in the proximal, medial , sandy and gravelly channels (e.g., Miall, 1985, Located on the opposing side of the Altiplano, and distal portions of a streamfl ow-dominated 1996; Hampton and Horton, 2007). Most sand the Western Cordillera is a ~70-km-wide range alluvial fan system (Table 2) that may have was deposited in amalgamated channel sys- consisting of Miocene–Quaternary vol canoes passed downslope into a braided fl uvial system. tems in which individual channels were <1.5 m and arc-derived ignimbrite sheets (Pareja et al., thick. Horizontally and trough cross-stratifi ed 1978; Isacks, 1988; Reutter et al., 1994; Wörner Lithofacies Association 1: Cross-Stratifi ed conglomerate lenses are interpreted as gravel et al., 2000a; Horton et al., 2002). Underlying Sandstone and Interbedded Mudstone deposited in longitudinal bars or lag depos- the volcanic cover rocks, the Western Cordillera This lithofacies association (Table 2) com- its (e.g., Miall, 1985). The limited occurrence is composed primarily of Precambrian granitic prises the lower ~425 m of the Peñas Formation of overbank sediments and absence of debris and gneissic basement. Evidence for this rela- and is composed of fi ne- to medium-grained fl ow deposits suggests a depositional environ- tionship includes outcrops of dated Precam- sandstones interbedded with mudstones. Indi- ment situated in the channel-dominated inter- brian rocks beneath volcanic cover (e.g., Troëng vidual sandstone beds are thickly to very thickly mediate portion of an alluvial fan, downslope et al., 1994; Wörner et al., 2000b), granitic and bedded (~0.3–1.5 m), extend laterally for tens of any proximal region affected by sediment gneissic xenoliths in Neogene igneous rocks of meters, and have basal contacts that generally gravity fl ows and upslope of the distal region (e.g., Jiménez De Ríos, 1992), Precambrian form concave-up irregular erosional surfaces in characterized by mixed fl oodplain and channel basement rocks from drill-hole samples (e.g., which the sandstone units cut into underlying environments. Lehmann, 1978), and Precambrian granite and mudstone or sandstone beds (Fig. 5A). Interbed- gneiss clasts in west-derived middle to upper ded mudstone layers typically occur as lenses a Lithofacies Association 3: Conglomerate and Tertiary strata in the Altiplano (e.g., Jiménez few tens of centimeters in thickness, although Interbedded Cross-Stratifi ed Sandstone De Ríos, 1992; Horton et al., 2002). locally thicker mudstone layers (up to 18 m) This lithofacies association defi nes the upper Although the composition of the Western may extend laterally for tens of meters. ~200 m of the Peñas Formation and is distin- Cordillera can be inferred from available data, Based on the observed lithofacies, this facies guished by a larger average grain size (Table 2). the structural and tectonic history is relatively association is interpreted as the product of This lithofacies association is composed of unknown as a result of the Neogene volcanic deposition in the distal zone of a streamfl ow- clast-supported, moderately sorted, pebble to cover. Synorogenic sedimentation represented dominated alluvial fan or a moderate-energy, cobble conglomerates displaying imbrication by foreland basin deposits derived from the sandy braided fl uvial system (e.g., Miall, 1985, of subrounded to rounded clasts and erosive

Geological Society of America Bulletin, September/October 2010 1447 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Murray et al.

A Peñas N

8 900 n = 110 B Peñas S 16

4 n = 104

5 800 8 700 10 n = 79 14 n = 41

18 6 n = 104 31 700 600 n = 30 6 n = 168 6

10

11 8 600 500

16 6 10 6 6 n = 113 n = 108 8 4 10 500 12 20 400 n = 112 20 cl svff m c vc g p cb b 400 19 14

n = 322 12 16 9

9 16 n = 219 300 Figure 4 (on this and following page). 17 Measured stratigraphic sections of 12 300 20 Peñas Formation at (A) Peñas N and (B) Peñas S localities and Aran- 12 16 juez Formation at (C) Aranjuez and (D) Pedregal localities, including 13 facies types, paleocurrent data, con- 200 24 glomerate clast-count results, and 40Ar/39Ar age data. Stratigraphic 200 13 n = 110 9 correlations for the Peñas Forma- 13 tion are based on the lowest con- 8 glomerate lenses and correspond to 13 a change in facies associations. 100 13

100 14 16 12 Peñas Fm. Peñas Fm. 9 0 Belén Fm. cl s vf f m c vc g p cb b 5 (Devonian) Lower section 0 poorly exposed cl s vf f mcvc g p cb b

1448 Geological Society of America Bulletin, September/October 2010 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Basin evolution in the northern Altiplano, Bolivia

C Aranjuez EXPLANATION

LITHOLOGY (lithofacies code [Table 1])

25.31 ± 0.06 Ma Clast-supported conglomerate (Gh, Gt) 300 (BM27060601) Clast-supported conglomerate (Gcm, Gco)

Matrix-supported conglomerate (Gmm) n = 147 Low-angle cross-bedded sandstone (Sl)

Massive sandstone (Sm) 200 Horizontally bedded sandstone (Sh) n = 65 Sandstone with interbedded siltstone (Sh-Flm)

Mudstone (Flm) n = 49

100 n = 90 Ash-fall tuff

8 n = 105 No exposure/covered

n = 20 GRAIN SIZE cl - clay Aranjuez Fm. s - silt Vertical scale in meters vf - very fine sand 0 f - fine sand Base of section covered cl s vf f mcvc g p cb b m - medium sand 0 cl svff m c vc g p cb b c - coarse sand vc - very coarse sand g - granule p - pebble cb - cobble D Pedregal b - boulder Paleocurrent direction from trough cross limbs 12 (n = number of measured limbs) Paleocurrent direction from clast imbrication 21 (n = number of measured clasts)

200 25.31 ± 0.06 Ma 40Ar/39Ar age (BM27060601)

CONGLOMERATE CLAST COMPOSITION n = 120 Quartzite 11 Green sandstone Gray siltstone 100 n = 115 Green siltstone Gray sandstone n = 168 Reddish-purple sandstone n = 128 Purple siltstone Number of clasts Aranjuez Fm. Tan siltstone Tan sandstone 0 n = 121 Green carbonate Sica Sica Fm. cl s vf f m c vc g p cb b Andesite/dacite (Devonian)

Figure 4 (continued).

Geological Society of America Bulletin, September/October 2010 1449 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Murray et al.

TABLE 1. DESCRIPTION AND INTERPRETATION OF SEDIMENTARY Lithofacies Association 4: Conglomerate and LITHOFACIES (AFTER MIALL, 1985; UBA ET AL., 2005) Interbedded Sandstone and Mudstone Facies edoc noitpircseD noitaterpretnI The Aranjuez Formation consists primarily Gcm Clast-supported, massive conglomerate. Poorly sorted, Clast-rich debris fl ows of poorly sorted, massive, disorganized con- subrounded to angular, none to very poorly developed normal to glomerates (Fig. 6A) interbedded with mas- inverse grading, pebbles to boulders. Gco Clast-supported, organized conglomerate. Poorly sorted, Traction bedload sive to poorly stratifi ed pebbly sandstones and subrounded to angular. Normal grading, pebbles to boulders. sandy mudstones (Fig. 6B and Table 2). Indi- Gmm Matrix-supported, massive conglomerate. Poorly sorted, Plastic debris fl ows vidual conglomerate beds are clast- or matrix- subrounded to angular. No grading, pebbles to cobbles. Gh Clast-supported, crudely horizontally stratifi ed conglomerate. Longitudinal bedforms, lag supported , have lenticular to sheet-like geom- Moderately sorted, well rounded to subrounded. Granules to deposits etries, and are laterally extensive for tens of cobbles with imbrications. meters. Mudstone and sandstone beds are Gt Clast-supported, trough cross-stratifi ed conglomerate. Moderately Minor channel fi lls, transverse sorted, well rounded to subrounded, pebbles to boulders. bars massive to thinly bedded and typically con- Sl Low-angle (<15°) cross-stratifi ed sandstone. Very fi ne to coarse- Channel fi lls, crevasse splays, tain very fi ne grained sand- to pebble-rich grained, trace pebbles. Moderately to well sorted. dune migration Sh Horizontally laminated sandstone. Very fi ne- to medium-grained, Planar bed fl ow, upper fl ow lenses that are laterally continuous for several trace cobbles and pebbles. Moderately to well sorted. regime meters to tens of meters. Some sandstone and Sm Massive sandstone. Fine- to coarse-grained, trace cobbles and Hyperconcentrated sediment- mudstone sections contain primary to slightly pebbles. Moderately to well sorted. gravity fl ows Ss Sandstone with basal scour surface. Very coarse to medium- Erosive channel fi lls reworked ash-fall tuffs up to 6 m thick that dis- grained, trace granules. Normal grading. play sharp, nonerosive basal contacts (Fig. 6C). Flm Massive or laminated siltstone to claystone. Overbank, abandoned channel or This facies association represents deposition suspension deposits by high-energy sediment and fl uid gravity fl ows in the medial to proximal zone of a mass-fl ow– dominated alluvial fan (e.g., Blair and McPherson, basal contacts (Fig. 5D). Individual conglomer- absence of overbank mudstones and diminished 1994). Disorganized, clast- and matrix-supported ate beds are up to 0.5 m thick and persist lat- occurrence of sandstone channels are consis- conglomerates are interpreted as the products of erally for tens of meters. Interbedded with the tent with a proximal depositional environment clast-rich and plastic debris fl ows, respectively, conglomerates are medium- to coarse-grained upslope of the distal region that is characterized with minor massive sandstones representing sandstones with trace pebbles and cobbles. by greater sand and mud deposition in mixed hyperconcentrated sediment-gravity fl ows. The Sandstone beds are medium to very thickly bed- channel and fl oodplain settings. more organized, clast-supported conglomerate ded (~0.2 to >2 m) and extend laterally for tens facies represents traction bedload sedimentation to a few hundred meters. Aranjuez Formation during waning sheet fl oods in poorly confi ned This facies association is interpreted as the channels (e.g., Blair and McPherson, 1994). result of deposition in shallow gravelly channels Two stratigraphic sections were measured Mudstones to fi ne-grained sandstones are inter- in the proximal zone of an alluvial fan system in the >340-m-thick Aranjuez Formation preted as deposits in interchannel overbank fl ow (e.g., Miall, 1985; Miall, 1996; Hampton and (Figs. 3C, 4C, and 4D). The base of the regions (e.g., Miall, 1985; Blair and McPherson, Horton, 2007). The clast-supported conglomer- Aranjuez Formation lies in angular uncon- 1994) that were capable of preserving primary to ates generally exhibit lenticular geometries and formity above the Devonian Sica Sica For- slightly reworked ash-fall tuffs. erosive basal contacts consistent with deposition mation and younger, subhorizontal middle in shallow channels of low sinuosity (e.g., Miall, Miocene–Pliocene La Paz Formation uncon- Depositional Synthesis 1985). More poorly organized, clast-supported formably overlaps tilted beds of the Aranjuez conglomerates and massive (structureless) Formation. The facies association of this for- On the basis of the observed sedimentary sandstone beds may suggest a limited amount of mation represents deposition in a mass-fl ow– lithofacies (Table 1; Figs. 5 and 6), the Peñas hyperconcentrated sediment gravity fl ows. The dominated alluvial fan (Table 2). Formation is considered to represent deposition

TABLE 2. LITHOFACIES ASSOCIATIONS AND RELATED LITHOFACIES Thickness Facies association Lithofacies )m( scitsiretcarahC noitaterpretnI 1: Cross-stratifi ed Sl, Ss, Sh, ~425 Moderately to well-sorted, fi ne- to medium-grained sandstones, tan to Deposition in sandy channel systems and sandstone and Sm, Flm orange; interbedded dark-red siltstones muddy overbank settings of the distal zone interbedded of a streamfl ow-dominated alluvial fan or mudstone a moderate-energy, sandy braided fl uvial system 2: Cross-stratifi ed Sl, Sh, Sm, ~500 Moderately sorted, medium- to coarse-grained sandstones, white to Deposition in sandy and gravelly channel sandstone and Gh, Gt, Flm red, with trace granules to cobbles; interbedded clast-supported systems, with limited muddy overbanks interbedded granule to cobble conglomerate lenses up to ~1 m thick, subrounded in the intermediate zone of a streamfl ow- conglomerate to rounded, clast imbrication; trace mudstone lenses dominated alluvial fan 3: Conglomerate and Gh, Gt, Sm, ~200 Moderately sorted, clast-supported pebble to cobble conglomerates, Deposition in shallow gravelly channel interbedded cross- Sh, Sl subrounded to rounded, clast imbrication, erosive basal contacts; systems in the proximal zone of a stratifi ed sandstone interbedded moderately sorted, medium- to coarse-grained streamfl ow-dominated alluvial fan sandstones, white to red, with trace pebbles and cobbles 4: Conglomerate Gcm, Gco, >340 Poorly sorted, massive, disorganized conglomerates, subrounded to High-energy, sediment and fl uid gravity and interbedded Gmm, Sh, angular, clast or matrix supported, may have very poorly developed debris fl ows in the intermediate to proximal sandstone and Sm, Flm normal or inverse grading; interbedded massive to thinly bedded zone of a mass-fl ow–dominated alluvial mudstone pebbly sandstones, sandy mudstones, and very thinly to thinly fan bedded fi ne-grained sandstones, tan to purple

1450 Geological Society of America Bulletin, September/October 2010 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Basin evolution in the northern Altiplano, Bolivia

A B

D

C

Figure 5. Photographs of representative lithofacies in the Peñas Formation. (A) Interbedded sandstone and mudstone of lithofacies associa- tion 1, representing deposition in the distal zone of a streamfl ow-dominated alluvial fan or braided fl uvial system. Roughly downdip view includes sandstones with basal scour surfaces (Ss) cutting into underlying mudstone (Flm) in center of photo and sandstone (Sl) in base of photo (arrows). Rock hammer for scale. (B) Representative exposure of medium to very thickly bedded, lateral continuous sandstone of lithofacies association 2, interpreted as deposition in an intermediate zone of a streamfl ow-dominated alluvial fan. Seated person (out- lined) for scale. (C) Stacked lenticular channel geometries observed in low-angle cross-stratifi ed sandstone (Sl) of the intermediate zone of a streamfl ow-dominated alluvial fan (lithofacies association 2). Compass for scale. (D) Trough cross-stratifi ed cobble conglomerate (Gt) displaying channel-fi ll geometry and erosive base (arrows), and interbedded sandstone (Sm, Sh) in lithofacies association 3, representing deposition in the proximal zone of a streamfl ow-dominated alluvial fan. Rock hammer for scale.

Geological Society of America Bulletin, September/October 2010 1451 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Murray et al.

A B

C

Figure 6. Photographs of representative lithofacies in the Aranjuez Formation. (A) Clast-supported disorganized conglomerate (Gcm) with subrounded to subangular, pebble to boulder clasts (top to left). Rock hammer for scale. (B) Clast-supported pebble to cobble conglomer- ate (Gco) with inverse to normal grading, and interbed- ded massive sandstone (Sm) and mudstone (Flm). Person for scale. (C) Two interbedded ash-fall tuff layers (white) with sharp basal contacts (arrows) within a section of massive to thinly bedded sandy mudstone. Lower tuff layer is ~6 m thick.

in a streamfl ow-dominated alluvial fan (Table 2) hibit confi ned fl ow in proximal environments association 3) contains predominantly conglom- that may have constituted part of a terminal fan that pass downslope into a distal region of un- erate and is distinguished by a channelized peb- system (e.g., Friend, 1978; Nichols , 1987; Kelly confi ned fl ow (Kelly and Olsen, 1993; Hamp- ble to cobble conglomerate facies (Fig. 5D) and and Olsen, 1993; Ridgway and DeCelles , 1993; ton and Horton, 2007). The apparent absence of an absence of mudstone facies. In comparison, Newell et al., 1999; Smith, 2000) adjacent to major fl uvial systems downslope of the alluvial channelized sandstone with thin granule to peb- a braided fl uvial system farther downslope. A fan facies associations in the Peñas Formation ble conglomerate lenses (Figs. 5B and 5C) are terminal fan commonly contains a fan-shaped, may suggest a terminal fan setting, partially the typical signatures of the medial zone of the distributary network of fl uvial channels in an analogous to fl uvial megafans of the modern streamfl ow-dominated alluvial fan (litho facies arid to semiarid environment where water in- foreland basin adjacent to the central Andes association 2). Both the medial and proximal fi ltrates and evaporates before escaping the (e.g., Horton and DeCelles, 2001; Leier et al., zones of the streamfl ow-dominated alluvial fan system by surface fl ow (Friend, 1978; Nichols, 2005; Uba et al., 2005). lithofacies associations are distinguishable from 1987; Kelly and Olsen, 1993; Newell et al., The three lithofacies associations of the Peñas either a distal alluvial fan or braided fl uvial en- 1999). Terminal fans tend to be sand dominated Formation can be distinguished from each other vironment (lithofacies association 1). Whereas and have a lower gradient in comparison to on the basis of grain sizes and sedimentary the proximal and medial lithofacies associations typical coarse-grained alluvial fans (Kelly and lithofacies (Table 2). The proximal zone of the include substantial conglomerate lithofacies, Olsen, 1993). Terminal fans also generally ex- streamfl ow-dominated alluvial fan (lithofacies the distal lithofacies association is composed

1452 Geological Society of America Bulletin, September/October 2010 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Basin evolution in the northern Altiplano, Bolivia of only medium-grained sandstone to siltstone TABLE 3. SUMMARY OF 40AR/39AR RESULTS (Fig. 5A). Additionally, the distal alluvial fan Sample Formation (locality) Lithology n MSWD Age (Ma) ±2σ (Ma) lithofacies association contains thick units BM29060605 Aranjuez Fm. (Jankho Khaua) Trachyte sill 13 1.8 27.42 0.09 BM29060602 Aranjuez Fm. (Jankho Khaua) Pyroclastic fl ow 14 1.0 27.41 0.06 (up to 18 m) of fl oodplain mudstone, whereas BM29060601 Aranjuez Fm. (Jankho Khaua) Trachyte sill 9 1.5 27.40 0.06 mudstone is limited to absent in the medial and BM29060603 Aranjuez Fm. (Jankho Khaua) Trachyte sill 9 1.0 27.34 0.07 BM29060604 Aranjuez Fm. (Jankho Khaua) Pyroclastic fl ow 9 0.8 27.31 0.07 proximal zones of the alluvial fan lithofacies BM27060601 Aranjuez Fm. (Aranjuez) Tuff 15 1.1 25.31 0.06 associations. The Aranjuez Formation is in- 6061402 Peñas Fm. (Cerro Santa Ana) Tuff 25 1.1 25.29 0.04 terpreted to represent deposition in the medial BM28060601 La Paz Fm. (Zona Sur) Tuff 12 0.6 14.11 0.07 to proximal zones of a mass-fl ow–dominated Note: n—number of sanidine grains analyzed; MSWD—mean square of weighted deviates. allu vial fan (e.g., Blair and McPherson, 1994). Deposits of coarse-grained debris fl ows and re- lated lithofacies in a distributary pattern charac- tion of the Aranjuez Formation (Matos, 2002) than the lowest levels in the west (~25.3 Ma age terize mass-fl ow–dominated alluvial fans (Blair yield nearly indistinguishable weighted mean for the lowermost exposed Peñas Formation), and McPherson, 1994; Miall, 1996). The mass- ages ranging from a maximum age of 27.42 ± we infer a general west-southwestward migra- fl ow–dominated alluvial fan facies association 0.09 Ma to a minimum age of 27.31 ± 0.07 Ma tion of the locus of basin subsidence, comparable of the Aranjuez Formation can be distinguished (Figs. 7A–7E; Table 3). An ash-fall tuff (sample in size and scale to the progradation of the al- from the streamfl ow-dominated alluvial fan BM27060601) from near the top of the measured luvial fan depositional systems toward the Alti- lithofacies associations of the Peñas Forma- section of the Aranjuez Formation (Figs. 3C plano. Moreover, growth structures (discussed tion by differences in types of lithofacies, grain and 4) has a weighted mean age of 25.31 ± below) observed in the Aranjuez Formation size, and degree of clast roundness. Whereas 0.06 Ma (Fig. 7F and Table 3). A tuff (sample indicate that shortening was contemporaneous the Aranjuez Formation includes disorganized 6061402) from the lowest section of the Peñas with sedimentation in the Altiplano at this time. matrix-supported conglomerates (Fig. 6A) diag- Formation exposed at Cerro Santa Ana (Fig. 3B) nostic of debris fl ows, lithofacies of the Peñas has an identical weighted mean age, within error , SYNDEPOSITIONAL AND Formation include stratifi ed, channelized con- of 25.29 ± 0.04 Ma (Fig. 7G and Table 3). The POSTDEPOSITIONAL STRUCTURES glomerates and sandstones with interbedded youngest dated tuff (sample BM28060601), overbank mudstones indicative of streamfl ow from the unconformably overlying La Paz For- Structural relationships observed in the Peñas processes. Additionally, the conglomerates of mation (Fig. 3C), has a weighted mean age of and Aranjuez formations suggest that shortening the Aranjuez Formation contain angular to sub- 14.11 ± 0.07 Ma (Fig. 7H and Table 3). occurred both during and after their deposition. rounded pebble to boulder clasts, whereas con- These new 40Ar/39Ar age results are generally Both the Peñas and Aranjuez formations are ex- glomerates of the Peñas Formation consist of consistent with 40Ar/39Ar step-heating results for posed in steeply dipping panels in close prox- well rounded to subrounded pebbles to cobbles. biotite grains obtained by Gillis et al. (2006) imity to west-southwest–directed thrust faults from reworked tuffs and tuff clasts in the Aran- of the central Andean backthrust belt (Figs. 2 DEPOSITIONAL AGE CONSTRAINTS juez Formation, which have late Oligocene ages and 3). In most localities, the formations occur ranging from 27.4 ± 1.1 Ma to 25.1 ± 1.1 Ma in the footwalls of thrust faults that are, in turn, New 40Ar/39Ar results for an igneous unit (2σ). Gillis et al. (2006) also reported a single carried in the hanging wall of additional thrust from the lowermost exposed basin fi ll and inter- early Oligocene age of 33.1 ± 1.6 Ma (their faults farther to the west-southwest, notably the bedded ash-fall tuffs provide constraints on the sample Taa06) for the Aranjuez Formation. Huarina thrust (Fig. 2). In this structural setting, depositional age of the Peñas and Aranjuez for- However, this sample of reworked tuff is located the Peñas and Aranjuez formations were sub- mations in the northeastern Altiplano ( Table 3 only ~40 m downsection of a sample yielding jected to signifi cant postdepositional deforma- and GSA Data Repository Table DR1)1. an age of 26.7 Ma ± 1.1 Ma (Gillis et al., 2004), tion. For example, a moderate to high degree 40Ar/39Ar laser total-fusion analyses were per- which would suggest an anomalously low accu- of internal deformation involving small-scale formed at the New Mexico Geochronology Re- mulation rate (<10 m/Myr) within the Altiplano folding and intraformational thrust faulting af- search Laboratory on a collection of individual hinterland, where Cenozoic accumulation rates fected the Aranjuez succession (Fig. 8A) in the sanidine crystals separated from eight samples. commonly exceed 500 m/Myr (Horton et al., vicinity of the tuff dated at 25.31 ± 0.06 Ma For each sample, weighted mean ages and 2σ 2001; Hampton and Horton, 2007). Given this (Fig. 7F and Table 3). However, the unconform- errors are calculated on the basis of 40Ar/39Ar contradiction, the sample’s relatively higher age ably overlying middle to upper Miocene La Paz results for 9 to 25 individual sanidine crystals errors, and the reworked nature of the tuff, we Formation, including a tuff-bearing section (Fig. 7 and Table 3). regard the early Oligocene 40Ar/39Ar age for this yielding an age of 14.11 ± 0.07 Ma (Fig. 7H The 40Ar/39Ar ages for the eight samples in this sample as an estimate of the maximum possible and Table 3), has undergone little to no defor- study range from late Oligocene to middle Mio- depositional age for the Aranjuez Formation, mation. These relationships indicate that post- cene. In the Jankho Khaua region (Fig. 2), fi ve with the late Oligocene representing the princi- depositional deformation affecting the Aranjuez samples (samples BM29060601–BM29060605) pal age of Aranjuez deposition. and Peñas formations occurred during the late from a hypabyssal trachyte sill and pyroclastic Based on the 40Ar/39Ar results presented here Oligocene to early Miocene, likely between fl ow contained within a steeply dipping sec- and by Gillis et al. (2006), deposition of the ~25 Ma and 15 Ma. Peñas and Aranjuez formations was clearly on- In addition to postdepositional structures, evi- 1GSA Data Repository item 2010105, 40Ar/39Ar going during the late Oligocene, with the princi- dence of growth strata suggests syndepositional analytical methods results (Table DR1), and sum- pal depositional ages ranging from ~27.5 Ma to shortening was active during late Oligocene mary of recalculated modal point-count data for sandstones (Table DR2), is available at http:// ~25 Ma. Because the lowest exposed stratigraphic deposition of the Aranjuez and Peñas formations. www.geosociety.org/pubs/ft2010.htm or by request to levels in the east (~27.4 Ma age for lower most Although not observed in the Peñas Formation, [email protected]. exposed Aranjuez Formation) are ~2 Myr older potentially due to insuffi cient exposure, the

Geological Society of America Bulletin, September/October 2010 1453 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Murray et al.

28.0 28.0 A BM29060605 (27.42 ± 0.09 Ma) E BM29060604 (27.31 ± 0.07 Ma)

27.0 27.0

26.0 26.0 Age (Ma)

25.0 25.0

24.0 24.0 28.0 28.0 B BM29060602 (27.41 ± 0.06 Ma) F BM27060601 (25.31 ± 0.06 Ma)

27.0 27.0

26.0 26.0 Age (Ma)

25.0 25.0

24.0 24.0 28.0 28.0 C BM29060601 (27.40 ± 0.06 Ma) G 6061402 (25.29 ± 0.04 Ma)

27.0 27.0

26.0 26.0 Age (Ma)

25.0 25.0

24.0 24.0 28.0 16.0 D BM29060603 (27.34 ± 0.07 Ma) H BM28060601 (14.11 ± 0.07 Ma)

27.0 15.0

26.0 14.0 Age (Ma)

25.0 13.0

24.0 12.0

Figure 7. Plots depicting laser total-fusion 40Ar/39Ar ages for individual sanidine crystals from eight samples (A–H). In each plot, the horizon- tal line represents the calculated weighted mean age (2σ) for each sample. Vertical lines represent 1σ error bars for individual crystal ages. (A–E) Samples BM29060601–BM29060605: igneous unit containing hypabyssal trachyte sill and pyroclastic fl ow in lowermost Aranjuez Formation, near Jankho Khaua (Fig. 2). (F) Sample BM27060601: tuff from uppermost Aranjuez Formation (Figs. 3C and 4). (G) Sample 6061402: tuff from lowest exposed Peñas Formation at Cerro Santa Ana (Fig. 3B). (H) Sample BM28060601: tuff from the La Paz Formation that unconformably overlies the Aranjuez Formation (Fig. 3C). See Table DR1 [see footnote 1] for individual 40Ar/39Ar age analyses.

Aranjuez Formation displays several indicators temporaneous with late Oligocene deposition of SEDIMENT PROVENANCE of structural growth during sedimentation. Along the Aranjuez Formation at ~27.5 Ma to 25 Ma, the Rio Khellkata river-cut exposures east of Rio although shortening likely persisted into early Sediment Dispersal Patterns La Paz (Fig. 8B), fanning bedding dips, abrupt Miocene time. Although there is no direct evi- lateral thickness variations, and internal angu- dence for the precise onset of deformation in the Paleocurrent measurements were collected lar unconformities are observed in the Aranjuez study area, it is conceivable that local shortening from the Peñas and Aranjuez formations from Formation. The presence of these growth struc- commenced prior to late Oligocene deposition of trough-cross limbs in sandstone and conglom- tures suggests that shortening was partially con- the Aranjuez Formation. erates (method I of DeCelles et al., 1983) and

1454 Geological Society of America Bulletin, September/October 2010 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Basin evolution in the northern Altiplano, Bolivia

Figure 8. (A) Photograph of intraforma- A tional thrust faults and mesoscopic folds SW NE in exposures of the Aranjuez Formation near Zona Sur (south La Paz), indicating signifi cant postdepositional deformation. Vehicles (outlined) and road for scale. (B) Photograph and line drawing of growth strata developed in the Aranjuez Forma- tion (Ta) in angular unconformity above Ta the Devonian Sica Sica Formation (Ds) in Rio Khellkata river-cut exposures east of Rio La Paz (Fig. 2). Fanning bedding dips, abrupt lateral thickness variations, and Ta internal angular unconformities (arrows) attest to syndepositional deformation. Per- son (outlined) in foreground is ~100 m in front of cliff face.

from conglomerate clast imbrication (Fig. 4). B Paleocurrent indicators for the measured sec- tions of the Peñas Formation show a change in the average fl ow direction with stratigraphic level (Fig. 9). The paleocurrent indicators in the lower ~425 m of the Peñas Formation show a wide range of fl ow directions, with the average fl ow toward the southeast (Fig. 9A). The paleocurrent directions in the upper part of the Peñas section (~425–1100 m) indicate predominantly west-southwest–directed fl ow with a moderately divergent dispersal pattern (Fig. 9B). A general westward fl ow direction is also measured in outcrops of the Peñas Forma- tion at Cerro Santa Ana (Fig. 3B). The upsec- tion change in sediment dispersal is considered to be the result of initial fl ow potentially par- allel to an uplifted source region followed by transverse fl ow approximately perpendicular to the source region. This pattern is consistent SW NE with the facies interpretations suggesting that deposition in the medial to proximal zones Cliff top of a streamfl ow-dominated terminal fan pro- graded over more distal deposits of a distal fan to braided fl uvial system. In other words, the temporal shift in depositional systems may coincide with the upsection replacement of southeastward, potentially axial fl ow by west- directed fl ow in diverging distributary channels of the medial to proximal fan. For the Aranjuez Ta Formation, paleocurrent measurements from Angular this study and Gillis et al. (2006) indicate an unconformity average south-southwest-directed paleofl ow in with Paleozoic strata a variable to poorly defi ned divergent pattern (Fig. 9C). Although the paleocurrent patterns Ds are not defi nitive, the facies associations and average fl ow directions for the Aranjuez For- mation suggest deposition in an alluvial fan system derived from the east-northeast. Cliff base

Geological Society of America Bulletin, September/October 2010 1455 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Murray et al.

Compositional Methods granites in the Eastern Cordillera (Horton et al., footnote 1]) for quartz-feldspar-lithic fragments 2002). Quartzite clasts are commonly associ- (Q-F-L %), monocrystalline quartz-feldspar- To determine the provenance of the Peñas ated with the Ordovician Amutara Formation, total lithic fragments (Qm-F-Lt %), and meta- and Aranjuez formations and possible strati- and carbonate clasts are likely derived from morphic-volcanic-sedimentary lithic fragments graphic trends, conglomerate and sandstone either the Carboniferous Copa cabana For- (Lm-Lv-Ls %) are based on methods defi ned compositional data were collected from the four mation or the Upper Cretaceous El Molino by Ingersoll et al. (1984) and Dickinson (1985). measured stratigraphic sections and addi tional Formation (e.g., Sempere, 1995; Suárez and isolated outcrops. Conglomerate clast counts Diaz, 1996; McQuarrie and DeCelles, 2001; Peñas Formation were conducted on the outcrop in 28 localities Horton et al., 2002; Gillis et al., 2006). Several at an approximate stratigraphic spacing of 20– types of sandstone clasts may be derived from Description 60 m and consisted of lithologic identifi cation Silurian–Devonian strata: Silurian sandstones Clastic provenance of the Peñas Formation of an average of 100 clasts per count. Although are generally gray to green (e.g., Catavi Forma- was determined from 16 conglomerate clast many conglomerate clasts cannot be identifi ed tion); Devonian sandstones are typically reddish counts (Fig. 10) and 34 sandstone point counts at the formation level with confi dence, some purple (e.g., Vila Vila Formation) and tan (e.g., (Fig. 11) for the measured sections. Overall, clast types may be attributed to specifi c rock Sica Sica and Colpacucho formations) (Suárez the dominant conglomerate clasts are quartzite, units or specifi c intervals of the Paleozoic strati- and Diaz, 1996). reddish-purple sandstone, and tan sandstone, graphic section. In the Altiplano, granite and Compositional data for 39 fi ne- to coarse- with minor amounts of gray siltstone, gray gneiss clasts are typically derived from Precam- grained sandstones were determined for sam- sandstone, and green sandstone (Figs. 4A, 4B, brian basement that may have been exposed in ples collected from measured sections at an and 10A). Granite and gneiss clasts were not the Western Cordillera or from Permo-Triassic ~50 m inter val. For each sample, a standard observed, and volcanic clasts are extremely petrographic thin section was cut, etched with rare. Similar clast compositions were observed concentrated hydrofl uoric acid, and stained for in two conglomerate clast counts at Cerro Santa potassium feldspar. Four-hundred points per Ana (Figs. 3B and 10B). Individual clast counts A Lower Peñas Fm. thin section were analyzed using a modifi ed of conglomerates in the Peñas N section gener- Gazzi-Dickinson method of point counting ally display an upsection decrease in the relative

Vector Mean = 131° ± 102° to determine composition (e.g., Gazzi, 1966; abundance of quartzite clasts, accompanied by Vector Magnitude = 22% Dickinson, 1970; Ingersoll et al., 1984; Dick- an increase in reddish-purple and tan sandstone Maximum Percentage = 14% inson, 1985). Framework grains (>0.0625 mm) clasts (Fig. 4A). n = 296 (22 stations) were classifi ed using the petrographic count- Compositionally, sandstones of the Peñas For- ing parame ters listed in Table 4. In cases of mation are sublitharenites and quartz arenites , dia genetic alteration, such as partial alteration refl ecting their high percentage of monocrystal- of feldspar to clay minerals, framework grains line quartz, limited lithic fragments, and lack of B Upper Peñas Fm. were counted as the original grain type. Lithic feldspar (e.g., Pettijohn et al., 1987). The com- fragments mainly consist of metamorphic positions do not vary signifi cantly upsection or Vector Mean = 256° ± 61° grains, argillaceous grains, and polycrystalline between stratigraphic sections, with average Vector Magnitude = 69% Maximum Percentage = 17% quartz grains. Another common framework Q-F-L % = 91-1-8 and average Qm-F-Lt % = n = 322 (29 stations) grain is monocrystalline quartz with rounded 88-1-11 (Fig. 11A and Table DR2 [see foot- syntaxial overgrowths. Matrix, authigenic note 1]). No samples display pressure solution cement , and pore spaces were also included in textures common in highly compacted sand- the point-counting parameters. Recalculated stone, and the interstitial space is predominantly C Aranjuez Fm. detrital modal percentages (Table DR2 [see occupied with infi ltrated mud or epimatrix (e.g.,

TABLE 4. PARAMETERS FOR SANDSTONE POINT COUNTS Vector Mean = 200° ± 97° Vector Magnitide = 20% lobmyS seirogetacniarG sretemarapdetaluclaceR Maximum Percentage = 20% Qmm Monocrystalline quartz Q-F-L: n = 168 (10 stations) Qmo Monocrystalline quartz with rounded syntaxial overgrowths Q = Qmm + Qmo + Qp + Qpt ztrauqenillatsyrcyloPpQ kF+pF=F cirbafetinotcethtiwztrauqenillatsyrcyloPtpQ mL+sL+vL=L Fp Plagioclase feldspar Fk Potassium feldspar Qm-F-Lt: stnemgarfcihtillassybapyh-cinaclovcirtiVvvL omQ+mmQ=mQ stnemgarfcihtillassybapyh-cinaclovcitisleFfvL kF+pF=F Figure 9. Paleocurrent data for measured stnemgarfcihtillassybapyh-cinaclovcitilorciMmvL tpQ+pQ+mL+sL+vL=tL sections of the Peñas and Aranjuez forma- Lvl Lathwork volcanic-hypabyssal lithic fragments Lmv Metavolcanic lithic fragments Lm-Lv-Ls: tions displaying vector mean and standard stnemgarfcihtilacimenillatsyrcyloPmmL amL+tmL+mmL+vmL=mL deviation, vector magnitude, maximum Lmt Quartz-feldspar-mica tectonite lithic fragments, including phyllite Lv = Lvv + Lvf + Lvm + Lvl percentage, and number of measurements Lma Quartz-feldspar-mica aggregate lithic fragments Ls = Lss + Lsc Lss Argillaceous-shale lithic fragments for each rose diagram. (A) Lower Peñas For- Lsc Extra-basinal carbonate lithic fragments mation (0 to ~425 m). (B) Upper Peñas M Monocrystalline mica Formation (~425–1100 m). (C) Combined D Dense minerals (pyroxenes, amphiboles, etc.) Misc. Miscellaneous and unidentifi ed grains paleocurrent data from this study and Gillis M/C Matrix (<0.0625 mm) and/or cement et al. (2006) for the Aranjuez Formation. Pore Pore space

1456 Geological Society of America Bulletin, September/October 2010 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Basin evolution in the northern Altiplano, Bolivia

A Peñas (N and S sections) B Cerro Santa Ana tectonite fabrics, and quartz-feldspar-mica tec- tonites (Fig. 11) is consistent with derivation from quartz-cemented sandstone, quartzite, and phyllite found in Paleozoic strata of the Eastern Cordillera. Additionally, polycrystalline quartz and monocrystalline quartz with undulatory extinction observed in these samples suggests that these grains were likely derived from a low- grade metamorphic source (Basu, 1985). The trace amounts of felsitic volcanic lithic grains n = 1568 n = 123 could be the erosional products of either Ceno- zoic volcanic rocks related to arc magmatism or Jurassic–Cretaceous hypabyssal intrusions and Quartzite Gray sandstone Green sandstone Reddish-purple sandstone fl ows exposed locally in the Eastern Cor dillera Gray siltstone Tan sandstone (e.g., Kennan et al., 1995; SERGEOMIN, 1997; DeCelles and Horton, 2003; Jiménez and Figure 10. Conglomerate clast compositional data for the Peñas López-Velásquez, 2008). Formation. (A) Total compositional data for measured sections of Conglomerate clast compositions (Fig. 10) the Peñas Formation. See Figure 3A for locations and Figure 4 for are consistent with sandstone compositions, individual clast counts. (B) Total compositional data for outcrops of and suggest derivation from Paleozoic rocks the Peñas Formation at Cerro Santa Ana. See Figure 3B for loca- of the Eastern Cordillera. The dominant clast tions and individual clast counts. lithology observed in the Peñas Formation is white to dark-gray quartzite, which is attrib- uted to the Ordovician Amutara Formation. Dickinson, 1970). Most monocrystalline quartz scales (Ingersoll et al., 1993). The recycled Reddish-purple and tan sandstone clasts were grains display straight extinction, but some provenance of these sandstones is verifi ed by likely derived from the Devonian Vila Vila and display undulatory and wavy extinction, with the dominance of metamorphic and sedimentary Sica Sica formations, respectively. Limited a few grains exhibiting strain lamellae. Some lithic fragments, and limited degree of volcanic clasts of gray siltstone, green sandstone, and monocrystalline quartz grains exhibit rounded detritus (e.g., Ingersoll, 1990). The presence of gray sandstone may have been derived from syntaxial overgrowths, suggesting of monocrystalline quartz with rounded syntaxial the Silurian Uncia and Catavi formations. The quartz-cemented sandstone. Polycrystalline overgrowths, polycrystalline quartz grains with lack of granite and gneiss clasts indicates that quartz with tectonite fabric is also present in the samples. Feldspar is rare in the sandstone Peñas S samples; where present, potassium feldspar Peñas N is more abundant than plagioclase. The domi- Qm Qm nant lithic fragments consist of metamorphic 0–299 m 0–299 m quartz-feldspar-mica tectonites and aggregates 300–599 m 300–599 m and sedimentary argillaceous-shale fragments, 600+ m 600+ m with trace felsitic volcanic lithic fragments (Fig. 11B). In both measured sections, the proportion Lm Lm of sedimentary lithic fragments increases upsec- tion (Fig. 11B).

Interpretation Based on conglomerate clast compositions A A (Fig. 10), sandstone petrofacies (Fig. 11), and west-directed paleofl ow direction (Fig. 9), the Peñas Formation was likely derived from recy- F Lt F Lt cling of quartz-rich Paleozoic metasedimentary and sedimentary rocks of the Eastern Cordi- llera. The average sandstone Qm-F-Lt % plots B B in the “recycled-orogenic” tectonic setting of Dickinson (1985); however, the sedimentary Lv Ls Lv Ls facies associations indicate a fi rst- to second- order alluvial sampling, which is not ideal to Figure 11. (A) Monocrystalline quartz-feldspar-total lithic fragments (Qm-F-Lt) and use with the third-order tectonic settings of (B) metamorphic-volcanic-sedimentary lithic fragments (Lm-Lv-Ls) ternary diagrams for the Dickinson Qm-F-Lt ternary diagram (e.g., 34 sandstone samples of the Peñas Formation collected from the Peñas N (left) and Peñas S Ingersoll, 1990; Ingersoll et al., 1993). On the (right) measured sections (Fig. 4). Note the upsection trends (arrows) toward greater sedi- other hand, “recycled-orogenic” settings tend to mentary lithic fragments. Recalculated modal percentages for samples are shown in Table produce uniform compositions at all sampling DR2 [see footnote 1].

Geological Society of America Bulletin, September/October 2010 1457 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Murray et al. sediments were not derived from Precambrian Aranjuez Pedregal basement in the Western Cordillera. Addition- A B ally, the lack of granite clasts implies that the abundant Permo-Triassic granites exposed today in the Eastern Cordillera were not ex- humed during deposition of the Peñas For- mation. The lack of carbonate clasts suggests that the Copacabana and El Molino formations were not exposed over signifi cant areas of the source region. The proportion of lithic fragment types found within the Peñas Formation sandstones changes upsection from predominantly metamorphic n = 476 n = 484 to roughly equal metamorphic and sedimen- tary fragments, suggesting that the source-rock Quartzite Purple siltstone composition evolved through time from primar- Green sandstone Tan sandstone Green siltstone Tan siltstone ily metasedimentary to mixed sedimentary and Gray sandstone Green carbonate metasedimentary (Fig. 11B). The compositional Gray siltstone Andesite/dacite trend in conglomerate clast types is consistent Reddish-purple sandstone with the sandstone petrographic data, showing Figure 12. Conglomerate clast compositional data for the Aranjuez an upsection reduction in Ordovician quartzite Formation. See Figure 3C for locations and Figure 4 for individual clast and increase in younger Devonian sandstone counts. (A) Total compositional data for the Aranjuez measured sec- (Figs. 4A and 4B). tion. (B) Total compositional data for the Pedregal measured section. Aranjuez Formation

Description wavy extinction. Polycrystalline quartz and Conglomerate clast compositions suggest Ten conglomerate clast counts and fi ve sand- monocrystalline quartz with rounded syntaxial erosion of Silurian to Devonian strata, with stone point counts provide provenance informa- overgrowths are less common framework con- limited Ordovician and Carboniferous prov- tion for the Aranjuez Formation. Reddish-purple stituents. The sandstone samples contain limited enance (Fig. 12). Green to gray sandstone and sandstone, green sandstone, and tan sandstone feldspar, with potassium feldspar more common gray to green siltstone clasts were likely eroded are the dominant clast types, with gray sand- than plagioclase. Lithic fragments are mostly from Silurian Catavi and Uncia formations, stone, gray siltstone, and tan siltstone repre- metamorphic quartz-feldspar-mica tectonites respectively. The eroded Devonian strata are senting minor components (Fig. 12). Clasts of and aggregates, with lesser amounts of sedi- represented by clasts of reddish-purple sand- quartzite and green siltstone with trace amounts mentary argillaceous-shale and felsitic volcanic stone (Vila Vila Formation), tan sandstone of green carbonate and andesite-dacite are only fragments (Fig. 13B). No clear stratigraphic (Sica Sica Formation), and tan siltstone (Belén observed in the Aranjuez section. Gneiss and trends in sandstone composition are recognized Formation). Quartzite clasts in the Aranjuez granite clasts are absent from these conglomer- for the Aranjuez Formation. section were likely derived from the Ordovi- ates. Individual clast counts show a poorly de- cian Amutara Formation, with green carbonate fi ned compositional trend of decreasing quartzite Interpretation clasts possibly representing minor input from clasts upsection in the Aranjuez measured sec- Similar to the Peñas Formation, the com- the Carboniferous Copacabana Formation or tion (Fig. 4C), and an upsection decrease in bination of conglomerate clast compositions Upper Cretaceous El Molino Formation. Gran- green sandstone accompanied by increased tan (Fig. 12), sandstone petrofacies (Fig. 13), and ite clasts are absent, indicating that the Permo- and reddish-purple sandstone in the lower 120 m westward-directed paleofl ow directions indicate Triassic granites in the Eastern Cordillera had of the Pedregal measured section (Fig. 4D). that the Aranjuez Formation was likely derived not yet reached the surface during deposition Sandstone facies are relatively uncommon from erosion of Paleozoic strata in the Eastern of the Aranjuez Formation, and that sediments in the conglomerate- and mudstone-dominated Cordillera. The predominantly metamorphic were not derived from Precambrian basement Aranjuez Formation. The sandstones are typi- to sedimentary lithic-rich compositions of exposed in the Western Cordillera. cally very fi ne to fi ne grained and are classi- these fi rst-order sandstone samples (Fig. 13) Similar to the Peñas Formation, upsection fi ed as lithic wackes based on their quartz- and suggest provenance from a recycled mixed trends of decreasing quartzite clasts in the Aran- lithic-rich, feldspar-poor composition and high metasedimentary-sedimentary source (e.g., Inger- juez measured section and decreasing green percentage of silt (<0.0625 mm) matrix of in- soll, 1990; Ingersoll et al., 1993). The relatively sandstone clasts accompanied by increasing tan fi ltrated mud and calcite cement (e.g., Petti- small amount of polycrystalline quartz indicates and reddish-purple sandstone clasts in the lower john et al., 1987). The modal composition of that the majority of quartz grains were subjected Pedregal measured section suggest relatively in- the Aranjuez sandstone samples is variable, to high strain prior to erosion (e.g., Basu, 1985). creased detrital input from younger stratigraphic with average Q-F-L % = 64-4-32 and average Monocrystalline quartz grains with rounded units with time. Although the Peñas and Aran- Qm-F-Lt % = 61-4-35 (Fig. 13A and Table syntaxial overgrowths derived from eroded juez formations are age-equivalent (Table 3), DR2 [see footnote 1]). Most monocrystalline quartz-cemented sandstone also supports the there is a difference in clast compositions ex- quartz grains display straight extinction, with interpretation of primarily sedimentary rocks in pressed as a greater proportion of quartzite a small percentage showing undulatory and the source region. clasts in the Peñas Formation (Figs. 10 and 12).

1458 Geological Society of America Bulletin, September/October 2010 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Basin evolution in the northern Altiplano, Bolivia

Qm (Fig. 4), with conglomerates absent in the with sedimentation (Fig. 14). In this scenario, lower portion but dominant near the top. This prior to and during earliest basin development, pattern is interpreted as an increase in the prox- thrust-induced denudation in the Eastern Cor- imity of the sediment source to the basin with dillera resulted in unroofi ng of Ordovician time. The observed facies associations refl ect strata (Fig. 14A). Forward propagation of west- Lm this temporal change in source-area proximity, southwest –directed thrust faults during Peñas and with an upsection facies transition from distal Aranjuez deposition increased the proximity of fan or braided fl uvial deposition to proximal, the sediment source to the basin (Fig. 14B), uplift- streamfl ow-dominated fan deposition. ing younger (Devonian) strata that resulted in an Provenance data from conglomerate clast increase in average grain size and erosion of these A counts (Figs. 4, 10, and 12) reveal an increase in younger rocks. Eventually, continued advance of the amount of sediments derived from younger the central Andean backthrust belt resulted in Paleozoic strata with time. For example, the fault propagation to the west-southwest of the F Lt lower Peñas Formation was primarily derived basin (Fig. 14C), initiating deposition of younger from Ordovician rocks, but detrital compositions synorogenic sedimentary strata (e.g., Coniri For- B upsection represent roughly equal contributions mation; Sempere et al., 1990; Horton et al., 2002) from Ordovician and Devonian rocks (Figs. 4A and hanging-wall uplift and deformation of the and 4B). The Aranjuez Formation also refl ects Peñas and Aranjuez deposits. This interpretation Lv Ls this trend, with an upsection reduction in Ordo- (Fig. 14) accounts for both the sedimentological vician and Silurian detritus relative to Devonian trend toward more proximal facies and the prov- Figure 13. (A) Monocrystalline quartz-feld- detritus (Figs. 4C and 4D). A similar trend is enance trend in which younger Paleo zoic rocks spar-total lithic fragments (Qm-F-Lt) and expressed in the sandstone petrographic data by were introduced into the sediment source area. (B) metamorphic-volcanic-sedimentary lithic an upsection shift in the ratio of sedimentary to Additionally, this interpretation is supported by fragments (Lm-Lv-Ls) ternary diagrams metamorphic lithic fragments (Fig. 11B). These the west-southwestward migration of the locus of for fi ve sandstone samples of the Aranjuez upsection provenance shifts seem counter- basin subsidence inferred from our new 40Ar/39Ar Formation collected from the Aranjuez and intuitive because simple erosional unroofi ng of a ages for the basin fi ll (Table 3). Pedregal measured sections (Fig. 4). Recal- single fault block would be expected to produce Previous studies have attributed growth of culated modal percentages for samples are an upsection trend in which detritus is systemati- the central Andean backthrust belt to emplace- shown in Table DR2 [see footnote 1]. cally older and from lower structural levels. ment of an eastward-propagating basement Several possibilities may explain this prov- megathrust over a 12 km ramp located near the enance trend, including shifts in drainage net- eastern Altiplano–Eastern Cordillera boundary This compositional difference potentially rep- works within an evolving sediment source area, (McQuarrie and DeCelles, 2001; McQuarrie, resents along-strike variations in source rock erosion through a single thrust sheet, or erosion 2002; McQuarrie et al., 2008). Continued tectonic exposure within the Eastern Cordillera. At face of newly uplifted exposures of younger Paleo- wedging on this basement megathrust allowed value, the clast compositions of the Peñas and zoic strata. In the fi rst case, changes within the slip to be transferred to the overlying Paleozoic Aranjuez formations suggest that late Oligo- source drainage network may have focused strata along décollement horizons located in cene rock uplift in the Eastern Cordillera was erosion in areas with Devonian strata, result- the Ordovician and Silurian formations, result- not regionally uniform, with deeper (Ordovi- ing in an increased proportion of younger sedi- ing in growth of the west-southwest–directed cian) stratigraphic levels exposed in the north mentary fragments. In the second case, erosion backthrust belt (McQuarrie, 2002; McQuarrie adjacent to the Peñas Formation, but principally through a single thrust sheet could produce an et al., 2008). However, the precise sequence of younger (Silurian) levels exposed farther south. initial infl ux of Ordovician quartzite derived deformation within that backthrust belt has been Despite these compositional variations, both the from the hanging wall followed by an increased diffi cult to unravel (e.g., Gillis et al., 2006). The Peñas and Aranjuez formations show consistent amount of footwall-derived Devonian detritus. stratigraphic and provenance relationships ob- upsection trends toward increased input from Although plausible, neither of these two options served within the Peñas and Aranjuez forma- relatively younger Paleozoic strata, suggestive can explain the temporal increase in grain size tions are consistent with a model of systematic of broadly similar unroofi ng patterns in com- represented by the upward coarsening trend in forward propagation of sequential thrust faults to parable structural systems. the basin. Therefore, we favor a third explana- the west-southwest (Fig. 14). This “in sequence” tion in which erosion of progressively younger model of thrusting and advance of the central DISCUSSION Paleozoic strata is the product of a systematic Andean backthrust belt toward the Altiplano may advance of fold-thrust deformation toward the help explain the partial discrepancy between pre- Basin Provenance basin, with the activation of new thrust faults in viously reported evidence for rapid exhumation greater proximity to the basin. that is derived from thermochronology compared The sedimentology, provenance, and 40Ar/39Ar to that derived from basin analysis. ages for the Peñas and Aranjuez formations Structural Evolution suggest late Oligocene–early Miocene deposi- Exhumational Record tion in a west-southwestward prograding sys- We propose that the mechanisms responsible tem of streamfl ow- and mass-fl ow–dominated for the observed sedimentological and prov- Thermochronological studies in the Eastern alluvial fans composed of sediments derived enance trends in the Peñas and Aranjuez forma- Cordillera suggest that a period of rapid cool- from the Eastern Cordillera. The Peñas For- tions are related to west-southwestward growth ing associated with rock uplift initiated during mation displays an upward coarsening trend of the central Andean backthrust belt concurrent the middle Eocene at ~40 Ma and continued

Geological Society of America Bulletin, September/October 2010 1459 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Murray et al.

Figure 14. Schematic cross sec- WSW Peñas and Aranjuez ENE tions depicting a simplifi ed formations sequence of fold-thrust defor- CENTRAL ANDEAN BACKTHRUST BELT mation in the central Andean A ALTIPLANO backthrust belt during late D Oligo cene to early Miocene basin DEVONIAN S development (Peñas and Aran- SILURIAN O juez formations). Vertical arrow ORDOVICIAN indicates location of measured sections with time. Approximate décollement horizons are based on balanced cross sections by B D McQuarrie (2002), McQuarrie D D S et al. (2008), and fi eld observa- S S O tions. (A) Early deformation in O the backthrust belt resulting in uplift and erosion of Silurian– Devonian strata, followed by exposure of Ordovician rocks, which dominated initial prov- C D enance in the basin. (B) West- D D D S southwestward propagation of S S S O thrust front into the Altiplano O causing uplift of previously fl at- lying Paleozoic strata and increasing the proximity of the source area, resulting in increases in average grain size and amount of Devonian- derived detritus. (C) Continued advance of thrust front inducing postdepositional deformation of the Peñas and Aranjuez formations and initiation of younger synorogenic sedimentation farther to the west-southwest. until ~25 Ma (Benjamin et al., 1987; McBride temporal delay would be observed between advance of deformation in the central Andean et al., 1987; Farrar et al., 1988; Masek et al., the onset of rapid exhumation in the core of the backthrust belt, away from the core of the East- 1994; Barnes et al., 2006; Gillis et al., 2006; backthrust belt, notably the Cordillera Real, ern Cordillera, to be the product of substantial Ege et al., 2007; Barnes et al., 2008; McQuarrie and the frontal zones of the backthrust belt crustal thickening within the Eastern Cor di- et al., 2008). Provenance data from this study along the margins of the Altiplano. Therefore, llera. Structural evidence also suggests that document the presence of Ordovician quartzite thermochronological evidence supporting the much of the upper-crustal shortening in the re- clasts in the lower Peñas Formation, which in- proposed middle Eocene onset of rapid cool- gion had terminated in the early Miocene time dicates that lower Paleozoic metasedimentary ing by uplift in the Cordillera Real (Gillis et al., (McQuarrie and DeCelles, 2001; Horton, 2005; rocks in the Eastern Cordillera were subjected 2006) should not be considered to be in confl ict Gillis et al., 2006). Therefore, we propose that to erosion during the late Oligocene (Figs. 10, with the late Oligocene infl ux of locally derived concentrated Oligocene–early Miocene short- 12, and 14A). This implies that considerable ex- sediment in intermontane or hinterland basins ening within the Eastern Cordillera and eastern humation and erosion of portions of the Eastern (MacFadden et al., 1985; Sempere et al., 1990; Altiplano resulted in suffi cient crustal thicken- Cordillera occurred prior to the deposition of Horton et al., 2002; this study) at the leading ing to generate large-scale isostatic uplift and the east-derived deposits in the Altiplano, pos- (west-southwestern) edge of the central Andean initial expression of the eastern margin of the sibly during an initial, middle Eocene to early backthrust belt. central Andean plateau. Oligocene period of uplift-related rapid cooling. However, exhumation was likely insuffi cient to Regional Implications CONCLUSIONS expose the Eastern Cordillera’s Permo-Triassic granites by the late Oligocene, because no gran- Growth structures, stratigraphic trends, and (1) Sedimentary lithofacies and facies asso- ite clasts are observed in the Peñas and Aran- 40Ar/39Ar ages for the Peñas and Aranjuez for- ciations identifi ed in the upper Oligocene to juez formations. Although there are no clear mations indicate syndepositional shortening lower Miocene Peñas and Aranjuez formations erosional products from the early phase of ex- during late Oligocene–early Miocene evo- of the northern Altiplano of Bolivia indicate humation affecting the Paleozoic section, one lution of a hinterland basin in the northern nonmarine deposition of coarse clastic sediment possibility is that the initial erosional products Altiplano. During this period, much of the within the central Andean backthrust belt. Ac- may have been deposited in wedge-top basins eastern Altiplano received an infl ux of clastic cumulation took place in the proximal, medial, closer to the core of the central Andean back- sediments derived from the Eastern Cordillera and distal zones of a streamfl ow-dominated thrust belt and were subsequently uplifted and (MacFadden et al., 1985; Sempere et al., 1990; terminal fan system, in a mass-fl ow–dominated eroded during migration of the deformation Horton et al., 2002; Gillis et al., 2004, 2006). alluvial fan, and potentially in braided fl uvial front toward the Altiplano. Given the sequen- Although several possible features may control environments at the downslope margins of the tial west-southwestward advance of the central the kinematic history of fold-thrust deforma- fans. Upsection transitions to more proximal Andean backthrust belt, it seems likely that a tion, we consider the west-southwestward facies are consistent with west-southwestward

1460 Geological Society of America Bulletin, September/October 2010 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Basin evolution in the northern Altiplano, Bolivia progradation of depositional systems and in- from thermochronological studies and the late Dickinson, W.R., 1970, Interpreting detrital modes of gray- creased proximity to the sediment source areas Oligocene infl ux of sediment in the Altiplano. wacke and arkose: Journal of Sedimentary Petrology, v. 40, p. 695–707. in the Eastern Cordillera. (5) The Peñas and Aranjuez formations in the Dickinson, W.R., 1985, Interpreting provenance relations (2) 40Ar/39Ar ages for igneous units within northern Altiplano are among the best exposed from detrital modes of sandstones, in Zuffa, G.G., ed., Provenance of Arenites: Dordrecht, D. Reidel, the lowermost exposed Aranjeuz Formation records of the middle Cenozoic infl ux of east- p. 333–361. and interbedded ash-fall tuffs within the Aran- derived sediments within a regional hinterland Ege, H., Sobel, E.R., Scheuber, E., and Jacobshagen, V., juez, Peñas, and overlying La Paz formations basin. This period of basin accumulation was 2007, Exhumation history of the southern Altiplano plateau (southern Bolivia) constrained by apatite provide key temporal constraints on the onset coeval with a signifi cant period of Oligocene– fission track thermochronology: , v. 26, of basin evolution, rate of basin advance, and early Miocene shortening in the central Andean p. TC1004, doi: 10.1029/2005TC001869. timing of upper-crustal shortening in the cen- backthrust belt. Advance of the deformation Farrar, E., Clark, A.H., Kontak, D.J., and Archibald, D.A., 1988, Zongo-San Gabon zone: Eocene foreland bound- 40 39 tral Andean backthrust belt. Ar/ Ar ages in front along the Altiplano–Eastern Cordillera ary of central Andean orogen, northwest Bolivia and the Aranjuez and Peñas formations range from boundary further suggests large-scale crustal southeast : Geology, v. 16, p. 55–58, doi: 10.1130/ 0091-7613(1988)016<0055:ZSGNZE>2.3.CO;2. ~27.5 Ma to 25 Ma, indicating consistent late thickening, and by inference, isostatic uplift and Friend, P.F., 1978, Distinctive features of some ancient river Oligocene accumulation. Deposition of the over- initial topographic growth of the eastern margin systems, in Miall, A.D., ed., Fluvial Sedimentology: lying La Paz Formation commenced in middle of the central Andean plateau. Canadian Society of Petroleum Geologists Memoir no. 5, p. 531–542. Miocene time. The 40Ar/39Ar ages from the ACKNOWLEDGMENTS Garzione, C.N., Molnar, P., Libarkin, J.C., and MacFadden, lowermost exposed strata in different localities B.J., 2006, Rapid late Miocene rise of the Bolivian show that initial Aranjuez deposition in proxi- Alti plano: Evidence for removal of mantle lithosphere: This research was partially supported by National Earth and Planetary Science Letters, v. 241, p. 543– mal settings predated initial Peñas deposition in Science Foundation grants EAR-0510441 and EAR- 556, doi: 10.1016/j.epsl.2005.11.026. distal settings by ~2 Myr, indicative of a west- 0908518 and a graduate student research grant from Garzione, C.N., Hoke, G.D., Libarkin, J.C., Withers, S., southwestward migration of the locus of basin the Geological Society of America. We thank Néstor MacFadden, B., Eiler, J., Ghosh, P., and Mulch, A., Jiménez, Sohrab Tawackoli, Robert Gillis, Raymond 2008, Rise of the Andes: Science, v. 320, p. 1304– subsidence. Growth strata and stratal overlap re- Ingersoll, Andrew Leier, Jesse Mosolf, and Nadine 1307, doi: 10.1126/science.1148615. lationships demonstrate that motion along fold- McQuarrie for benefi cial discussions. Dana Roeber Gazzi, P., 1966, Le arenarie del fl ysch sopracretaceo Murray and Nelsón Ayala provided assistance in the dell’Appennino modenese: Correlazioni con il fl ysch thrust structures of the central Andean backthrust di Monghidoro: Mineralogica et Petrographica Acta, belt was contemporaneous with late Oligocene fi eld. Constructive reviews by Jason Covault, Andrew v. 12, p. 69–97. Hanson, Abhijit Basu, and Charles Mitchell helped to early Miocene deposition of the Aranjuez and GEOBOL (Servicio Geológico de Bolivia), 1995a, Carta improve the manuscript. Geologica de Bolivia, Hoja : Servicio Geo- Peñas formations, but had ceased prior to middle logico de Bolivia, scale 1:100,000. Miocene deposition of the La Paz Formation. REFERENCES CITED GEOBOL (Servicio Geológico de Bolivia), 1995b, Carta (3) Sediment provenance information derived Geologica de Bolivia, Hoja de La Paz: Servicio Geo- Allmendinger, R.W., Jordan, T.E., Kay, S.M., and Isacks, logico de Bolivia, scale 1:100,000. from conglomerate clast compositions, sand- B.L., 1997, The evolution of the Altiplano-Puna pla- GEOBOL (Servicio Geológico de Bolivia), 1995c, Carta stone petrographic data, and paleocurrent ori- teau of the central Andes: Annual Review of Earth and geologica de Bolivia, Hoja Milluni: Servicio Geo- entations show that Paleozoic sedimentary and Planetary Sciences, v. 25, p. 139–174, doi: 10.1146/ logico de Bolivia, scale 1:100,000. annurev.earth.25.1.139. Ghosh, P., Garzione, C.N., and Eiler, J.M., 2006, Rapid up- metasedimentary rocks of the Eastern Cordi- Barnes, J.B., Ehlers, T.A., McQuarrie, N., and Pelletier, J.D., lift of the Altiplano revealed through 13C-18O bonds in llera were the principal sediment sources for the 2006, Variations in Eocene to recent erosion across the paleosol carbonates: Science, v. 311, p. 511–515, doi: central Andean fold-thrust belt, northern Bolivia: Im- 10.1126/science.1119365. Peñas and Aranjuez formations of the northern plications for plateau evolution: Earth and Planetary Gillis, R.J., Horton, B.K., and Grove, M., 2004, Exhuma- Altiplano. The compositional provenance data Science Letters, v. 248, p. 118–133, doi: 10.1016/ tion history and basin development along the eastern reveal a complex unroofi ng pattern in which j.epsl.2006.05.018. margin of the central Andean plateau, Bolivia: Geo- Barnes, J.B., Ehlers, T.A., McQuarrie, N., O’Sullivan, P.B., logical Society of America Abstracts with Programs, an upsection increase in younger detritus is ex- and Tawackoli, S., 2008, Thermochronometer record v. 36, p. 433. pressed by an increased proportion of Devonian of central Andean Plateau growth, Bolivia (19.5°S): Gillis, R.J., Horton, B.K., and Grove, M., 2006, Thermo- clasts relative to Ordovician clasts and by an in- Tectonics, v. 27, doi: 10.1029/2007TC002174. chronology, geochronology, and upper crustal structure Basu, A., 1985, Reading provenance from detrital quartz, in of the Cordillera Real: Implications for Cenozoic exhu- crease in the ratio of sedimentary to metasedi- Zuffa, G.G., ed., Provenance of Arenites: Dordrecht, mation of the central Andean plateau: Tectonics, v. 25, mentary lithic fragments. These provenance D. Reidel, p. 231–247. doi: 10.1029/2005TC001887. Beck, S.L., Zandt, G., Myers, S.C., Wallace, T.C., Silver, P.G., González, M., Diaz-Martinez, E., and Ticlla, L., 1996, trends require the introduction of younger rocks and Drake, L., 1996, Crustal thickness variations in the Comen tarios sobre la estratigrafi a del Silurico y Devo- in the source area and, therefore, are not compat- Central Andes: Geology, v. 24, p. 407–410, doi: 10.1130/ nico del norte y centro de la Cordillera Oriental y Alti- ible with simple unroofi ng of a single fault block. 0091-7613(1996)024<0407:CTVITC>2.3.CO;2. plano de Bolivia: Ponta Grossa, Brazil, Simposio Sul Benjamin, M.T., Johnson, N.M., and Naeser, C.W., 1987, Re- Americano do Siluro-Devoniano, Anais, p. 117–130. (4) The complex upsection shift in prove- cent rapid uplift in the Bolivian Andes: Evidence from Gubbels, T.L., Isacks, B.L., and Farrar, E., 1993, High-level nance and general west-southwestward progra- fi ssion-track dating: Geology, v. 15, p. 680–683, doi: surfaces, plateau uplift, and foreland development, dation of depositional systems can be attributed 10.1130/0091-7613(1987)15<680:RRUITB>2.0.CO;2. Bolivian central Andes: Geology, v. 21, p. 695–698, Blair, T.C., and McPherson, J.G., 1994, Alluvial fans and doi: 10.1130/0091-7613(1993)021<0695:HLSPUA> to sequential activation of fold-thrust structures their natural distinction from rivers based on morphol- 2.3.CO;2. within the central Andean backthrust belt. Ad- ogy, hydraulic processes, sedimentary processes, and Hampton, B.A., and Horton, B.K., 2007, Sheetfl ow fl uvial facies assemblages: Journal of Sedimentary Research, processes in a rapidly subsiding basin, Altiplano pla- vance of upper-crustal shortening toward the v. A64, p. 450–489. teau, Bolivia: Sedimentology, v. 54, p. 1121–1147, doi: leading (west-southwestern) edge of the back- DeCelles, P.G., and Giles, K.N., 1996, Foreland basin sys- 10.1111/j.1365-3091.2007.00875.x. thrust belt provides a mechanism for both the in- tems: Basin Research, v. 8, p. 105–123, doi: 10.1046/ Hoke, G.D., and Garzione, C.N., 2008, Paleosurfaces, paleo- j.1365-2117.1996.01491.x. elevation, and the mechanisms for the late Miocene creased proximity of sediment source areas and DeCelles, P.G., and Horton, B.K., 2003, Early to middle topographic development of the Altiplano plateau: the introduction of younger Paleozoic detritus Tertiary foreland basin development and the history of Earth and Planetary Science Letters, v. 271, p. 192– from more-frontal thrust sheets. At a regional Andean crustal shortening in Bolivia: Geological Soci- 201, doi: 10.1016/j.epsl.2008.04.008. ety of America Bulletin, v. 115, p. 58–77, doi: 10.1130/ Horton, B.K., 1998, Sediment accumulation on top of the scale, a long-term west-southwestward shift in 0016-7606(2003)115<0058:ETMTFB>2.0.CO;2. Andean orogenic wedge: Oligocene to late Miocene deformation could explain the ~10–15 Myr dis- DeCelles, P.G., Langford, R.P., and Schwartz, R.K., 1983, basins of the Eastern Cordillera, southern Bolivia: Two new methods of paleocurrent determination from Geological Society of America Bulletin, v. 110, crepancy between the estimated middle Eocene trough cross-stratification: Journal of Sedimentary p. 1174–1192, doi: 10.1130/0016-7606(1998)110 (~40 Ma) onset of rapid exhumation inferred Petrol ogy, v. 53, p. 629–642. <1174:SAOTOT>2.3.CO;2.

Geological Society of America Bulletin, September/October 2010 1461 Downloaded from gsabulletin.gsapubs.org on July 6, 2010 Murray et al.

Horton, B.K., 2005, Revised deformation history of the Leier, A.L., DeCelles, P.G., and Pelletier, J.D., 2005, Pettijohn, F.J., Potter, P.E., and Siever, R., 1987, Sand and central Andes: Inferences from Cenozoic foredeep Mountains, monsoons, and megafans: Geology, v. 33, Sandstone (Second Edition): New York, Springer- and intermontane basins of the Eastern Cordillera, p. 289–292, doi: 10.1130/G21228.1. Verlag, 553 p. Bolivia: Tectonics, v. 24, p. TC3011, doi: 10.1029/ Lohmann, H.H., 1970, Outline of tectonic history of Bolivian Reutter, K.J., Döbel, R., Bogdanic, T., and Kley, J., 1994, 2003TC001619. Andes: American Association of Petroleum Geologists Geological map of the Central Andes between 20° and Horton, B.K., and DeCelles, P.G., 1997, The modern fore- Bulletin, v. 54, p. 735–757. 26° S 1:1,000,000, in Reutter, K.J., Scheuber, E., land basin system adjacent to the Central Andes: MacFadden, B.J., Campbell, K.E., Ciffelli, R.L., Siles, and Wigger, P.J., eds., Tectonics of the Southern Central Geology, v. 25, p. 895–898, doi: 10.1130/0091-7613 O., Johnson, N.M., Naeser, C.W., and Zeitler, P.K., Andes: New York, Springer-Verlag, Enclosure 1. (1997)025<0895:TMFBSA>2.3.CO;2. 1985, Magnetic polarity stratigraphy and mamma- Ridgway, K.D., and DeCelles, P.G., 1993, Stream-dominated Horton, B.K., and DeCelles, P.G., 2001, Modern and ancient lian fauna of the Deseadan (late Oligocene–early alluvial fan and lacustrine depositional systems in fl uvial megafans in the foreland basin system of the Cen- Miocene) Salla beds of northern Bolivia: The Jour- Cenozoic strike-slip basins, Denali fault system, Yukon tral Andes, southern Bolivia: Implications for drainage nal of Geology, v. 93, p. 223–250, doi: 10.1086/ Territory, Canada: Sedimentology, v. 40, p. 645–666, network evolution in fold-thrust belts: Basin Research, 628950. doi: 10.1111/j.1365-3091.1993.tb01354.x. v. 13, p. 43–63, doi: 10.1046/j.1365-2117.2001.00137.x. Masek, J.G., Isacks, B.L., and Fielding, E.L., 1994, Erosion Roeder, D., 1988, Andean-age structure of Eastern Cor di- Horton, B.K., Hampton, B.A., and Waanders, G.L., 2001, and tectonics at the margins of continental plateaus: llera (province of La Paz, Bolivia): Tectonics, v. 7, Paleogene synorogenic sedimentation in the Alti- Journal of Geophysical Research, v. 99, p. 13,941– p. 23–39, doi: 10.1029/TC007i001p00023. plano plateau and implications for initial moun- 13,956, doi: 10.1029/94JB00461. Sempere, T., 1995, Phanerozoic evolution of Bolivia and ad- tain building in the central Andes: Geological Matos, R., 2002, Mapeo geológico de La Paz sur, Cordi llera jacent regions, in Tankard, A.J., Suárez, R., and Welsink, Society of America Bulletin, v. 113, p. 1387–1400, doi: Oriental de Bolivia, in Argollo, J., ed., Revista Bo- H.J., eds., Petroleum Basins of South America: Ameri- 10.1130/0016-7606(2001)113<1387:PSSITA>2.0.CO;2. liviana de Geociencias: La Paz, Bolivia, Universidad can Association of Petroleum Geologists Memoir 62, Horton, B.K., Hampton, B.A., LaReau, B.N., and Baldellon, Mayor de San Andrés, p. 34–44. p. 207–230. E., 2002, Tertiary provenance history of the north- McBride, S.L., Clark, A.H., Farrar, E., and Archibald, D.A., Sempere, T., Hérail, G., Oller, J., and Bonhomme, M.G., 1990, ern and central Altiplano (central Andes, Bolivia): A 1987, Delimitation of a cryptic Eocene tectono-thermal Late Oligocene-early Miocene major tectonic crisis and detrital record of plateau-margin tectonics: Journal of domain in the Eastern Cordillera of the Bolivian Andes related basins in Bolivia: Geology, v. 18, p. 946–949, Sedimentary Research, v. 72, p. 711–726, doi: 10.1306/ through K-Ar and 40Ar-39Ar step-heating: Journal of the doi: 10.1130/0091-7613(1990)018<0946:LOEMMT> 020702720711. Geological Society, v. 144, p. 243–255, doi: 10.1144/ 2.3.CO;2. Ingersoll, R.V., 1990, Actualistic sandstone petrofacies: gsjgs.144.2.0243. SERGEOMIN (Servicio Nacional de Geología y Técnico de Discriminating modern and ancient source rocks: McQuarrie, N., 2002, The kinematic history of the central Minas), 1997, Mapas tematicos de recursos minerales Geology, v. 18, p. 733–736, doi: 10.1130/0091-7613 Andean fold-thrust belt, Bolivia: Implications for de Bolivia, hoyas La Paz y Copacabana: Servicio Na- (1990)018<0733:ASPDMA>2.3.CO;2. building a high plateau: Geological Society of America cional de Geologia y Mineria, scale 1:250,000. Ingersoll, R.V., Bullard, T.F., Ford, R.L., Grimm, J.P., Bulletin, v. 114, p. 950–963, doi: 10.1130/0016-7606 Smith, G.A., 2000, Recognition and signifi cance of stream- Pickle, J.D., and Sares, S.W., 1984, The effect of grain (2002)114<0950:TKHOTC>2.0.CO;2. fl ow-dominated piedmont facies in extensional basins : size on detrital modes: A test of the Gazzi-Dickinson McQuarrie, N., and DeCelles, P., 2001, Geometry and struc- Basin Research, v. 12, p. 399–411, doi: 10.1046/ point-counting method: Journal of Sedimentary Petrol- tural evolution of the central Andean backthrust belt, j.1365-2117.2000.00125.x. ogy, v. 54, p. 103–116. Bolivia: Tectonics, v. 20, p. 669–692, doi: 10.1029/ Suárez, R., and Diaz, E., 1996, Léxico estratigráfico de Ingersoll, R.V., Kretchmer, A.G., and Valles, P.K., 1993, 2000TC001232. Bolivia: Revista Técnica de Yacimientos Petrolíferos The effect of sampling scale on actualistic sandstone McQuarrie, N., Horton, B.K., Zandt, G., Beck, S., and Fiscales Bolivianos, v. 17, p. 7–227. petrofacies: Sedimentology, v. 40, p. 937–953, doi: DeCelles, P.G., 2005, Lithospheric evolution of the Swanson, K.E., Noble, D.C., McKee, E.H., Sempere, T., 10.1111/j.1365-3091.1993.tb01370.x. Andean fold-thrust belt, Bolivia, and the origin of Martinez, C., and Cirbian, M., 1987, Major revisions Isacks, B.L., 1988, Uplift of the central Andean plateau the central Andean plateau: Tectonophysics, v. 399, in the age of rock units and tectonic events in the north- and bending of the Bolivian orocline: Journal of Geo- p. 15–37, doi: 10.1016/j.tecto.2004.12.013. ern Altiplano basin of Bolivia: Geological Society of physical Research, v. 93, p. 3211–3231, doi: 10.1029/ McQuarrie, N., Barnes, J.B., and Ehlers, T.A., 2008, Geo- America Abstracts with Programs, v. 19, p. 456. JB093iB04p03211. metric, kinematic, and erosional history of the central Troëng, B., Soria, E., Claure, H., Mobarec, R., and Murillo, Jiménez, N., and López-Velásquez, S., 2008, Magmatism Andean Plateau, Bolivia (15–17°S): Tectonics, v. 27, F., 1994, Descubrimiento de basamento Precámbrico in the Huarina belt, Bolivia, and its geotectonic im- doi: 10.1029/2006TC002054. en la Cordillera Occidental Altiplano de los Andes plications: Tectonophysics, v. 459, p. 85–106, doi: Miall, A.D., 1977, A review of the braided-river depositional Bolivianos: XI Congreso Geológico de Bolivia, Santa 10.1016/j.tecto.2007.10.012. environment: Earth-Science Reviews, v. 13, p. 1–62, Cruz, Memorias, p. 231–237. Jiménez De Ríos, G., 1992, Acerca del basamento Pre- doi: 10.1016/0012-8252(77)90055-1. Uba, C.E., Heubeck, C., and Hulka, C., 2005, Facies cámbrico en el bloque Andino de Bolivia: Sociedad Miall, A.D., 1985, Architectural-element analysis: A new analysis and basin architecture of the Neogene Sub- Geo lógica Boliviana, Boletín (X Congreso Geológico method of facies analysis applied to fl uvial deposits: andean synorogenic wedge, southern Bolivia: Sedi- Boliviano, La Paz), v. 27, p. 88–93. Earth-Science Reviews, v. 22, p. 261–308, doi: 10.1016/ mentary Geology, v. 180, p. 91–123, doi: 10.1016/ Jordan, T.E., 1995, Retroarc foreland and related basins, 0012-8252(85)90001-7. j.sedgeo.2005.06.013. in Busby, C.J., and Ingersoll, R.V., eds., Tectonics of Miall, A.D., 1996, The Geology of Fluvial Deposits: Berlin, Wörner, G., Hammerschmidt, K., Henjes-Kunst, F., Lezaun, Sedimentary Basins: Cambridge, Blackwell Science, Springer-Verlag, 581 p. J., and Wilke, H., 2000a, Geochronology (40Ar/39Ar, p. 331–362. Newell, A.J., Tverdokhlebov, V.P., and Benton, M.J., 1999, K-Ar and He-exposure ages) of Cenozoic magmatic Kelly, S.B., and Olsen, H., 1993, Terminal fans—A review Interplay of tectonics and climate on a transverse fl u- rocks from northern Chile (18–22°S): Implications with reference to Devonian examples: Sedimentary vial system, Upper Permian, Southern Uralian foreland for magmatism and tectonic evolution of the central Geology, v. 85, p. 339–374, doi: 10.1016/0037-0738 basin, Russia: Sedimentary Geology, v. 127, p. 11–29, Andes: Revista Geológica de Chile, v. 27, p. 205–240. (93)90092-J. doi: 10.1016/S0037-0738(99)00009-3. Wörner, G., Lezaun, J., Beck, A., Heber, V., Lucassen, F., Kennan, L., Lamb, S., and Rundle, C., 1995, K-Ar dates Nichols, G.J., 1987, Structural controls on fl uvial distribu- Zinngrebe, E., Rössling, R., and Wilke, H.G., 2000b, from the Altiplano and Cordillera Oriental of Bolivia: tary systems—The Luna System, northern Spain, in Precambrian and Early Paleozoic evolution of the Implications for Cenozoic stratigraphy and tectonics: Ethridge, F.G., Flores, R.M., and Harvey, M.D., eds., Andean basement at Belen (northern Chile) and Cerro Journal of South American Earth Sciences, v. 8, Recent Developments in Fluvial Sedimentology: So- Uyarani (western Bolivia Altiplano): Journal of South p. 163–186, doi: 10.1016/0895-9811(95)00003-X. ciety of Economic Paleontologists and Mineralogists American Earth Sciences, v. 13, p. 717–737, doi: Lamb, S., and Hoke, L., 1997, Origin of the high plateau in Special Publication, p. 269–277. 10.1016/S0895-9811(00)00056-0. the central Andes, Bolivia, South America: Tectonics, Pareja, J., Vargas, C., Suárez, R., Ballón, R., Carrasco, R., v. 16, p. 623–649, doi: 10.1029/97TC00495. and Villaroel, C., 1978, Mapa geológico de Bolivia y MANUSCRIPT RECEIVED 13 JULY 2009 Lehmann, B., 1978, A Precambrian core sample from the memoria explicativa: Yacimientos Petrolíferos Fiscales REVISED MANUSCRIPT RECEIVED 16 OCTOBER 2009 MANUSCRIPT ACCEPTED 23 OCTOBER 2009 Altiplano/Bolivia: Geologische Rundschau, v. 67, Bolivianos y Servicio Geológico de Bolivia, scale p. 270–278, doi: 10.1007/BF01803266. 1:1,000,000. Printed in the USA

1462 Geological Society of America Bulletin, September/October 2010