Você está na página 1de 26

Basin Research (2006) 18, 145170, doi: 10.1111/j.1365-2117.2006.00291.

Evolution of the late Cenozoic Chaco foreland basin, Southern Bolivia


Cornelius Eji Uba 1 , Christoph Heubeck and Carola Hulka
Institut fr Geologische Wissenschaften, Freie Universitt Berlin, Berlin, Germany

ABSTRACT
Eastward Andean orogenic growth since the late Oligocene led to variable crustal loading, exural subsidence and foreland basin sedimentation in the Chaco basin.To understand the interaction between Andean tectonics and contemporaneous foreland development, we analyse stratigraphic, sedimentologic and seismic data from the Subandean Belt and the Chaco Basin.The structural features provide a mechanism for transferring zones of deposition, subsidence and uplift.These can be reconstructed based on regional distribution of clastic sequences. Isopach maps, combined with sedimentary architecture analysis, establish systematic thickness variations, facies changes and depositional styles.The foreland basin consists of ve stratigraphic successions controlled by Andean orogenic episodes and climate: (1) the foreland basin sequence commences between 27 and 14 Ma with the regionally unconformable, thin, easterly sourced uvial Petaca strata. It represents a signicant time interval of low sediment accumulation in a forebulge-backbulge depocentre. (2) The overlying 14^7 Ma- old Yecua Formation, deposited in marine, uvial and lacustrine settings, represents increased subsidence rates from thrust-belt loading outpacing sedimentation rates. It marks the onset of active deformation and the underlled stage of the foreland basin in a distal foredeep. (3) The overlying 7^6 Ma- old, westerly sourced Tariquia Formation indicates a relatively high accommodation and sediment supply concomitant with the onset of deposition of Andeanderived sediment in the medial-foredeep depocentre on a distal uvial megafan. Progradation of syntectonic, wedge- shaped, westerly sourced, thickening- and coarsening-upward clastics of the (4) 6^2.1 Ma- old Guandacay and (5) 2.1 Ma-to -Recent Emborozu Formations represent the propagation of the deformation front in the present Subandean Zone, thereby indicating selective trapping of coarse sediments in the proximal foredeep and wedge-top depocentres, respectively. Overall, the late Cenozoic stratigraphic intervals record the easterly propagation of the deformation front and foreland depocentre in response to loading and exure by the growing Intra- and Subandean fold-and-thrust belt.

INTRODUCTION
Foreland basin systems develop as a result of exural warping of the lithosphere in response to supralithospheric and sublithospheric orogenic wedging (DeCelles & Giles, 1996; Pner et al., 2002). Lithospheric exure under static loads generates down-bending exure proximal to the orogen, which migrates as the load advances. Foreland basins therefore exhibit a characteristic asymmetric cross- section. Their sedimentary ll generally preserves and records a detailed exural response of the continental lithosphere to orogenic loading (Beaumont, 1981; Jordan, 1981; Tankard, 1986). The lithospheric response to thrusting varies between and within the foreland basin system
Correspondence: Cornelius Eji Uba, Institut fr Geologische Wissenschaften, Freie Universitt Berlin, Malteserstrasse 74 100, 12249 Berlin, Germany. E-mail: uba@geo.uni-potsdam.de 1 Present address: Institut fr Geowissenschaften, Universitt Potsdam, Karl-Liebknecht Str. 24/25,14476 Potsdam/Golm, Germany.
r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd

but is mainly controlled by the elastic thickness of the lithosphere and the applied loads (Watts, 1992, 2001). DeCelles & Giles (1996) characterized foreland basin systems into four dierent depocentres: wedge-top, foredeep, forebulge and backbulge. Each depocentre exhibits distinctive internal architecture, sedimentology and structure. Accommodation space is created by combined static and dynamic subsidence (DeCelles & Giles, 1996; Catuneanu et al., 1997). The Chaco foreland basin of the central South America is a classic example of a foreland basin system in a retro arc position. It can be subdivided into the Interandean Zone, the Subandean Zone and the Chaco plain tectono morphologic units (Uba et al., 2005) (Fig. 1). The basin formed during the late Cenozoic (Sempere et al., 1990; DeCelles & Horton, 2003) in response to Nazca-South American plate convergence and its related eastward interaction with the Brazilian shield. Detailed structural studies in the Interandean and Subandean Zones documented structural styles and timing of deformation (Sempere et al., 1990;

145

C. E. Uba et al.

Fig. 1. Tectonic map of Bolivia modied after Baby et al. (1995, 1997), Kley et al. (1997), and Mller et al. (2002) showing major morphotectonic divisions and location of the study area (in square).

Baby et al., 1992, 1997; Belotti et al., 1995; Dunn et al., 1995; Roeder & Chamberlain, 1995; Welsink et al., 1995; Moretti et al., 1996; Kley et al., 1996, 1999; Mller et al., 2002; Echavarria et al., 2003). The west-to - east variation in heat ow and thermal gradient, coupled with the hydrocarbon potential, thermal maturity and exhumation history of the Subandean Zone and Chaco plain, have also been examined (Baby et al., 1995; Dunn et al., 1995; Moretti et al., 1996; Somoza, 1998; Husson & Moretti, 2002; Echavarria et al., 2003; Ege, 2004). Furthermore, the uplift history of the central Andes has been analysed and documented (Isacks, 1988; Gubbels et al., 1993; Kennen et al., 1997; Gregory-Wodzicki, 2001). Echavarria et al. (2003) examined the accumulation rates of Neogene strata in the Subandean Zone and postulated two-phase deformation, similar to the documented three-phase tectonic subsidence history in the Chaco basin by Coudert et al. (1995). However, not much has been done on the foreland basin evolution and the dynamic interaction between the Andean fold^thrust-belt and the foreland basin. The internal architecture of a foreland basin ll, as well as its age, lithofacies and thickness, reects how the leading edge of the orogen advances over the foreland basin in time and space. These geometries and ages are best studied through a combination of outcrop, seismic, and well data that provide, in their combination, a unique archive of basin response to orogenic growth (e.g. Schlunegger et al., 1997; Alves et al., 2003; Echavarria et al., 2003). Our

study utilizes 2-D industry seismic data, logs from selected exploration wells, and stratigraphic proles from outcrops to establish such history for the Chaco foreland basin of southeastern Bolivia. Our objectives are (1) to examine the spatial and temporal changes in facies, texture and sedimentation rate of the foreland strata, (2) to explore links between stages of Andean structural evolution and basin- stratigraphic response and (3) to dene and characterize temporal and spatial distribution and migration of the depocentre as related to the deformational front.These objectives will contribute to a better understanding of tectonic processes because the stratigraphic sections record the basinward propagation of thrust sheets and their leading depositio nal systems. Furthermore, a deformational sequence is straightforward to examine and to interpret in situations where variations in sediment composition, palaeo -drainage network, uvial patterns and thickness can be taken as a proxy to accumulation and sedimentation rate (Wadworth et al., 2003; Jones et al., 2004).

GEOLOGICAL HISTORY
The Subandean fold-and-thrust belt and the Chaco foreland basin
The Andean orogeny created and subsequently progressively deformed the Chaco foreland basin since the late

146

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin

Fig. 2. Geological and structural map of the study area (modied from Suarez-Soruco, 2000) showing the data set and localities of measured sections mentioned in the text: 1, Abapo; 2,Tatarenda; 3, Saipuru; 4, Piriti; 5, San Antonio; 6, Oquitas; 7, Choreti; 8, Itapu; 9, Ivoca; 10, Cuevo; 11, Boyuibe; 12, Iguamirante; 13, Machareti; 14, Angosto del Pilcomayo (Villamontes); 15, Puesto Salvacion; 16, Zapaterimbia; 17, Rancho Nuevo; 18, Sanadita; 19, San Telmo; 20, Nogalitos; 21, Emborozu .

Oligocene (Oller, 1986; Sheels, 1988; Sempere et al., 1990; Baby et al., 1992; He rail et al., 1996). This foreland basin forms the easternmost part of the Andean orogen, which developed from the Cretaceous in the Altiplano (Horton & DeCelles, 2001; DeCelles & Horton, 2003) as a result of subduction of the Nazca plate below the South American plate and the simultaneous subduction of the Brazilian Shield at an initially low rate of 5.8 cm year 1 to a subsequent maximum rate of up to 15.2 cm year 1 (in the late Oligocene; Somoza, 1998). Andean deformation commenced in the west with formation of the Altiplano basin and foreland sedimentation, with the depocentre probably at the present Eastern Cordillera (Sempere et al., 1997; DeCelles & Horton, 2003; Elger et al., 2005). During the Cretaceous-Eocene, the area of present-day southern Bolivia

was already part of a foreland system but was presumably included in a large intracontinental plain of non-deposition. The geology of the Bolivian Andes is classied into six tectonomorphic units, of which three units participate in the late Cenozoic foreland system (Fig. 1). Sedimentary units pertaining to the Chaco basin occur (west to east) from the Inter-Andean Fault (IAT), through the Subandean Zone, and below the Chaco plain to its onlap on the Brazilian Shield and the Alto de Izozog basement high. The western part of this basin is deformed by the Subandean fold-and-thrust-belt and is still undergoing active shortening at its leading edge (Fig. 2; Oller, 1986; Sheels, rail et al., 1996). Late Cenozoic se1988; Baby et al., 1992; He dimentary strata are commonly well exposed along anks

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

147

C. E. Uba et al.

Fig. 3. Structural balanced cross- section of the Central Andes from Altiplano to Chaco plain at 211S. Modied after Kley et al. (1999) and Elger et al. (2005). UKFZ, Uyuni-Khenayani Fault Zone; SVT, SanVicenteThrust; CYT, Camarga-Yavi Thrust; IAT, Interandean Thrust; SAT, Subandean Thrust; PF, Pajonal Fault; PBF, Palos Blanco Fault. See Fig. 2 for location.

of the major leading syn- and anticlines near the wedgetip of the Subandean zone. This region, between 64130 0 and 62130 0 E and 18145 0 and 22130 0 S, forms the principal study area. Chaco foreland basin sedimentation is assumed to have begun approximately 27 Ma ago near the Eastern Cordillera (Sempere et al., 1990), as a result of eastward migration of the deformation front ahead of the Interandean and Subandean Zones (Sempere et al., 1990; Husson & Moretti, 2002; DeCelles & Horton, 2003; Echavarria et al., 2003; Ege, 2004). Late Cenozoic strata show characteristic westward thickening. It stands to reason that Chaco foreland basin strata had also been deposited in considerable thickness in the region occupied by the present-day Subandean Zone before its uplift and incorporation into the eastwardmigrating orogenic wedge. Although these deposits are well preserved in the Subandean Zone, in some areas the coarse-grained proximal foreland basin deposits have been eroded. The eroded proximal basin sedimentation is as a result of erosion removal after deformation and subsequent propagation of the fold^thrust belt (Burbank & Raynolds, 1988). A rough estimate of their original thickness (ca. 3^5 km) can be obtained by reconstructing thermalgradient- calibrated sedimentary thickness from AFT samples of the youngest Mesozoic strata in the Subandean Zone (Ege, 2004).

Magnitude and timing of shortening


Lithospheric thickening and corresponding shortening in the fold-and-thrust belt of the Subandean zone, reconstructed from structural balanced cross- sections (e.g. Sempere et al., 1990; Kley et al., 1996, 1997), began east of the Eastern Cordillera in the late Miocene. However, widespread shortening there started only in the Oligocene (Baby et al., 1992; Gubbels et al., 1993; Dunn et al., 1995; Kley, 1996; Jordan et al., 1997; Kley et al., 1997; McQuarrie, 2002). Since then, continuous eastward propagation of thrusting, accompanied by large- scale folding, produced a generally eastward-younging synorogenic wedge (Moretti et al., 1996; DeCelles & Horton, 2003; Echavarria et al., 2003).

During the late Oligocene, the Eastern Cordillera was the focus of pronounced shortening (Kley et al., 1997; Horton, 1998). Figure 3 shows structural styles and major thrust sheets, illustrating that the Subandean Zone is deformed by mostly in- sequence, thin- skinned thrust sheets that include north-northeast-trending ramp anticlines and passive roof duplexes (Baby et al., 1992, 1997; Belotti et al., 1995; Dunn et al., 1995; Kley et al., 1996, 1999; Echavarria et al., 2003). This progressive thin- skinned deformation is recorded in a suite of angular unconformities and stratigraphically distinct foreland packages. A total shortening of 210^336 km is postulated for the Central Andes (Baby et al., 1992; Moretti et al., 1996; McQuarrie & DeCelles, 2001; Mller et al., 2002; Elger et al., 2005) together. The Interandean and Subandean Zones take up 140 and 86 km shortening at 201S and 221S, respectively (Baby et al., 1997). This matches well with a total shortening of 140 km at 211S in the Interandean and Subandean zones together (Kley et al., 1997). Moretti et al. (1996) calculated a peak shortening rate between 6 and 2.1 Ma, followed by a minimum shortening rate between 2.1 Ma and the present in the Subandean Zone. Their values, however, disagree with the estimates by Echavarria etal. (2003), who postulate two periods of high shortening rates (11 and 8 mm year 1) at 9^7 and 2^0 Ma, respectively, separated by an in-between low of 0^5 mm year 1 at 221300 latitude.These contradictions may be due to the paucity of direct age dates for the deformation and the inherent variability of geologic cross- section construction and interpretation. The variations in shortening values (Moretti et al., 1996; Echavarria et al., 2003) and the resulting inferred time periods of uplift (Sempere et al., 1990; Baby et al., 1992, 1997; Dunn et al., 1995, Kley et al., 1997) suggest diachronous movement on individual thrust sheets. For example, Echavarria et al. (2003) attribute the 2-Ma- shortening event to thrust reactivation in the south-western Subandean Zone ( 22130 0 ), whereas Moretti et al. (1996) interpreted the 2.1 Ma shortening as a major displacement event synchro nous with folding and uplift of the leading Aguarage range (Fig. 2). No data are available to constrain the time of formation of the ramp anticlines of the western Chaco plain, which are clearly visible on industry reection- seis-

148

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin

Fig. 4. Stratigraphy of the Subandean Zone and Chaco basin ll and rose diagrams summarize the palaeocurrent directions. Ages are based on Marshall et al. (1993), Moretti et al. (1996), Echavarria et al. (2003), Hulka (2005), and Hulka et al. (in press).

mic proles but have as yet only an indistinct and low to pographic expression. Anticline cores of these structures are morphologically expressed in a N^S trend of low hills, cut by east^west-trending gullies.These structures appear to be actively forming and indicate the continuous eastward growth of the Andes onto the Brazilian Shield.

DATA AND METHODS


The data set compiled for this study includes measured sections, seismic data and well logs. T wenty- one stratigraphic sections along major rivers, small streams and road cuts in the Subandean foothills were measured and sampled for lithologic, sedimentologic and biostratigraphic data (Fig. 2) to document architectural style and basin geometry. In addition, we interpreted 45 wire-line logs and their well reports from hydrocarbon industry exploration wells and tied them to 42800 km of 2-D industry seismic proles. Wire-line logs (g-ray, resistivity and sonic), combined with well reports, provide ne vertical details of wells and lithology resolution and thus complement seismic data for a better understanding of the subsurface geology. Similar methods were used by Schlunegger et al. (1997) and Alves et al. (2003) to study the Upper Marine Molasse Group of the North Alpine foreland basin and the Lusitanian rift basin of West Iberia, respectively. Wire-line log analysis and well reports were used in combination with seismic facies attributes to delineate dierent stratigraphic packages and to correlate them to the ve late Cenozoic formations. The seismic data, wire-line logs and well reports were provided by Chaco S.A. and Yacimientos Petroleros Fiscales de Bolivia (YPFB), Santa Cruz. The seismic lines cover mostly the Chaco plain where outcrop is poor or absent, and partially extend into the

foothills of the Subandean Zone. They dene the regional stratigraphic architecture of the late Cenozoic basin ll. We used regional isopach trends as a proxy for accommo dation space (e.g. Wadworth et al., 2003), and traced their thickness variations from vertical facies associations. Synthetic seismogram and check- shots from well logs were used to perform time-to -depth conversion from two way-travel time (TWT, in ms).

FORELAND LITHOSTRATIGRAPHY
The up to 7.5-km-thick (Emborozu section), eastwardthinning strata of the Chaco foreland basin are largely composed of siliciclastic non-marine redbeds with minor shallow-marine strata. We used a detailed stratigraphy after Suarez Soruco (2000) that is principally based on lithology, with only minor modications (Fig. 4).The basin ll includes (from base to top) the Petaca, Yecua, Tariquia, Guandacay and Emborozu Formations. Age dating of these formations has proved dicult and principally relies on a combination of mammal biostratigraphy and radio metric dating of rare tus (e.g. Marshall et al., 1993; Marshall & Sempere, 1991; Moretti et al., 1996; Echavarria et al., 2003; Hulka, 2005). Notwithstanding the recent by published new 40Ar/39Ar radiometric ages for the late Cenozoic units in the Bolivian Subandean zone by Hulka (2005), no complete and precise chronology for the basin ll is yet available. Therefore, we used published ages to document the late Cenozoic lithostratigraphy of the southern Bolivia. However, the ages should be applied with caution. Parts of the formations are suspected to be diachronous, not only younging west-to - east, as could be expected, but possibly also north-to - south (Echavarria et al., 2003). In addition, the stratigraphy is complicated by several nearly basinwide low-angle unconformities.

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

149

C. E. Uba et al.

Fig. 5. Selected outcrop photographs showing (a) clast- supported reworked pedogenic conglomerate facies of the Petaca Fm (see hammer in circle for scale). (b) Shallow-marine-lacustrine mudstone-dominated facies with thin-bedded ooid-, shell hash-dominated sandstone bed of the Yecua Fm (arrow). (c) Channelized sandstone beds with desiccation cracks (arrow) of theTariquia Fm. (d) Alternation of conglomerate and sandstone beds of the Guandacay Fm. (e) Sheet-like cobble-boulder-dominated conglomerate bed with thin beds of sandstone of the Emborozu Fm.

Petaca formation
Cenozoic sedimentation in the Chaco basin commenced during the late Oligocene (assumed ca. 27 Ma; Marshall et al., 1993; Moretti et al., 1996) with the deposition of the

up to 250-m-thick Petaca Formation (Gubbels et al., 1993; Sempere, 2000). This formation unconformably overlies Mesozoic eolian strata (Sempere, 1995). The lower part of the Petaca Formation consists of greenish grey, white and light purple basal calcrete. The calcrete consist of isolated

150

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin

Fig. 6. Correlated proles of stratigraphic sections of the late Cenozoic strata in the northern study area 191S. See Fig. 2 for location.

to clustered, blocky to massive and bracciated nodules (Uba et al., 2005).The calcrete body is overlain by horizontal to disorganized clast- supported reworked-pedogenic conglomerate (Fig. 5a) composing of poorly sorted, densely packed clasts of poorly rounded intraformational reworked calcrete nodules and subordinate chert. The conglomerates show sharp and erosive bases. Medium to very- coarse-grained sandstone and sandy to ne-grained mudstone mark the upsection lithology of the Petaca Formation.The calcareous, red to grey, bioturbated sandstone is characterized by tabular to lenticular beds, trough cross-, planar and horizontal stratications, as well as rip-up clasts at the base. The massive, laminated mud-

stone bodies have bioturbation, minor desiccation cracks and padogenesis. The formation thins towards the centre of the study area (Villamontes-Camiri axis, Fig. 6).The reworked pedogenic conglomerate and sandstone bodies show channel and bedform architectural elements and an overall ning-upward sequence. Cross- stratication in sandstones indicates a westward-directed drainage (Fig. 4). Uba et al. (2005) attributed the thick calcrete horizons to well-developed palaeosols, indicating 0 or low sedimentation in an arid to semiarid climate in which evaporation generally exceeded precipitation (e.g. Cecil, 1990). The lithofacies and architectural elements in the Petaca conglomerate and sandstone indicate variable high- energy

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

151

C. E. Uba et al.

Fig. 7. Correlated proles of stratigraphic sections of the late Cenozoic rocks in the southern study area 211S. See Fig. 2 for location.

stream ows in a channelized setting. Uba et al. (2005) and Marshall et al. (1993) interpret the Petaca strata as having been deposited by braided stream. The ning-upward trend and changes in bedform represent a decrease in ow strength or depth as a result of waning of ood intensity (Miall, 1996). The occasional occurrence of successions of palaeosols indicates predominantly non-deposition and surface exposure. This is supported by desiccation marks, bioturbation and purple colour (e.g. Miall, 1996; Retallack, 1997). The contact between the Petaca Formation and underlying eolian strata is a regional erosional unconformity that may have formed as a far- eld response to early Andean tectonics (Sempere etal.,1990; Dunn etal.,1995). Marshall et al. (1993) reported reptilian and mammal bone

fragments of late Oligocene to late Miocene age, found close to the Aguarague range in conglomerate (Sempere et al.,1990; Marshall & Sempere,1991). However, as the age of the basal calcretes has not been ascertained, the onset of deposition is poorly constrained.

Yecua formation
The up to 600 -m-thick Yecua Formation (Padula & Reyes, 1958) overlies the Petaca Formation with an indistinct lowangle erosional unconformity.The Yecua Formation shows a west-to - east and northeast-to - southwest facies variation. North of Camiri, it consists of red-green to brown sandstone^mudstone couplets (Fig. 5b) showing herring-

152

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin


bone cross- stratication, laminated, convolute, aser, wavy and lenticular bedding in ning- and coarseningupward successions. This lithofacies also consists of gypsum veins, syndepositional structures, bioturbation and desiccation marks (see Fig. 5b). Fossils include bivalves, the foraminifera Globigerinacea and Corbicula, the ostracodes genera Cypridelis and Heterocypris, pelecypods, gastropods, cirripeds, decapods, crabs, sh skeleton fragments, ooids, shell hash and terrestrial plants (Marshall & Sempere, 1991; Marshall et al., 1993; Hulka et al., in press). In the western and southern part of the study area (south of Camiri), the Yecua Formation consists of red to light brown, lenticular and very ne- to medium-grained sandstone interbedded with red to light-brown, ripple-laminated sandstone couplets. The proportion of mudstone bodies dominate over the sandstone (Fig. 5b). The sandstones show erosive channel structure and ning-upward trends. These sand bodies contain cross-bedding, climbing ripples, gypsum veins, rip-up clasts and burrows. Mudstone^ sandstone couplets contain mottled soil, desiccation cracks and extensive burrows. In general, the sandstone proportion, bed thickness and the channel pro portion of the Yecua Formation increase upsection and to wards the west (Figs 6 and 7). We interpret the fossiliferous and varicoloured Yecua facies in the north of the study area as deposits of lacustrine, tidal, shoreline and brackish to shallow marine environments, in agreement with the previous work by Marshall et al. (1993), Hulka et al. (in press) and Uba et al. (2005), and is supported by the presence of lacustrineshallow marine fossils and the lithofacies. The mudstonedominated terrestrial facies of the Yecua Formation to the west and south are products of uvial overbank and channel processes, with occasional lacustrine and mudat settings. Hulka et al. (in press) placed these variations in a regional context and argued that the marginal marine facies of the Yecua Formation represented a marine incursion from the northeast along the axis of the developing foreland basin as far as Camiri.The age of the Yecua strata has variably been estimated based on ostracodes and foraminifera to be 14^7 Ma (Padula & Reyes, 1958; Marshall et al., 1993; Hulka et al., in press) and 11^7 Ma (Moretti et al., 1996). Recently published 40Ar/39Ar radiometric ages of 10.49 0.33 and 9.41 0.52 Ma (Hulka, 2005) on interbedded tus in the Yecua Formation uvial facies from the Emborozu and Nogalitos sections matches the estimated biostratigraphic age from the marine facies. By analogy, these ages are considered herein to correlate with the uvial-and-lacustrine Yecua- equivalent strata (Tariquia Formation of Bolivian nomenclature; Russo, 1959; Ayaviri, 1964; Moretti et al., 1996; Suarez Soruco, 2000) near Argentinas border with Bolivia that yielded an age of 9.95 0.34 Ma (Echavarria et al., 2003). dational contact. The Tariquia Formation is characterized by thick- and thin-bedded sandstone bodies interbedded by laminated mudstone and very ne-grained sandstone (Uba et al., 2005). The light brown, light yellow and red, well- sorted, very ne- to medium-grained sandstone bodies range between 0.5 and 15 m thickness and consist of sharp erosional base, ribbon and channel geometry (Fig. 5c), and extend laterally for hundreds of meters. Sandstone units have massive bedding, planar, trough cross- and climbing ripple sedimentary structures. Intraformational rip-up clasts and reworked calcareous no dules are common. The sandstone bodies have multistorey channel architecture and an overall coarseningand thickening-upward trend (Fig. 5d).There is an upward increase in the degree of vertical stacking, bed thickness and lateral interconnectedness in the sandstone unit. Overall, the mean grain size, channel interconnectedness, sandstone proportion and thickness of the Tariquia Formation increase towards the west (Figs 6 and 7). The massive, laminated- or ripple- stratied interbedded mudstone^ sandstone couplets show sheet geometry and are laterally extensive. Taenidium barreti trace fossils (Buatois et al., in press) in the thick-bedded channelized sandstone and in mudstone and sandstone couplets are common. The trace fossils disrupt the primary sedimentary structures. A distinguishing feature of the Tariquia Formation is the presence of abundant mudcracks (Fig. 5c, arrow), occasional syndepositional deformation, and poorly developed palaeosols that are more dominant in the mudstone^ sandstone couplets. Palaeocurrent measurements indicate a mean transport towards the east (Fig. 4). In theTariquia Formation, the sandstone proportion and size and bed thickness increase towards the west (Figs 6 and 7). The Tariquia Formation is interpreted to represent a range of processes that operate in a large uvial system (Uba et al., 2005). Thick-bedded sandstones were deposited within major channels, whereas the thin-bedded sandstone units indicate deposits from crevasse channels. In overbank areas, mudstone and sheet sandstone were deposited by crevassing splay and suspension fallout. Following Uba et al. (2005), we interpret theTariquia Formation as a product of a low-gradient, high- sedimentation, channelized anastomosing stream and associated thick oodplains on a distal uvial megafan. This interpretation is supported by the laterally extensive channel geometry, aggrading thick oodplain deposit (ponded area), vertical channel stacking, frequent crevassing and avulsion, and a general lack of lateral channel migration architecture (e.g. Smith, 1986; Makaske et al., 2002; Bridge, 2003). The inferred depositional processes and lithofacies are similar to active modern meganfans in the Chaco plain that receive sediment from large uvial networks (Horton & DeCelles, 2001). Avulsion and crevassing result in the development of new channels on the overbank, whereas the active channels are abandoned (Smith, 1986; Makaske etal., 2002).The abundant rip -up clasts that may have been formed by erosional scouring of overbank sediments and

Tariquia formation
TheTariquia Formation (Russo,1959; Ayaviri,1964) is up to 3800-m-thick and overlies the Yecua Formation with gra-

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

153

C. E. Uba et al.
high degree of bioturbation in theTariquia Formation suggest long periods of channel abandonment and colonization by insects (Buatois et al., in press). The welldeveloped upward coarsening and thickening trend suggests a systematic stratigraphic development governed by either long-term eastward propagation of the fold^thrustbelt and/or the expansion of drainage networks. Uba et al. (2005) postulated a shift in climate from a semi-arid to a humid condition during the deposition of the Tariquia strata. The Tariquia Formation age is late Miocene (Chasicoan-Huaquerian) based on biostratigraphy (Marshall & Sempere, 1991), in agreement with a single apatite ssion-track age of 7 Ma from Mica (Moretti et al., 1996). In addition, Moretti et al. (1996) assumed 6 Ma as the upper age limit of the Tariquia Formation. In the absence of a well- constrained age for this unit, we use the imprecise age of 7^6 Ma for the deposition duration for this formation. Celles & Horton, 2003). Lenses of coal suggest the presence of a ponded area and vegetation, and therefore, a humid palaeoclimate (Uba et al., 2005). Vertical stacking and aggradation of channels into overbank deposits imply crevassing and avulsion, indicating periodic abandonment of active channels. The weakly developed palaeosol and poorly preserved bioturbation may suggest a high overbank aggradation rate (Bridge, 2003).The contact between the Tariquia and the overlying Guandacay Formation is unconformable (Moretti et al., 1996; Echavarria et al., 2003), approximately 6 Ma in age (Moretti et al., 1996), and is marked by a distinct increase in mean grain size. Hulka (2005) estimated the top of this formation at 2.1 0.2 Ma based on 40Ar/39Ar dating of tu at its contact to the Em section (Fig. 2).The age of borozu Formation in the Abapo the Guandacay Formation is therefore late Miocene to Early Pliocene (6^2.1 Ma).

formation Emborozu Guandacay formation


The up-to -1500-m-thick Guandacay Formation consists of conglomerate, sandstone and mudstone (Jimenez-Miranda & Lopez-Murillo,1971) (Fig. 5d).The granule- cobble conglomerate shows sheet-like and lenticular geometry, clast- supported, polymictic, a coarsening- and thickening-upward trend, massive to inversely graded, well-developed imbrication and basal scour surfaces. Gravel bedforms and poorly developed lateral accretion surfaces are common architectural elements. The dominantly medium- to very- coarse-grained sheet-like sandstones are moderately to well sorted, and are laterally extensive for several hundreds of metres (Fig. 5d). The sandstone bodies consist of trough cross-, planar, ripple and horizontal stratication, and occasional stringers of pebbles. Thick interbedded mudstones and sandstone are massive to laminated and laterally continuous for several tens or hundreds of metres. Lenses of thin coal seams, poorly preserved bioturbation, and weakly developed mottled soils are present (Uba et al., 2005).The conglomerates generally thicken and coarsen upsection and to the west. The conglomerate and sandstone bodies show, like the Tariquia Formation, an upward increase in stacked packages, lateral interconnectedness, and multi- to single- storey channel systems that grade into the interbedded mudstone and sandstone. Palaeocurrent measurements indicate a northeast-to - southeast-directed ow (Fig. 4). The conglomerate and sandstone lithofacies provide evidence of deposition in uctuating, high- energy, bedload-dominated large uvial channels, anked by oodplains, and zones of incipient soil development (Uba et al., 2005). The dimensions of channel lls and the types of sedimentary structures in the Guandacay Formation suggest large discharges (e.g. Horton & DeCelles, 2001). Consequently, Uba etal. (2005) envision a proximal braided setting on a medial uvial megafan, similar to those that deposited the Camargo Formation and that drain the modern central Andean (Horton & DeCelles, 2001; DeThe Emborozu Formation (Ayaviri, 1967) is exposed only Section) and within synclines in in the northeast (Abapo the southwestern (Emborozu and Nogalitos; Fig. 2) study area. The up-to -2000-m-thick, conglomeratic upwardcoarsening strata of this formation cap the foreland stratigraphic succession in the Chaco Basin. In outcrops near the present Subandean topographic front, growth structures occur (Echavarria et al., 2003), documenting a syndeformational origin. The Emborozu Formation is dominated by an up-to -60 -m-thick, cobble-boulder conglomerate that reaches at least 153 cm in diameter (Fig. 5e; Uba et al., 2005). This laterally extensive (several hundreds of metres) conglomerate shows sharp erosive scour surfaces, sheet-like to lenticular single- channel geometry, inverse and normal grading, and moderately to poorly developed imbrications. This conglomerate lithofaces is associated with up-to 6 -m-thick, coarse- to very- coarsegrained, sheet-like sandstone with horizontal, trough cross-, ripple- and planar stratication. The singlestorey, vertically stacked conglomerate and sandstone bodies grade into medium- to very coarse-grained, rippled, massive-laminated interbedded sandstone and mudstone in which poorly preserved burrows and coal lenses occur. Upsection and to the west, the thickness, lateral continuity, amalgamation and maximum grain size of the conglomerate and sand bodies increase and the percentage of overbank nes decreases. The palaeodrainage pattern shows a northeast-to - southeast-directed ow (Fig. 4). Uba et al. (2005) interpret the Emborozu Formation as a uctuating- energy, bedload, proximal uvial system of successions of large, isolated to amalgamated channels. The presence of a thick to subordinate oodplain and the lateral extent suggests deposits on a proximal uvial megafan (Horton & DeCelles, 2001; Uba et al., 2005).The sharp scour surfaces may represent discrete channels or an amalgamation of scour as a result of avulsion events and channel abandonment. The Emborozu Formation overlies the

154

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin

Fig. 8. Major package boundaries and their characteristics recognized on seismic sections, well logs, and interpreted lithology.The base and top of each package is dened.

Guandacay Formation with a well-dened regional angular unconformity that marks its base in seismic sections (Moretti et al., 1996; Echavarria et al., 2003). Moretti et al. (1996) previously estimated the age of this contact based on Ar^Ar on mica to be 3.3 Ma. However, a new section; Ar^Ar on mica) estimate 2.1 0.2 Ma (tu; Abapo by Hulka (2005) agrees relatively well with 1.8 Ma documented by Echavarria et al. (2003) for correlative strata in Argentina. Consequently, 2.1 0.2 Ma is used herein as the basal age of the formation.

Package N1
The base of N1is a prominent, readily traceable reector of high amplitude, medium frequency and medium continuity across most of the study area, marking the contact between the late Cenozoic foreland basin and underlying Mesozoic strata (Fig. 8a). In some seismic sections, the underlying Mesozoic strata show diuse toplap and truncations with low angular geometry. In wire-line logs, this contact shows an abrupt increase from 100 to 170 O m in ILD and an immediate drop from 90 to 40 in DTresponses (e.g. Fig. 8a). A clear dierentiation can be made between the top of N1 and the base of N2 as a result of a pronounced medium- to high-amplitude, continuous and medium- to high-frequency reector that is easily identied and correlated throughout all seismic sections (Fig. 10). Wire-line logs show a sharp increase from 30 to 120 API in the GR curve and from 50 to 70 in the DT curve, coupled with an abrupt increase to 90 in the ILD curve. The thinness of this package does not allow a detailed seismic facies characterization. However, in some areas, N1 displays internally lateral extensive, low- to medium-amplitude, subparallel, discontinuous, low-frequency reectors (Fig. 10). Among them, a wedge- shaped set reaches up to ca. 200 m (  0.2 ms) thickness across a broad area in the western Chaco plain and gradually pinches out to 0 with onlap terminations upon reaching the Alto de Izozog high (Fig. 10). Wire-line logs through N1 (Fig. 11) generally indicate low GR value (30^60 API), average 90^55 DT, and 90^170 O m ILD values. The GR curves show cylindrical shape characteristic. The toplap reection terminations and truncation of N1 on the underlying Mesozoic strata indicate an uncon-

SEISMIC STRATIGRAPHY OF THE LATE CENOZOIC DEPOSITS


A good well-to - seismic tie and the lateral continuity of horizons allowed interpretation of the visible geometric features on the seismic lines. After interpreting seismic lines and wire-line logs, we subdivided the foreland-basin ll into ve regionally mappable packages, numbered sequentially N1 to N5. These are delineated by discontinuities that coincide with changes in seismic facies and that can be correlated with wire-line logs. Seismic facies attributes include prominent reectors, termination geometry (onlap, toplap, downlap and truncation), reection conguration, and external form. Not all onlap and truncation geometries could be mapped in the seismic sections due to limited vertical resolution combined with small unit thickness. Figures 8 and 9 illustrate the most characteristic seismic facies features and g -ray (GR), resistivity (ILD) and sonic (DT) log responses of the ve packages.The corresponding lithofacies and depositional environments are calibrated by well data.

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

155

C. E. Uba et al.

Fig. 9. g-ray, resistivity, and sonic records for IGR 01 well and the interpreted lithology correlated to the Angosto del Pilcomayo section located approximately 40 km farther south.

formity. It could also be a condensed or non-deposition surface with minor erosion (e.g. Mitchum et al., 1977; Sheri & Geldart, 1995). However, the quality of the seismic data and the overall low thickness of N1do not allow a clear dierentiation between toplap and erosional truncation. Distal onlap and the overall wedge form of N1 on the Alto de Izozog High indicate basin progradation. We interpret package N1 as a sand-dominated aggradational uvial system.This interpretation is based on seismic facies characteristic (variable-amplitude, discontinuous and low-

frequency reection) combined with low GR and DT and the cylindrical shape of the GR curve, implying that this package consists mainly of relatively high- energy uvial deposits (Badley, 1985; Cant, 1992; Emery & Myers, 1996). The cylindrical shape of GR logs suggests an aggrading braided uvial system (Cant, 1992; Emery & Myers, 1996). The high acoustic impedance variation between the overlying mudstone-dominated N2 and the underlying, sandstone-dominated Mesozoic rocks also suggests a change in lithology and probably a high degree of cementation or

156

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin

Fig. 10. Segment of a W-E uninterpreted and interpreted migrated seismic line along approximately 201 45 0 S showing the ve late Cenozoic sequences and thrusting and folding of the foreland sequences.The location of the line is shown in Fig. 2.

pedogenesis. We interpret the strong seismic facies and sharp wire-line log relations at the contact of N1 and N2 to reect an unconformity. We correlate N1 based on the seismic facies and wire-line log characteristics mentioned above as the subsurface equivalent of the Petaca Formation.

Package N2
The top of N2 is a laterally continuous, high-to -moderate-amplitude reector (Fig. 8a). The GR, ILD and DT logs do not show a sharp dierence but rather a gradual response at the contact to N3 (Fig. 8a). N2 is overall wedgeshaped, with a maximum thickness of more than 450 m (0.20 ms) in the west.To the east, N2 terminates with onlap geometries on the Alto de Izozog High, where it overlies N1 and pinches out onto the Mesozoic strata (Figs 10 and 11). Its internal seismic facies are: to the west, N2 displays internal seismic reections that show variable-amplitude, discontinuous, subparallel, low-frequency, low vertical spacing and chaotic pattern. Low- scale hummocky clino forms dip at variable angles ( 1^21); to the east, low- scale complex sigmoid- oblique seismic reections also occur. In contrast, to the east, low-angled clinoforms, discontinuous, low-amplitude, semi transparent and chaotic reections, coupled with low acoustic impedance, occur (Figs 8a and 10). Wire-line logs show 60^120 API in GR, 95^85 O m in ILD and 65^85 in DT values. However, GR and

DT values decrease and increase upsection, respectively. GR indicate marked thin spikes and large percent of high to low values (80 : 20) in the N2 package (Fig. 9). However, the low GR and high DT values increase upsection. GR logs show an irregular or serrated response (Cant, 1992; Emery & Myers, 1996) and small- scale variability in values as indicated by numerous thin cycles with ning-upward trends (Fig. 11). The variable-amplitude, discontinuous, semi-transparent, internal structure, higher GR and lower DTvalues are typical of a poorly stratied mudstone-dominated system, deposited mainly by suspension settling and subordinate channel settings (Cant, 1992; Alves et al., 2003). Based on the seismic facies and well-log characteristics, we can interpret the N2 package as deposition in varied settings such as shallow marine, lacustrine and uvial environments (e.g. Badley, 1985; Cant, 1992; Alves et al., 2003; Hofmann et al., 2006), with aggrading uvial setting dominantly in the west and south of the study area.The interpretation of varied-depositional settings is further supported by thin sandstone intervals in wire-line logs, a relatively high and serrated GR response, sigmoid- oblique and hummocky clinoforms, varied-amplitude and low frequency (Sangree & Widmier,1977; Badley,1985; Cant,1992; Sheri & Geldart, 1995). As N2 thickens westward towards its depocentre, it develops varied-amplitude and lowcontinuity reections and dened- clinoforms. The small- scale variability observed in the GR, DT and ILD

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

157

C. E. Uba et al.

Fig. 11. Segment of a W-E uninterpreted and interpreted migrated seismic line along approximately 19130 0 S showing pinch out of the late Cenozoic sequences.The location of the line shown is in Fig. 2.

proles largely represents variation in depositional energy associated with high-frequency cyclicity (Cant, 1992; Emery & Myers, 1996). The gradual GR and DT responses at N2^N3 contact suggest a fairly steady change of depositional environment between these two packages.The eastward thinning indicates the presence of a palaeo -high near the eastern border of the study area before the deposition of N2. The upsection decrease in the GR and increase in the DT values, which can imply upsection increase in sandstone proportion, reects a basinwide shift in facies. The characteristic mudstone-dominated seismic facies and wire-line log attributes of N2 package are analogous to the Yecua Formation.

Package N3
The base of N3 is a variable continuous and varied-amplitude reector; the GR and DTshow decrease and decrease in values, respectively (Figs 8 and 9).The laterally continuous, moderate- to high-amplitude reector marks the top of this package and a transitional contact to N4. In wireline logs, this contact is marked by a relatively sharp low GR and high ILD and DT responses (Figs 8b and 9). N3 shows a maximum thickness of 1500 m (  0.75 ms) in the western study area, thinning gradually eastward to

pinch out at the Alto de Izozog basement high, where it onlaps and overlies Mesozoic strata (Fig. 10). Internally, N3 displays varied- seismic facies; in the lower portion of the section, it shows low- to medium-amplitude, discontinuous, subparallel, low-frequency, semi-transparent and hummocky reectors (Figs 10 and 12). However, the seismic facies changes upward to more moderate-low continuous, varied-amplitude, less chaotic and less hummocky reectors and wedge- sheet external forms. The proportion of clinoform, hummocky, low-frequency reectors increases to the east. The clinoforms show eastoriented downlap onto, and appear to coalesce with, medium-amplitude reectors. The log character of the N3 package is distinguished from the underlying N2 package because it contains relatively lower GR (30^90 API), higher DT (70^100) and higher ILD (95^160) responses. In addition, it shows a thicker and larger percent of a low GR response compared with the N2 package, with the percent and thickness of low GR and DTresponses increasing signicantly upward and to the west, where it reaches tens of metres in thickness (Figs 9, 12 and 14). GR and DT curves have both serrated and occasional bell shapes. We interpret the varied-amplitude, subparallel, moderately to discontinuous seismic facies, coupled with low GR

158

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin

Fig. 12. W-E wire-line logs at about 211S illustrating aspects of sequence boundaries.The depositional sequences identied can be correlated to near outcrops. See Fig. 2 for location.

and high DT response of package N3, as alternations of sandstone and mudstone deposited by sandstone-dominated uvial deposition (e.g. Sangree & Widmier, 1977; Cant, 1992; Sheri & Geldart, 1995; Emery & Myers, 1996). The low to high amplitude, moderate^low continuity, lens- sheet external forms, and serrated-bell shape GR and DTresponses suggest channels aggrading into oodplains (Cant, 1992). This interpretation is further supported by the upsection change in seismic facies (e.g. chaotic, hummocky, semi-transparent, combined with upsection decrease in the GR values and thickness of low GR response) that indicate an upsection increase in the pro portion of sandbodies and bed thickness. The serratedand bell- shaped GR responses suggest multiple ningupward trends and variable depositional energy. The semi-transparent and chaotic seismic features represent lack of stratication. The westward-thickening wedgeshaped geometry of N3, the eastwardly oriented clino forms, and the wedge form suggest deposition by pro -

gradation from the west. The upsection increase in clino forms reectors, variable seismic characteristics and pro portion of sandstone and GR thickness within N3 suggest a strong progradational pulse concomitant with a basinward shift in facies and depocentre location. We assign the N3 package to theTariquia Formation because of interpretation of characteristic of the seismic facies expressions and wire-line logs identied in this package.

Package N4
The top of package N4 is characterized by a prominent, high-amplitude, continuous reector that can be mapped and correlated throughout all seismic sections. This top contact is marked by local toplap and truncation terminations of N4 reectors on N5 (Figs 8d and 13, inset photo), accompanied by abrupt breaks on GR, DT and ILD logs (Figs 9, 12 and 14). Figures 8d and 13 show that this contact is a well-dened angular unconformity. The base of N4 is

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

159

C. E. Uba et al.

Fig. 13. Segment of a W-E uninterpreted and interpreted migrated seismic line along approximately 211S showing the structural styles and the ve sequences.The location of the line is shown in Fig. 2.

delimited by a moderate to high-amplitude, continuous reector.Wire-line logs indicate a decrease and an increase in the values of GR and DT, respectively. The thickness variation of Package N4 is similar to that of N3, with a broad area in the west, where the thickness exceeds approximately1500 m ( 750 ms) thinning to the east to zerol at the Alto de Izozog high. At this basement high, N4 also onlaps and overlies N2, N3 and Mesozoic strata (Fig. 11). Internally, N4 shows parallel to subparallel, variable-amplitude and -frequency, and moderate^low continuity reectors (Figs 8c, 9, 10 and 13). Package N4 shows a wedgeshaped, vertically spaced reections, suggesting several tens of metre- scale bedding. In the southern part of the study area, N4 includes a growth structure near the LaVertiente Fault (Fig.12), showing an upsection decrease in the inclination of reectors and onlap geometry (Fig.12 inset). However, this fault is limited to the southern part of the study area and is not recognized further north (e.g. Fig. 10). The N4 package is identied in wire-line logs by low GR (30^70 API), high DT (105^140) and high ILD (95^ 130) responses (Figs 12 and 14). Figure 9 show that both the GR and SP log curves show an upsection decrease and increase in response, respectively, and have cylindrical

and bell shapes (Figs 9, 12 and 14). However, the percent of low GR values in N4 are relatively higher and thicker than in the underlying N3 package (Figs 9 and 14). The low GR and DT intervals have a serrated shape. The high GR and DT values, cylindrical shape and thickness, combined with parallel- subparallel, variedamplitude, moderate^low continuous and wedge- shaped seismic facies suggest a thick intercalation of sandstone with a conglomerate-dominated uvial environment, probably in a braided setting (Cant, 1992; Emery & Myers, 1996).The upward increase in the high GR log response at the base of each cylindrical- or bell- shaped unit coupled with its heterogeneity and vertical spacing in seismic lines may indicate conglomerate lithofacies.We interpret the intercalated-thin- serrated GR and DTresponse as alternating sand- and mudstone-bodies of overbank deposits (Cant, 1992; Emery & Myers, 1996). The overall decrease in GR and increase in DT curves and the upsection increase in thickness indicate a thickening- and coarsening-upward trend. The seismic facies attributes and wireline log characteristics of package N4 described above are analogous to the sandstone- conglomerate-dominated Guandacay Formation. The top of N4 marks a local angu-

160

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin

Fig. 14. W-E wire-line logs at about 191S illustrating aspects of sequence boundaries.The depositional sequences identied can be correlated to near outcrops. See Fig. 2 for location.

lar unconformity.The westward thickening, wedge geometry of the package suggests that the depocentre is located to the west. The pinch out and onlap of N4 at the Alto de Izozog High indicate that of this basement palaeohigh existed before deposition of N4. The onlap and thinning of the N4 reectors on the La Vertiente Fault are attributed to syndepositional deformation. The pronounced top reector implies a strong acoustic impedance contrast and likely represents the erosional surface or an angular regio nal unconformity (Fig. 12, inset; Dunn et al., 1995; Moretti

et al., 1996; Horton & DeCelles, 1997; Echavarria et al., 2003).

Package N5
The base of N5 is dened by a pronounced thick high-amplitude, continuous, high-frequency reector, with onlaps on the underlying N4 package. The top of the N5 package is not well dened and consists of medium- to variableamplitude and moderately continuous, high frequency

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

161

C. E. Uba et al.
reectors. However, well-log data are not available for the top of N5. This package shows a maximum thickness of 2000 m (1.0 ms) in the west and thins eastward to 0^200 m ( 0.1ms) at the Alto de Izozog high, where it occasionally shows a progressive onlap on N4. The internal seismic facies includes laterally extensive, sub- to parallel reectors of variable amplitude, low continuity and low frequency (Figs 8d, 9 and 12). The external forms present are sheet and wedge geometries.The seismic sections show a large vertical spacing of reectors (Figs 8d and 13). Growth structures above the tips of major fault-propagation folds display characteristic fan- shaped reectors with upward-decreasing dip. Only one of these faults (Mandeyepecua Fault) can be mapped throughout the study area (Figs10 and13). No wire-line data are available to ascertain the N5 lithological and sedimentological characteristics. Notwithstanding the absence of wire-line logs, we interpret the N5 package to consist of an alternation of conglomerate and sandstone deposited in a conglomeratedominated uvial setting (e.g. Cant, 1992), probably a large alluvial fan. This interpretation is supported by the subparallel-to -parallel, variable-amplitude, and low continuous, sheet- to wedge- shaped, and high vertical spacing of reectors, which probably suggest thick conglomerate lithofacies. The overall thickening- and coarsening-upward trend, wedge- shaped external form, and pronounced upsection increase in vertical spacing in seismic section are interpreted as a continuing basinward shift in grain size.The seismic facies expressions and alluvial setting interpretation for the N5 package permit its correlation with the surface Emborozu Formation. ing a structural high. This ridge is anked by up to 250 m of strata cratonward and 4100 m of strata orogenward, respectively. Further east, the Petaca thins to 0 at the Alto de Izozog structural high. This nding updates a previous view expressing a lack of thickness variations for this formation (e.g. Gubbels et al., 1993; Moretti et al., 1996). The very low available accommodation space during this period was probably a result of a long time interval of basin stability and non- to low subsidence in this distal part of the basin, augmented by scarce sediment supply inferred from successions of palaeosols (Gubbels etal.,1993; Horton etal., 2001). Gubbels et al. (1993) and Moretti et al. (1996) reported a similar lack of thickness variation for the Yecua Formation. In contrast, Fig. 15b shows a distinctive westward thickening, reaching a maximum thickness of 600 m in outcrop (e.g. Emborozu section).This westward thickening and increase in sandstone proportion of the Yecua strata is also documented in seismic sections (Figs 6, 7, 10 and 13). During 14^7 Ma, exural foreland basin subsidence as a result of thrusting episode that was probably centred in the present-day Interandean- or Subandean Zone (Coudert et al., 1995; Moretti et al., 1996; Echavarria et al., 2003) led to creation of accommodation space in the western part of the basin. Flemings & Jordan (1989) and Sinclair et al. (1991) show that thrusting event results in an increase of the ratio between tectonic subsidence and sediment ux. The dominantly ne-grained sediments in the more distal and relatively sandy facies in the west suggest that the depositional slope was probably too low to produce large coarse-grained deposits or there was a long lag-time between erosion and more coarse-grained sedimentation as documented in other foreland basins (e.g. Blair & Bilo deau, 1988; Jones et al., 2004). The rst appearance of oro genward increases in thickness and sandstone proportion indicates a change in locus of deposition. The three subsequent isopach maps (7^6, 6^2.1and 2.1^ 0; Fig. 15c^ e) show a similar and regular westward-thickening trend, with a maximum thickness of 3800, 2000 and 1500 m, respectively, near the western limit of our study area in the Subandean Ranges. The thick Tariquia strata exhibit a relatively high sediment accumulation rate of 1mm year 1 corresponding during this time in the Subandean Zone (Coudert et al., 1995; Echavarria et al., 2003). The high accommodation space and sedimentation rate could be linked to the elastic exural model from Flemings & Jordan (1989). According to their model, ero sion of uplifted area and subsequent transportation and deposition of sediment in foreland basin decrease the oro genic load and increase the sediment load within the basin, thereby resulting in an increased basin wavelength and an increase in sediment supply, causing migration of basin-magin facies.We proposed that the high sedimentation rate was not only as a result of high topography, as well as a shift in climate from a semi-arid to a humid condition (Uba et al., 2005), which led to high precipitation and high denudation, thus high sediment supply.The high denudation in the west and high sedimentation in the central part

DISCUSSION
Overall stratigraphic pattern
In constructing the isopach maps (Fig. 15), we compiled thickness information for each formation derived from outcrop, depth- converted seismic, wire-line logs and well report data. The thickness values are not corrected for the eect of compaction due to limited postdepositional burial.We constructed isopach maps for ve time periods (occasionally poorly) constrained by age estimates for the formations (Marshall & Sempere, 1991; Marshall et al., 1993; Moretti et al., 1996; Echavarria et al., 2003; Hulka, 2005; Hulka et al., in press). We take thickness variations (Fig.15) through time as a proxy for available accommodation space to infer that the Chaco foreland basin experienced variations in creation of accommodation space since the late Oligocene (e.g. Wadworth et al., 2003). The early basin history is illustrated in a single map (Fig. 15a) spanning more than 13 Ma. A second map (Fig. 15b) spans a 7 Ma- time period. In contrast, the nal three time intervals represent only 3^1 Ma each (Fig. 15c^ e). As expected, these ve maps show distinctive thickness patterns. The isopach map of the 27^14 Ma- old Petaca Formation (Fig. 15a) shows a regional and broad area along the Villamontes-Camiri axis with o50 m of strata, possibly reect-

162

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin

Fig. 15. Isopach maps of the ve late Cenozoic units of the Chaco foreland basin in the study area based on measured surface sections, interpreted industry seismic data, and well logs. (a) 27^14 Ma Petaca Fm. (b) 14^7 Ma Yecua Fm. (c) 7^6 Ma Tariquia Fm. (d) 6^2.1 Ma Guandacay FM. (e) 2.1 Ma Emborozu Fm.

of the basin, coupled with a humid climate, might have also resulted in the migration of the proximal Guandacay conglomerate facies into the region as rapid unloading

outpaced loading (e.g. Blair & Bilodeau, 1988; Catuneanu, 2004). The Tariquia and Guandacay Formations both clearly exhibit a regional asymmetrical geometry and thin

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

163

C. E. Uba et al.
once more, the dominance of ne-grained rocks also agrees with models of facies patterns on the distal margins of underlled foreland basin models (e.g. Blair & Bilodeau, 1988; Sinclair, 1997). During the deposition of the Tariquia, Guandacay and Emborozu Formations, the predominantly uvial deposits, coarsening-upward trend, increase in single-interconnected- channel geometry and high avulsion frequency indicate that the Chaco foreland basin shifted to an overlled stage. Furthermore, this stage is expressed by a predominance of a transverse sediment supply from the mountain belt, and a gradual decrease in accommodation space (e.g. Sinclair & Allen, 1992; Jordan, 1995; Catuneanu, 2004). The transition from an underlled to an overlled stage in a foreland basin system is controlled by a decrease in the rate of exural subsidence, a decrease in sediment bypass and an increase in exhumation (Flemings & Jordan, 1989; Sinclair & Allen, 1992; Catuneanu, 2004).

Alto de Izozog
The Alto de Izozog is a large, topographical high between 550 and 800 m elevation and a width of ca. 300 km (Horton & DeCelles, 1997). It forms a NNE-SSW-trending structural high bordering the eastern limit of the study area. Its uplift mechanism and timing is debated. The Alto de Izo zog has been interpreted as a recent forebulge depocentre (e.g. Coudert et al., 1995; Moretti et al., 1996; Horton & DeCelles, 1997; DeCelles & Horton, 2003). However, the interpreted seismic lines show that all late Cenozoic, Mesozoic and even post-Carboniferous strata onlap and pinch out on this structure (Fig. 10), thereby suggesting a pre-Mesozoic origin. Husson & Moretti (2002) reported a general geothermal gradient of up to 50 1C km 1 and a heat ow of more than 100 mWm 2 at the Alto de Izozog.These values are extremely high compared with the gradient and heat ow values of 26 1C km 1 and 52 mWm 2 from wells further to the west (Husson & Moretti, 2002). Husson & Moretti (2002) also pointed out that these high heat values are abnormal for a forebulge depocentre. In addition, the distance from the Alto de Izozog to the deformation front is rather short. At its minimum, only ca. 70 km separate the exposed basement rocks from the to pographic front of the Subandean ranges, implying that the combined width of the wedge-top and foredeep depo centre reaches barely 100 km (Figs 11 and 15).Theoretically, this short distance is possible, but will imply a very low elastic thickness and require a thicker basin sedimentary ll (e.g. Watts, 2001). In contrast, a high elastic thickness of 460 km (Stewart & Watts, 1997; Tassara, 2005) is observed in the southern Central Andes. The low crossstrike width between assumed forebulge location and deformation front strongly disagrees with values of ca. 4300 km for most other foreland basin systems (e.g. DeCelles & Giles, 1996; DeCelles & Horton, 2003). We found no evidence of forebulge migration since the late Miocene from our interpretation of the seismic data,

Fig. 15. Continued

to 0 at the Alto de Izozog basement high. In contrast, the seismic facies of the Emborozu Formation shows a combination of symmetrical and asymmetrical geometries.

Underfilled and overfilled stages of the Chaco foreland basin


The sedimentary and seismic interpretations, as summarized in Fig. 15, show an overall asymmetrically westwardthickening wedge, a decrease in depositional energy with distance from the deformational front, variable sediment supply and axial to transverse sediment dispersal. Deposits of the Petaca Formation, displaying sedimentation on a very low topographic gradient (Fig. 15a), minor erosion, interbasinal-reworked pedogenic conglomerate, terrestrial condition and a low sediment thickness (250^ 0 m), suggest an overlled stage of the embryonic and still extremely distal Chaco foreland basin because of its consistent easterly transverse sediment dispersal and sedimentary style (e.g. Flemings & Jordan, 1989; Jordan, 1995). During the deposition of the Yecua Formation, basin drainage changed from a transverse to an axial pattern (Fig. 4). This, together with the deposition of mudstonedominated lacustrine and marginal marine facies, indicate an underlled stage (e.g. Flemings & Jordan, 1989), resembling the underlled phase of the Western Taiwan and Camargo Basins, Bolivia (Covey, 1986; DeCelles & Horton, 2003). Because the sediment-accumulation rate decreased (600 m in 7 Ma) and the uvial pattern was modied

164

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin

Fig. 16. Structural cross- sections (modied after Dunn et al., 1995; Moretti et al., 1996; Baby et al., 1997; Kley et al., 1999) of the evolutionary model illustrating the eastward migration of the deformation front and foreland basin depocentres in time and space with evolution of the Andean fold^thrust belt. EC, Eastern Cordillera; IA, Interandes; SZ, Subandean Zone; PF, Pajonal Fault; PBF, Palos Blanco Fault.

although Coudert et al. (1995) estimated 90 km of forebulge migration, based on limited seismic data. A back and forth jump in forebulge location (Waschbusch & Royden, 1992), as observed in the late Devonian/early Mississippian Antler orogeny of the western United States (Giles & Dickinson, 1995), would require such a role of the Alto de Izozog since Mesozoic time, as observed from the onlap and pinch out relationships clearly visible on seismic lines. However, no equivalent pre-Mesozoic foreland basin system is known below the Chaco foreland basin.We therefore consider the Alto de Izozog an unlikely candidate for a recent forebulge but rather advocate a yet-to -be-dened, pre-Mesozoic continental-interior uplift mechanism.

Depocentre migration through time


The late Cenozoic strata express the foreland basin geo metry and sedimentation pattern in four depocentres (backbulge, forebulge, foredeep and wedge-top; DeCelles & Giles, 1996). We interpret the Petaca Formation as an Oligo -Miocene backbulge depocentre east of its Villamontes-Camiri structural high and the axis itself, with only ca. 50 m thickness of the Petaca Formation, as the forebulge (Fig. 16a). The forebulge was likely very low in topo graphic relief and was therefore subjected only to minor erosion. Its preservation is probably a result of forebulge exural migration through the study area between 20

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

165

C. E. Uba et al.

Fig. 17. Diagram illustrating (a) timing and total shortening rate compiled from Echavarria et al. (2003) and Oncken et al. (in press); and the possible major thrusting episode. (b) The rate and propagation of exhumation front of the Central Andes (after Ege, 2004).

and 10 Ma (e.g. Crampton & Allen, 1995; White et al., 2002), similar to the forebulge migration of the Camargo basin (Horton & DeCelles, 1997; DeCelles & Horton, 2003).The presence of a forebulge-backbulge depocentre during Petaca time ( 27^14 Ma) is also supported by the westwarddirected palaeocurrent directions. This interpretation supports the previously predicted backbulge in this region by Mcquarrie et al. (2005). The exural migration of the forebulge into the backbulge area led to minor uplift and moderate erosion. This may explain the presence of an erosional unconformity (see Moretti et al., 1996; Echavarria et al., 2003). The marginal marine, lacustrine and uvial facies of the Yecua Formation indicate a low ratio between sediment supply and accommodation space, and implies the pro gressive migration of the foredeep into the study area by 14^7 Ma. During the deposition of this formation, Coudert et al. (1995) and Echavarria et al. (2003) documented a subsidence rate of 1m Ma 1. Exhumation and structural data indicate that the Subandean Zone began to be deformed and exhumed in this interval (Kley et al., 1996; Moretti et al., 1996; Echavarria et al., 2003; Ege, 2004). The underlled stage of the basin is due to the several-millionyears time lag between loading to the west of the widening basin and its subsequent inll by prograding sediment wedges (e.g. Blair & Bilodeau, 1988). Our interpretation of the Yecua Formation as the ll of a distal foredeep contrasts with its interpretation as a backbulge depocentre by Marshall et al. (1993), but agrees with the interpretation of DeCelles & Horton (2003). The distal foredeep development in the study area may be time- correlative to the wedge-top depocentre in the Camargo basin (DeCelles & Horton, 2003). The thickening- and coarsening-upward Tariquia Formation represents a medial-foredeep depocentre ll (Fig. 16c). This interpretation is supported by the long-lasting and substantial creation of accommodation space and a

high accumulation rate (e.g. Echavarria et al., 2003), a resulting westward increase in large- scale sandstone-dominated facies, and its variable uvial pattern. Palaeocurrents clearly indicate for the rst time a signicant Andean pro venance (Uba et al., 2005). During Guandacay time (6^2.1 Ma), the proximal foredeep depocentre had arrived in the study area (Fig. 16d). Westward thickening, consistent eastward-directed palaeocurrents, and a westward increase in the proportion of conglomerates provide further evidence for the presence of the proximal foredeep. Strikingly similar facies and geometries of foredeep depocentres have been documented by Flemings & Jordan (1989), Sinclair & Allen (1992), DeCelles & Horton (2003) and Catuneanu (2004) for other foreland basins worldwide. We interpret the Emborozu Formation as representing the wedge-top depocentre (Fig. 16e). These thickeningand coarsening-upward strata are apparently regionally restricted, related to specic thrusts, and show wedgeshaped, high-amplitude reectors and growth structures above active blind thrusts (e.g. Fig. 12). The contact between the Guandacay and the Emborozu formations is a progressive regional angular unconformity that marks the transfer from foredeep to wedge-top depocentre.

Regional tectonic implications


The propagation of a foreland basin system depocentres is related to the migration of the orogenic load and to the lithospheric exural response to crustal load and erosional unloading (Jordan et al., 1988; Sinclair & Allen, 1992; DeCelles & Giles, 1996; Pner et al., 2002; DeCelles & Horton 2003; Catuneanu, 2004). Figure 17a shows a compilation of total- shortening and shortening rates in the Subandean (Echavarria et al., 2003; Elger et al., 2005; Oncken et al., in press), whereas Fig. 17b displays the timing of deformation and the propagation of the exhumation front based on apatite ssion track analysis (Ege, 2004) between

166

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin


Tupiza (Eastern Cordillera) and Villamontes (western Chaco plain) corresponding to eastward propagation of the deformation front since the late Oligocene.The gure shows that the shortening rate increased markedly around 27 (?), 10 and 2.1 Ma. The basal palaeosols of the Petaca Formation indicate long periods of low sediment accumulation and suggest little to no structural activity (e.g. Gubbels et al., 1993). However, the lack of age constraints makes it dicult to relate it to Andean tectonics. We speculate that the succession of palaeosols may be older than the Andean orogeny (Cretaceous? or Eocene?). The subsequent reworking of the palaeosols and the sand-mudstone deposition may represent the rst inuence of distant Andean tectonics.We relate this tectonic episode to a major thrusting event that is represented by high shortening and exhumation rates in Fig.17 (Oncken etal., in press; Ege, 2004). It is expressed by the onset of thrusting to the west in the Tupiza region in rail et al., 1996; Kley et al., 1997) the Eastern Cordillera (He that produced 55 km of shortening and a low crustal load (Gregory-Wodzicki, 2000). Consequently, this shortening and low crustal load produced low topography that resulted in a small-magnitude exural wavelength, which probably caused the foreland basin system to migrate into the study area.This situation supports a correlation of the late Oligocene-late Miocene forebulge/backbulge development to the Cayara and Camargo foredeep depocentre farther to the west, as proposed by DeCelles & Horton (2003) and Mcquarrie et al. (2005). According to Horton (1998), Mcquarrie (2002) and Mller et al. (2002), the sedimentary basins of the Tupiza region are associated with fold^thrust deformation, whereas apatite ssion track ages from the Eastern Cordillera show a decrease in cooling ages from 38 to 17 Ma (Ege, 2004). During the deposition of the Petaca Formation, the structural, sedimentolo gical and thermochronologic data indicate major structural growth and crustal thickening within the Eastern Cordillera. Figure 17 shows a pronounced increase in shortening rate, coupled with an increase in exhumation rate during the late Miocene, which has been attributed to low-anglebasement thrusting and the arrival of the deformation front in the westernmost part of the Subandean Zone (e.g. Gubbels et al., 1993; Coudert et al., 1995; Moretti et al. 1996; Kley et al., 1997; Echavarria et al., 2003; Ege, 2004). Yecua strata record basement-imbricated uplift in the Interandean and eastern propagation of the fold^thrust system in the eastern Interandean or Subandean Zone. Coudert et al. (1995) and Echavarria et al. (2003) suggest that the sedimentation rate increased rapidly during this time. However, Echavarria et al. (2003) suggest that during this time, the rate of uplift in the Subandean Zone must have been less than the rate of sedimentation. We attribute the deposition of the Tariquia and Guandacay Formations to exural response due to the high sedimentation as a result of high erosion of shortened and uplifted regions in the Interandean and Subandean ranges during tectonic quiescence, notwithstanding the occurrence of minor thrusting (e.g. Echavarria et al., 2003) in the basin, such as the development of the local, 6^2.1 Maold La Vertiente structure (Moretti et al., 1996; Fig. 12). It suggests that the 14^7 Ma major thrust episode might have produced a very large crustal load and therefore, a large wavelength shortly before the onset of deposition of the Tariquia Formation, which resulted in high denudation in the western part of the Subandean Zone and increased accommodation space and deposition of more than 3500m-thick- sediments in 7^6 Ma in the central part of the Subandean Zone. During this time, the young structures were just beginning to grow in the study area (e.g. Moretti et al., 1996; Kley et al., 1999; Ege, 2004). The Emborozu Formation marks both the reactivation of thrusting in the west (Emborozu section) and the arrival of the deformation front at the western Chaco plain (Aguarague range). However, the timing of this thrusting (Gubbels et al., 1993; Moretti et al., 1996; Echavarria et al., 2003; Ege, 2004) remains debated. Moretti et al. (1996) and Gubbels et al. (1993) estimate the age of the thrusting to postdate the formation and uplift of the leading large anticline (Aguaragua range; Fig. 2) at 3.3 Ma age of mica on tu, whereas Echavarria et al. (2003) postulated a younger age of approximately 2.5 Ma for the in- sequence thrust, with out- of- sequence reactivation of older structures in the west at 2^2.2 Ma. Our study agrees with the results of Echavarria et al. (2003) that 2.1 Ma (herein constrained) Emborozu strata to the east reect in- sequence fold^ thrust propagation into the basin, which led to Aguarague range uplift, although the equivalent strata to the west represent the reactivation of the older Nogalitos range (Fig. 2), thus forming an out- of sequence intermontane basin (e.g. Echavarria et al., 2003).

CONCLUSIONS
The combination of seismic stratigraphy and outcrop facies interpretation argues for a close interaction between Andean fold^thrust belt deformation and Chaco foreland basin development since the late Oligocene, resulting in a group of eastward-migrating foreland system depocentres, driven primarily by crustal shortening and tectonic loading. Signicant exural subsidence developed since the late Miocene. Variably created accommodation space was predominantly lled by non-marine siliciclastics. The sedimentary ll of the Chaco foreland basin can be assigned to dierent depocentres starting with the Petaca Formation in the backbulge and forebulge, through distal-, medial- and proximal foredeep by Yecua, Tariquia and Guandacay Formations, respectively, and nally to wedgetop deposition of the Emborozu Formation. Three major tectonic episodes are expressed in the foreland strata: (1) the late Oligocene uplift of the Eastern Cordillera initiated the foreland development and is expressed as relatively steep to low-angle basement thrusts; (2) late Miocene formation of the Intrandean/Subandean foldand-thrust belt led to a pronounced underlled stage;

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

167

C. E. Uba et al.
and (3) late Pliocene shortening generated the overlled, coarse clastic wedges of the Emborozu Formation. Foreland development and depocentre migration agree well with fold-and-thrust belt exhumation rates. The late Cenozoic Chaco basin is a classical example of a foreland basin system.The overall coarsening- and thickening-upward trend and stratigraphic architectures document a propagating fold^thrust belt and corresponding foreland basin depocentres (DeCelles & Giles, 1996). Similar migration of depocentre with time and foreland basin architecture has been recorded for numerous other basins worldwide, such as the Karoo foreland basins (Catuneanu et al., 1999), Taiwan (Covey, 1986) and North Alpine (Schlunegger et al., 1997; Pner et al., 2002).
Badley, M.E. (1985) Practical Seismic Interpretation. Intern. Human Res. Develop. Corp., Boston. Beaumont, C. (1981) Foreland basins. Geophys. J. Roy. Astron. Soc., 65, 291^329. Belotti, H.J., Saccavino, L.L. & Schachner, G.A. (1995) Structural styles and petroleum occurrence in the Subandean fold and thrust belt of northern Argentina. In: Petroleum Basins rez Soruco & H.J. of South America (Ed. by A.J. Tankard, R. Sua Welsink), Am. Assoc. Pet. Geol. Mem., 62, 545^555. Blair, T.C. & Bilodeau, W.L. (1988) Development of tecto nic cyclotherms in rift, pull-apart, and foreland basins: sedimentary response to episodic tectonism. Geology, 16, 517^520. Bridge, J.S. (2003) Rivers and Floodplains: Forms, Processes, and Sedimentary Records. Blackwell, Oxford, 491pp. ' ngano, M.G, Hulka, C. & HeuBuatois , L.A., Uba, C.E., Ma beck, C. (in press). Deep bioturbation in continental environments: evidence from Miocene uvial deposits of Bolivia. In: Ichnology at the Crossroads: A Multidimensional Approach to the Science of Organism^Substrate Interactions (Ed. by R. Bromley, L.A. Buatois, J.J. Genise, M.G. Ma ngano & R. Melchor), SEPM Special Publ.. Burbank, D.W. & Raynolds , R.G.H. (1988) Stratigraphic keys to the timing of thrusting in terrestrial foreland basins; applications to the northwestern Himalaya. In: New Perspectives in Basin Analysis (Ed. by K.L. Kleinspehn & C. Paola) SpringerVerlag, New York. Cant, D.J. (1992) Subsurface facies analysis. In: Facies Models: Response to Sea Level Changes (Ed. by R.G. Walker & N.P. James), Geol. Assoc., 27^45. Catuneanu, O. (2004) Retroarc foreland systems ^ evolution through time. J. Afr. Earth Sci., 38, 225^242. Catuneanu, O., Beaumont, C. & Waschbusch, P. (1997) Interplay of static loads and subduction dynamics in foreland basins: reciprocal stratigraphies and missing peripheral bulge. Geology, 25, 1087^1090. Catuneanu, O., Sweet, A.R. & Miall, A.D. (1999) Concept and styles of reciprocal stratigraphies: western Canada foreland basin. Terra Nova, 11, 1^8. Cecil, C.B. (1990) Paleoclimate controls on stratigraphic repetition of chemical and siliciclastic rocks. Geology, 18, 533^536. Coudert, L., Frappa, M., Viguier, C. & Arias , R. (1995) Tectonic subsidence and crustal exure in the Neogene Chaco Basin of Bolivia. Tectonophysics, 243, 277^292. Covey, M. (1986) The evolution of foreland basin to steady state: evidence from the western Taiwan foreland basin. In: Foreland Basins (Ed. by P. Allen & P. Homewood), Spec. Publ. Int. Assoc. Sediment., 8, 77^90. Crampton, S.L. & Allen, P.A. (1995) Recognition of forebulge unconformities associated with early stage foreland basin development: example from north Alpine foreland basin. Am. Assoc. Pet. Geol. Bull., 79, 1495^1514. Decelles , P.G. & Giles , K.A. (1996) Foreland basin systems. Basin Res., 8, 105^123. Decelles , P.G. & Horton, B.K. (2003) Early to MiddleTertiary foreland basin development and the history of Andean crustal shortening in Bolivia. Geol. Soc. Am. Bull., 115, 58^77. Dunn, J.F., Hartshorn, K.G. & Hartshorn, P.W. (1995) Structural styles and hydrocarbon potential of the Subandean thrust belt of Southern Bolivia. In: Petroleum Basins of South America (Ed. by A.J.Tankard, R. Suarez Soruco & H.J.Welsink), Am. Assoc. Pet. Geol. Mem., 62, 523^543.

ACKNOWLEDGEMENTS
This paper is part of a PhD thesis by the rst author at the Freie Universitt Berlin, Germany. The authors were supported nancially by the DFG through the Sonderforschungsbereich (SFB) 267 and logistically by Chaco S.A., Santa Cruz, Bolivia.We are indebted to Oscar Aranibar, Fernando Alegria and Nigel Robinson of Chaco S.A. for their assistance.Thanks are also due to David T uno Ba nzer of Yacimientos Petroleros Fiscales de Bolivia (YPFB), Santa Cruz, Bolivia, for providing some of the seismic lines. We also thank Harald Ege and AndresTassara (Freie Universitt Berlin) for contributing shortening and AFT data and for helpful and stimulating discussions on Andean geodynamics. We are grateful for comments provided by the reviewers Jonas Kley, Fritz Schlunegger and Patrice Baby, which greatly improved the early revisions of this manuscript.

REFERENCES
Ayaviri, A. (1964) Geolog| a del a rea deTarija, entre los r| os Pilaya ^ Pilcomayo y R| o Bermejo. Informe InternoYPFB (GXG-996). Ayaviri, A. (1967) Estratigraf| a del Subandino meridional. Informe InternoYPFB (GXG-1215). Alves , T.M., Manuppella, G., Gawthorpe, R.L., Hunt, D.W. & Montiero, J.H. (2003) The depositional evolution of diaper- and fault-bounded rift basins: examples from the Lusitanian Basin of West Iberia. Sediment. Geol., 162, 273^303. Baby, P., He rail, G., Salinas , R. & Sempere, T. (1992) Geometry and kinematic evolution of passive roof duplexes deduced from cross- section balancing: example from the foreland thrust system of the southern Bolivian Subandean Zone. Tectonics , 11, 523^536. Baby, P., Moretti, I., Guillier, B., Limachi, E., Mendez, E., Oller, J. & Specht, M. (1995) Petroleum system of the northern and central Bolivian Subandean Zone. In: Petroleum Basins rez Soruco & H.J. of South America (Ed. by A.J. Tankard, R. Sua Welsink), Am. Assoc. Pet. Geol. Mem., 62, 445^458. Baby, P., Rochat, P., Mascle, G. & He rail, G. (1997) Neogene shortening contribution to crustal thickening in the back-arc of the Central Andes. Geology, 25, 883^886.

168

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Evolution of the late Cenozoic Chaco foreland basin


ndez, R., Allmendinger, R. & ReyEchavarria, L., Herna nolds , J. (2003) Subandean thrust and fold belt of northwestern Argentina: geometry and timing of the Andean evolution. Am. Assoc. Pet. Geol. Bull., 87, 965^985. Ege, H. (2004) Exhumations- und Hebungsgeschichte der zentralen Anden in Sdbolivien (211S) durch Spaltspur-Thermo chronologie an Apatit. PhD Thesis, Freie Universitt, Berlin. Elger, K., Oncken, O. & Glodny, J. (2005) Plateau- styles accumulation of deformation: the southern Altiplano. Tectonics, 24, TC4020, doi:10.1029/2004TC001675. Flemings , P.B. & Jordan, T.E. (1989) A synthetic stratigraphic model of foreland basin development. J. Geophys. Res., 94, 3851^3866. Emery, D. & Myers , K.J. (1996) Sequence Stratigraphy. Blackwell Publishers, Oxford, 297pp. Giles , K.A. & Dickinson, W.R. (1995) The interplay of eustasy and lithospheric exure in forming stratigraphic sequences in foreland settings: an example from the Antler foreland, Nevada and Utah. In: Stratigraphic Evolution of Foreland Basins (Ed. by G.M. Ross), Spec. Publ. SEPM, 52, 187^211. Gregory-Wodzicki, K.M. (2000) Uplift history of the central and northern Andes: a review. Geol. Soc. Am. Bull., 112,1091^1105. Gubbels , T.L., Isacks , B.L. & Farrar, E. (1993) High-level surface, plateau uplift, and foreland development, Bolivian central Andes. Geology, 21, 695^698. He rail, G., Oller, J., Baby, P., Bonhomme, M.G. & Soler, P. (1996) Strike- slip faulting, thrusting and related basins in Cenozoic evolution of the southern branch of the Bolivian oro cline. Tectonophysics , 259, 201^212. Hofmann, M.H., Hendrx, M.S., Moore, J.N. & Sperazza, M. (2006) Late Pleistocene and Holocene depositional history of sediments in Flathead Lake, Montana: evidence from high-resolution seismic reection interpretation. Sediment. Geol., 184, 111^131. Horton, B.K. (1998) Sediment accumulation on top of the Andean orogenic wedge: Oligocene to late Miocene basins of the Eastern Cordillera, southern Bolivia. Geol. Soc. Am. Bull., 110, 1174^1192. Horton, B.K. & Decelles, P.G. (1997) The modern foreland basin system adjacent to the Central Andes. Geology, 25, 895^898. Horton, B.K. & Decelles , P.G. (2001) Modern and ancient uvial megafans in the foreland basin system of the central Andes, southern Bolivia: implications for drainage network evo lution in fold^thrust belts. Basin Res., 13, 43^63. n, Horton, B.K., Hampton, B.A., Lareau, B.N. & Baldello E. (2001) Tertiary provenance history of the northern and central Altiplano (Central Andes, Bolivia): a detrital record of plateau-margin tectonics. J. Sediment. Res., 72, 711^726. Hulka, C. (2005) Sedimentary and tectonic evolution of the Cenozoic Chaco foreland basin, Southern Bolivia. PhD Thesis, Freie Universitt Berlin. Hulka, C., Grfe, K.U., Sames , B., Heubeck, C. & Uba, C.E. (in press). Depositional setting of the middle to late Miocene Yecua Formation of the central Chaco foreland basin, Bolivia. J. South Am. Earth Sci. Husson, L. & Moretti, I. (2002) Thermal regime of fold and thrust belts ^ an application to the Bolivian sub Andean zone. Tectonophysics , 345, 253^280. Isacks , B.L. (1988) Uplift of the central Andean plateau and bending of the Bolivian orocline. J. Geophys. Res., 93, 3211^3231. Jimenez-Miranda, F. & Lopez-Murillo, R. (1971) Estratigraf| a de algunas secciones del subandino sur y centro. Informe Interno YPFB. Jones , M.A., Heller, P.L., Roca, E., Garce s , M. & Cabrera, L. (2004) Time lag of syntectonic sedimentation across an alluvial basin: theory and example from Ebro Basin, Spain. Basin Res., 16, 467^488. Jordan, T.E. (1981) Thrust loads and foreland basin evolution, Cretaceous, western United States. Am. Assoc. Pet. Geol. Bull., 65, 2506^2520. Jordan, T.E. (1995) Retroarc foreland and related basins. In: Tectonics of Sedimentary Basins (Ed. by C.J. Busby & R.V. Ingersoll), pp. 331^362. Blackwell Scientic, Oxford. Jordan, T.E., Flemings , P.B. & Beer, J.A. (1988) Dating of thrust-fault activity by use of foreland basin strata. In: New perspectives in basin analysis (Ed. by K. Kleinspehn & C. Paola), Springer^Verlag, New York, 307^330. Jordan, T.E., Reynolds , J.H. & Erikson, J.P. (1997) Variability in age of initial shortening and uplift in the central Andes, 16^ 331300 S. In: Tectonic Uplift and Climate Change (Ed. by W.F. Ruddiman), pp. 41^61. Plenum Press, New York. Kennen, L., Lamb, S.H. & Hoke, L. (1997) High-altitude paleosurfaces in the Bolivian Andes: evidence for late Cenozoic surface uplift. In: Palaeosurfaces: Recognition, Reconstruction, and Palaeoenvironmental Interpretation (Ed. by M. Widdowson), Spec. Publ. Geol. Soc. Am., 120, 307^323. Kley, J. (1996) Transition from basement-involved to thinskinned thrusting in the Cordillera Oriental of southern Bolivia. Tectonics, 15, 763^775. Kley, J., Gangui, A.H. & Krger, D. (1996) Basement-involved blind thrusting in the eastern Cordillera Oriental, southern Bolivia: evidence from cross- sectional balancing, gravimetric and magnetotelluric data. Tectonophysics, 259, 171^184. Kley, J., Monaldi, C.R. & Salfity, J.A. (1999) Along- strike segmentation of the Andean foreland: causes and consequences. Tectonophysics , 301, 75^94. Kley, J., Mller, J., Tawackoli, S., Jacobshagen, V. & Manutsoglu, E. (1997) Pre-Andean and Andean-Age Deformation in the Eastern Cordillera of Southern Bolivia. J. South Am. Earth Sci., 10, 1^19. Makaske, B., Smith, D.G. & Berendsen, H.J.A. (2002) Avulsions, channel evolution and oodplain sedimentation rates of the anastomosing upper Columbia River, British Columbia, Canada. Sedimentology, 49, 1049^1071. Marshall, L.G. & Sempere, T. (1991) The Eocene to Pleisto cene vertebrates of Bolivia and their stratigraphic context: a review. In: Fosiles y facies de Bolivia:V ertebrados (Ed. by R. SuarezSoruco), Rev.Te cn. deYPFB, 11, 631^652. Marshall, L.G., Sempere, T. & Gayet, M. (1993) The Petaca (late Oligocene-middle Miocene) and Yecua (late Miocene) formations of the Subandean-Chaco basin, Bolivia, and their tectonic signicance. Doc. Lab. Ge ol. Lyon, 125, 291^301. Mcquarrie, N. (2002) The kinematic history of the central Andean fold^thrust belt, Bolivia: implications for building a high plateau. Geol. Soc. Am. Bull., 114, 950^963. Mcquarrie, N. & Decelles , P.G. (2001) Geometry and structural evolution of the central Andean backthrust belt, Bolivia. Tectonics, 20, 669^692. Mcquarrie, N., Horton, B.K., Zandt, G., Beck, S. & Decelles , P.G. (2005) Lithospheric evolution of the Andean fold^ thrust belt, Bolivia, and the origin of the central Andean plateau. Tectonophysics, 399, 15^37. Miall, A.D. (1996) The Geology of Fluvial Deposits. Springer-Verlag, Berlin. Mitchum Jr, R.M., Vail, P.R. & Sangree, J.B. (1977) Seismic stratigraphy and global changes of sea level, part 6:

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

169

C. E. Uba et al.
stratigraphic interpretation of seismic reection patterns in depositional sequences. In: Seismic Stratigraphy ^ Application to Hydrocarbon Exploration (Ed. by C.E. Payton), Am. Assoc. Pet. Geol. Mem., 26, 516. Moretti, I., Baby, P., Mendez, E. & Zubieta, D. (1996) Hydro carbon generation in relation to thrusting in the Subandean Zone from 181 to 221S, South Bolivia. Pet. Geosci., 2, 17^28. Mller, J.P., Kley, J. & Jacobshagen, V. (2002) Structure and Cenozoic kinematics of the Eastern Cordillera. T ectonics, 21, 1^27. Oller, J. (1986) Consideraciones generales sobre la geologia y estratigraa de la Faja Subandina norte. MSc Thesis Universite Mayor de San Andres, La Paz. Oncken, O., Hindle, D., Kley, J., Elger, K. & Victor, P. (in press). Deformation of the Central Andean upper plate system^facts, ction, and constraints for plateau models. In: Deformation Processes in the Andes (Ed. by O. Oncken, G. Chong, G. Franz, P. Giese, H.J. Gtze,V. Ramos, M. Strecker & P.Wigger), Front. Earth Sci. Monogr. Ser. 1. n al Le xico EsPadula, L.E. & Reyes , F.C. (1958) Contribucio tratigra co de las Sierras Subandinas, Repu blica de Bolivia. Bolet| n Te cn.YPFB, 1, 9^70. Pfiffner, O.A., Schlunegger, F. & Buiter, S. (2002) The Swiss Alps and their peripheral foreland basin: stratigraphic response to the deep crustal processes. Tectonics, 21, 3.1^3.16. Retallack, G.J. (1997) A Color Guide to Paleosols. Wiley, Chichester, 175pp. Roeder, D. & Chamberlain, R.L. (1995) Structural geology of Sub Andean fold and thrust belt in northwestern Bolivia. In: Petroleum Basins of South America (Ed. by A.J.Tankard, S. Suarez & H.J.Welsink), Am. Assoc. Pet. Geol. Mem., 62, 459^479. Russo, A. (1959) Estructura y estratigraf| a del a rea de Agua Salada. Bolet| n Te c.YPFB, 3, 13^35. Sangree, J.B. & Widmier, J.M. (1977) Seismic stratigraphy and global changes of sea level, part: seismic interpretation of clastic depositional facies. In: Seismic Stratigraphy ^ Applications to Hydrocarbon Exploration (Ed. by C.E. Payton), Am. Assoc. Pet. Geol. Mem. 26 (pp. 516. Schlunegger, F., Leu, W. & Matter, A. (1997) Sedimentary sequences, seismic facies, subsidence analysis, and evolution of the Burdigalian upper Marine Grop, central Switzerland. Am. Assoc. Pet. Geol. Bull., 81, 1185^1207. Sempere, T. (2000) Sediment accumulation on top of the Andean orogenic wedge; Oligocene to late Miocene basins of the Eastern Cordillera, southern Bolivia; discussion and reply. Geol. Soc. Am. Bull., 112, 1752^1759. Sempere, T. (1995) Phanerozoic evolution of Bolivia and adjacent regions. In: Petroleum Basins of South America (Ed. by A.J. Tan rez Soruco & H.J. Welsink), Am. Assoc. Pet. Geol. kard, R. Sua Mem., 62, 207^230. Sempere, T., He rail, G., Oller, J. & Bonhomme, M.G. (1990) Late Oligocene^Early Miocene major tectonic crisis and related basins in Bolivia. Geology, 18, 946^949. Sempere, T., Butler, R.F., Richards, D.R., Marshall, L.G., Sharp, W. & Swisher, C.C., III (1997) Stratigraphy and chronology of the upper Cretaceous-lower Paleogene strata in Bolivia and northwest Argentina. Geol. Soc. Am. Bull., 109, 709^727. Sheffels , B. (1988) Structural Constraints on Crustal Shortening in the Bolivian Andes. PhD Thesis, Massachusetts Institute of Technology, Boston. Sheriff, R.E. & Geldart, L.P. (1995) Exploration Seismology. Cambridge University Press, Cambridge. Sinclair, H.D. (1997) Tectonostratigraphic model for underlled peripheral foreland basins: an Alpine perspective. Geol. Soc. Am. Bull., 109, 324^346. Sinclair, H.D. & Allen, P.A. (1992) Vertical vs horizontal mo tions in the Alpine orogenic wedge: stratigraphic response in the foreland basin. Basin Res., 4, 215^232. Sinclair, H.D., Coakley, B.J., Allen, P.A. & Watts , A.B. (1991) Simulation of foreland basin stratigraphy using a diusion model of mountain belt uplift and erosion: an example from the central Alps, Switzerland. Tectonics , 10, 599^620. Smith, D.G. (1986) Anastomosing river deposits, sedimentation rates and basin subsidence, Magdalena River, northwestern Columbia, South America. Sediment. Geol., 46, 177^196. Somoza, R. (1998) Updated Naszca (Farallon)-South America relation motions during the last 40 My: implication for mountain building in the central Andean region. J. South Am. Earth Sci., 11, 211^215. Stewart, J. & Watts , A.B. (1997) Gravity anomalies and spatial variations of exural rigidity at mountain ranges. J. Geophys. Res., 102, 5327^5352. Suarez Soruco, R. (2000) Le xico estratigra co de Bolivia. Rev. Tec.YPFB, 17, 1^213. Tankard, A.J. (1986) On the depositional response to thrusting and lithospheric exure: examples from the Appalachian and Rocky Mountain basins. In: Foreland Basins (Ed. by P. Allen & P. Homewood), Spec. Publ. Int. Assoc. Sediment., 8, 369^392. Tassara, A. (2005) Interaction between the Nazca and South American plates and formation of the Altiplano -Puna plateau: review of a exural analysis along the Andean margin (151^ 341S). Tectonophysics , 399, 39^57. Uba, C.E., Heubeck, C. & Hulka, C. (2005) Facies analysis and basin architecture of the Neogene Subandean synorogenic wedge, southern Bolivia. Sediment. Geol., 180, 91^123. Wadworth, J., Boyd, R., Diessel, C.F.K. & Leckie, D.A. (2003) Stratigraphic styles of coal and non-marine strata in high-accommodation setting: Falher Member and Gates Formation Lower Cretaceous), western Canada. Can. Assoc. Pet. Geol. Bull, 51, 275^303. Waschbusch, P.J. & Royden, L.H. (1992) Spatial and temporal evolution of foredeep basins: lateral strength variations and inelastic yielding in continental lithosphere. Basin Res., 4, 179^196. Watts , A.B. (1992) The eective elastic thickness of the litho sphere and evolution of foreland basin. Basin Res., 4, 169^178. Watts , A.B. (2001) Isostasy and Flexure of the Lithosphere. Cambridge University Press, Cambridge, UK, 458 pp. White, T., Furlong, K. & Arthur, M. (2002) Forebulge migration in the Cretaceous Western Interior basin of the central United States. Basin Res., 14, 43^54. Welsink, H.J., Martinez, E., Aranibar, O. & Jarandilla, J. (1995) Structural inversion of a Cretaceous rift basin, southern Altiplano, Bolivia. In: Petroleum Basins of South America (Ed. by rez-Soruco & H.J. Welsink), Am. Assoc. Pet. A.J. Tankard, R. Sua Geol. Mem., 62, 305^324.

Manuscript received 7 July 2005; Manuscript accepted17 April 2006

170

r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170

Você também pode gostar