evolutia bazinului de foreland chaco

Upload: ultimatereader

Post on 06-Apr-2018

230 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    1/26

    Evolution of the late Cenozoic Chaco foreland basin,Southern BoliviaCornelius Eji Uba1, Christoph Heubeck and Carola Hulka

    Institut fr GeologischeWissenschaften, FreieUniversitt Berlin, Berlin, Germany

    ABSTRACTEastward 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 ofve 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 Andean-

    derived 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 s elective

    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 deformationfront 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 exuralwarp-

    ing of the lithosphere in response to supralithospheric

    and sublithospheric orogenic wedging (DeCelles & Giles,

    1996; Pner etal., 2002). Lithospheric exure under static

    loads generates down-bending exure proximal to the

    orogen, which migrates as the load advances. Forelandbasins 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 thrust-

    ing varies between and within the foreland basin system

    but is mainly controlled by the elastic thickness of the

    lithosphere and the applie d loads (Watts, 1992, 2001). De-

    Celles & Giles (1996) characterized foreland basin systems

    into fourdierent depocentres: wedge-top, foredeep, fore-

    bulge and backbulge. Each depocentre exhibits distinctive

    internal architecture, sedimentology and structure. Ac-

    commodation space is created by combined static anddynamic subsidence (DeCelles & Giles, 1996; Catuneanu

    etal., 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 etal.,1990; De-

    Celles& Horton, 2003) in response to Nazca-South Amer-

    ican plate convergence and its related eastward interaction

    with the Brazilian shield.Detailed structural studies in the

    Interandean and Subandean Zones documented structur-

    al styles and timing of deformation (Sempere et al., 1990;

    Correspondence: Cornelius Eji Uba, Institut fr GeologischeWissenschaften, Freie Universitt Berlin, Malteserstrasse 74 -100, 12249 Berlin, Germa ny. E-mail: [email protected] -po tsdam.de1 Present address: Institut fr Geowissenschaften, UniversittPotsdam, Karl-Liebknecht Str. 24/25,14476 Potsdam/Golm, Ger-

    many.

    BasinResearch (2006) 18, 145170, doi: 10.1111/j.1365-2117.2006.00291.x

    r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd 145

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    2/26

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    3/26

    Oligocene (Oller, 1986; Sheels, 1988; Sempere etal., 1990;

    Baby et al., 1992; He rail et al., 1996). This foreland basinforms 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 year1 to a subse-

    quent maximum rate of up to 15.2 cm year1 (in the late

    Oligocene; Somoza, 1998). Andean deformation com-

    menced in the west with formation of the Altiplano basin

    and foreland sedimentation, with the depocentre probably

    at the present Eastern Cordillera (Sempere etal.,1997; De-

    Celles & Horton, 2003; Elger et al., 2005). During the Cre-

    taceous-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-deposi-tion.

    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 Suban-

    dean 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 Suban-

    dean fold-and-thrust-belt and is still undergoing active

    shortening at its le ading edge (Fig. 2; Oller, 1986; Sheels,

    1988; Baby etal.,1992; He rail etal.,1996). Late Cenozoic se-

    dimentary strata are commonly well exposed along anks

    Fig. 2. Geological and structural map of the study area (modied from Suarez-Soruco, 2000) showing the data s et 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 .

    r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170 147

    Evolutionof thelateCenozoicChaco foreland basin

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    4/26

    of the major leading syn- and anticlines near the wedge-

    tip of the Subandean zone. This region, between $641300

    and 621300E and 181450 and 221300S, forms the principal

    study area.

    Chaco foreland basin sedimentation is assumed to have

    begun approximately 27 Ma ago near the Eastern Cordil-

    lera (Sempere etal., 1990), as a result of eastward migration

    of the deformation front ahead of the Interandean and

    Subandean Zones (Sempere etal.,1990; Husson & Moretti,

    2002; DeCelles & Horton, 2003; Echavarria et al., 2003;

    Ege, 2004). Late Cenozoic strata show characteristic west-

    ward thickening. It stands to reason that Chaco foreland

    basin strata had also been deposited in considerable thick-

    ness in the region occupied by the present-daySubandean

    Zone before its uplift and incorporation into the eastward-

    migrating orogenic wedge. Although these deposits are

    well preserved in the Subandean Zone, in some areas thecoarse-grained proximal foreland basin deposits have

    been eroded.The eroded proximal basin sedimentation is

    as a result of erosion removal after deformation and subse-

    quent propagation of the fold^thrust belt(Burbank & Ray-

    nolds, 1988). A rough estimate of their original thickness

    (ca. 3^5 km) can be obtained by reconstructing thermal-

    gradient-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 c orresponding shortening in

    the fold-and-thrust belt of the Subandean zone, recon-

    structed 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, wide-

    spread 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 generallye astward-younging synorogenic wedge (Moret-

    ti et al., 1996; DeCelles & Horton, 2003; Echavarria et al.,

    2003).

    During the late Oligocene, the Eastern Cordillera was

    the focus of pronounced shortening (Kleyetal.,1997; Hor-

    ton, 1998). Figure 3 shows structural styles and major

    thrust sheets, illustrating that the Subandean Zone is de-

    formed by mostly in-sequence, thin- skinned thrust

    sheets that include north-northeast-trending ramp anti-

    clines and passive roof duplexes (Baby etal.,1992,1997; Be-

    lotti et al., 1995; Dunn et al., 1995; Kley et al., 1996, 1999;

    Echavarria et al., 2003).This progressive thin-skinned de-

    formation is recorded in a suite of angular unc onformities

    and stratigraphically distinct foreland packages. A total

    shortening of 210 336 km is postulated for the Central An-

    des (Baby etal., 1992; Moretti etal.,1996; McQuarrie & De-

    Celles, 2001; Mlleretal., 2002; Elger etal., 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 at211S in the Interandean and Subandean zones

    together (Kleyetal.,1997). Moretti etal. (1996) calculated a

    peak shortening rate between 6 and 2.1Ma, 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 year1)

    at 9^7 and 2^0 Ma, respectively, separated by an in-be-

    tween low of 0^5 mm year1 at 221300 latitude.These con-

    tradictions 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 etal., 1996;

    Echavarria etal., 2003) and the resulting inferred time per-

    iods 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,Echa-

    varria et al. (2003) attribute the 2-Ma-shortening event to

    thrust reactivation in the south-western Subandean Zone

    ( $221300), 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-

    Fig. 3. Structural balanced cross- section of the Central Andes from Altiplano to Chaco plain at 211S. Modied after Kley etal. (1999)

    and Elger et al. (2005). UKFZ, Uyuni-Khenayani Fault Zone ; SVT, San VicenteThrust ; CYT, Cama rga-Yavi Thrust ; IAT, Inte randean

    Thrust ; SAT, Subandean Thrust ; PF, Pajonal Fault; PBF, Palos Blanco Fault. See Fig. 2 for location.

    r 2006 The Authors. Journal compilat ion r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145 170148

    C. E. Ubaet al.

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    5/26

    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 east-

    ward growth of the Andes onto the Brazilian Shield.

    DATA AND METHODS

    The data set compiled for this study includes measuredsections, seismic data and well logs. Twenty- one strati-

    graphic sections along major rivers, small streams and

    road cuts in the Subandean foothills were measured and

    sampled for lithologic, sedimentologic and biostrati-

    graphic 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 ex-

    ploration wells and tied them to42800 km of 2-D indus-

    try s eismic proles. Wire-line logs (g-ray, resistivity and

    sonic), combined with well reports, provide ne vertical

    details of wells and lithology resolution and thus comple-

    ment seismic data for a better understanding of thesubsurface geology. Similar methods were used by Schlu -

    negger etal. (1997)and Alves etal. (2003)to study the Upper

    Marine Molasse Group of the North Alpine foreland basin

    and the Lusitanian rift basin of West Iber ia, respectively.

    Wire-line log analysis and well reports were used in com -

    bination with seismic facies attributes to delineate dier-

    ent 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 Ya-

    cimientos 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. Syn-

    thetic 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), eastward-

    thinning 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; Mar-shall & 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

    southernBolivia.However, the ages should be appliedwith

    caution. Parts of the formations are suspected to be dia-

    chronous, not only younging west-to- east, as could be ex-

    pected, but possibly also north-to- south (Echavarria etal.,

    2003). In addition, the stratigraphy is complicated by sev-

    eral nearly basinwide low-angle unconformities.

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

    r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170 149

    Evolutionof thelateCenozoicChaco foreland basin

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    6/26

    Petaca formation

    Cenozoic sedimentation in the Chaco basin commenced

    during the late Oligocene (assumed ca. 27Ma; 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

    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

    sandston e bed of the Yecua Fm (arrow). (c) Channeli zed sandstone beds with de siccatio n cracks (arrow) of theTariquia Fm. (d)

    Alternation of conglomerate and s andstone beds of the Guandacay Fm. (e) Sheet-like cobble-boulder-dominated conglomerate bed

    with thin beds of sandstone of the Emborozu Fm.

    r 2006 The Authors. Journal compilat ion r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145 170150

    C. E. Ubaet al.

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    7/26

    to clustered, blocky to massive and bracciated nodules

    (Uba etal., 2005).The calcrete body is overlain by horizon-

    tal to disorganized clast-supported reworked-pedogenic

    conglomerate (Fig.5a) composing of poorly sorted, densely

    packed clasts of poorly rounded intraformational re-

    worked calcrete nodules and subordinate chert. The con-

    glomerates show sharp and erosive bases. Medium to

    very-coarse-grained sandstone and sandy to ne-grained

    mudstone mark the upsection lithology of the Petaca For-

    mation.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 re-

    worked 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 etal. (2005) attributed the thick calcrete horizons to

    well-developed palaeosols, indicating 0 or low sedimenta-

    tion in an arid to semiarid climate in which evaporation

    generally exceeded precipitation (e.g. Cecil, 1990). The

    lithofacies and architectural elements in the Petaca con-

    glomerate and sandstone indicate variable high-energy

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

    r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170 151

    Evolutionof thelateCenozoicChaco foreland basin

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    8/26

    stream ows in a channelized setting. Uba etal. (2005) andMarshall 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 un-

    derlying eolian strata is a regional erosional unc onformity

    that may have formed as a far- eld response to early An-

    dean tectonics (Sempereetal.,1990;Dunn etal.,1995). Mar-

    shall et al. (1993) reported reptilian and mammal bone

    fragments of late Oligocene to late Miocene age, foundclose 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 low-

    angle erosional unconformity. The Yecua Formation shows

    a west-to-east and northeast-to-southwest facies varia-

    tion. North of Camiri, it consists of red-green to brown

    sandstone^mudstone couplets (Fig. 5b) showing herring-

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

    r 2006 The Authors. Journal compilat ion r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145 170152

    C. E. Ubaet al.

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    9/26

    bone cross-stratication, laminated, convolute, aser,

    wavy and lenticular bedding in ning- and coarsening-

    upward successions.This lithofacies also consists of gyp-

    sum veins, syndepositional structures, bioturbation and

    desiccation marks (see Fig. 5b). Fossils include bivalves,

    the foraminifera Globigerinacea and Corbicula, the ostra-

    codes genera Cypridelis and Heterocypris, pelecypods,

    gastropods, cirripeds, decapods, crabs, sh skeleton frag-ments, 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 andvery ne- to medium- grained sand-

    stone interbedded with red to light-brown, ripple-lami-

    nated sandstone couplets. The proportion of mudstone

    bodies dominate over the sandstone (Fig. 5b). The sand-

    stones show erosive channel structure and ning-upward

    trends.These sand bodies contain cross-bedding, climb-

    ing ripples, gypsum veins, rip-up clasts and burrows.

    Mudstone sandstone couplets contain mottled soil, de-

    siccation 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 lacus-

    trine, tidal, shoreline and brackish to shallow marine en-

    vironments, 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 lacustrine-

    shallow marine fossils and the lithofacies.The mudstone-

    dominated terrestrial facies of the Yecua Formation to the

    west and south are products of uvial overbank and chan-nel processes, with occasional lacustrine and mudat set-

    tings. Hulka et al. (in press) placed these variations in a

    regional context and argued that the marginal marine fa-

    cies of the Yecua Formation represented a marine incur-

    sion 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 fora-

    minifera to be $14 7 Ma (Padula & Reyes, 1958; Marshall

    etal.,1993; Hulka etal., in press) and11 7Ma (Moretti etal.,

    1996). Recently published 40Ar/39Ar radiometric ages of

    10.49 0.33 and 9.41 0.52 Ma ( Hulka, 2005) on inter-

    bedded tus in the Yecua Formation uvial facies fromthe Emborozu and Nogalitos sections matches the esti-

    mated biostratigraphic age from the marine facies. By ana-

    logy, 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 Ar-

    gentinas border with Bolivia that yielded an age of

    9.95 0.34 Ma ( Echavarria etal., 2003).

    Tariquia formation

    TheTariquia Formation (Russo,1959; Ayaviri,1964) is up to

    3800- m- thick and overlies the Yecua Formation with gra-

    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. Intra-

    formational rip-up clasts and reworked calcareous no-

    dules are common. The sandstone bodies have multi-

    storey channel architecture and an overall coarsening-

    and 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 (Bua-

    tois etal., in press) in the thick-bedded channelized sand-

    stone and in mudstone and sandstone couplets are

    common.The trace fossils disrupt the primary sedimen-

    tary 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 mea-

    surements indicate a mean transport towards the east

    (Fig. 4). In theTariquia Formation, the sandstone propor-

    tion 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 depos-

    ited within major channels, whereas the thin-bedded

    sandstone units indicate deposits from crevasse channels.

    In overbank areas,mudstone and sheet sandstone were de-

    posited by crevassing splayand suspension fallout. Follow-

    ingUba etal. (2005),we interpret theTariquia Formation as

    a product of a low-gradient, high- sedimentation, channe-

    lized anastomosing stream and associated thick ood-

    plains on a distal uvial megafan. This interpretation issupported by the laterally extensive channel geometry, ag-

    grading 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 in-

    ferred depositional processes and lithofacies are similar

    to active modern meganfans in the Chaco plain that re-

    ceive sediment from large uvial networks (Horton & De-

    Celles, 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

    r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170 153

    Evolutionof thelateCenozoicChaco foreland basin

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    10/26

    high degree of bioturbation in theTariquia Formation sug-

    gest long periods of channel abandonment and coloniza-

    tion by insects (Buatois et al., in press). The well-

    developed upward coarsening and thickening trend sug-

    gests a systematic stratigraphic development governed by

    either long-term eastward propagation of the fold^thrust-

    belt and/or the expansion of drainage networks. Uba et al.

    (2005) postulated a shift in climate from a semi-arid to ahumid condition during the deposition of the Tariquia

    strata. The Tariquia Formation age is late Miocene

    (Chasicoan-Huaquerian) based on biostratigraphy (Mar-

    shall & 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 ab-

    sence of a well-constrained age for this unit, we use the

    imprecise age of 7^6 Ma for the deposition duration for

    this formation.

    Guandacay formation

    The up-to-1500-m-thick Guandacay Formation consists

    of conglomerate, sandstone and mudstone (Jimenez-Mir-

    anda & Lopez-Murillo,1971) (Fig. 5d).The granule-cobble

    conglomerate shows sheet-like and lenticular geometry,

    clast-supported, polymictic, a coarsening- and thicken-

    ing-upward trend, massive to inversely graded, well-de-

    veloped imbrication and basal s cour 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 sandstonebodies consist of trough cross-,planar, ripple and horizon-

    tal 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 coalseams,poorly pre-

    served bioturbation, and weakly developed mottled soils

    are present (Uba etal., 2005).The conglomerates generally

    thicken and coarsen upsection and to the west. The con-

    glomerate 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 andsandstone. Palaeocurrent measurements indicate a north-

    east-to - southeast-directed ow (Fig. 4).

    The conglomerate and sandstone lithofacies provide

    evidence of deposition in uctuating, high-energy, bed-

    load-dominated large uvial channels, anked by ood -

    plains, 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 proximalbraided

    setting on a medial uvial megafan, similar to those that

    deposited the Camargo Formation and that drain the

    modern central Andean (Horton & DeCelles, 2001; De-

    Celles & Horton, 2003). Lenses of coal suggest the pre-

    sence of a ponded area and vegetation, and therefore, a

    humid palaeoclimate (Uba et al., 20 05). 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 over-

    bank aggradation rate( Bridge, 2003).The contact betweenthe Tariquia and the overlying Guandacay Formation is

    unconformable (Moretti et al., 1996; Echavarria et al.,

    2003), approximately 6 Ma in age (Moretti etal., 1996), and

    is marked by a distinct increase in mean grain size. Hulka

    (2005) estimated the top of this formation at 2.1 0.2Ma

    based on 40Ar/39Ar dating of tu at its contact to the Em-

    borozu Formation in the Abapo section (Fig. 2).The age of

    the Guandacay Formation is therefore late Miocene to

    Early Pliocene (6^2.1Ma).

    Emborozu formation

    The Emborozu Formation (Ayaviri, 1967) is exposed only

    in the northeast (Abapo Section) and within synclines in

    the southwestern (Emborozu and Nogalitos ; Fig. 2) study

    area. The up-to-2000-m-thick, conglomeratic upward-

    coarsening strata of this formation cap the foreland strati-

    graphic succession in the Chaco Basin. In outcrops near

    the present Subandean topographic front, growth struc-

    tures occur (Echavarria et al., 2003), documenting a

    syndeformational origin. The Emborozu Formation is

    dominatedby an up-to- 60-m -thick,cobble-bouldercon-

    glomerate 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 scoursurfaces, 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-coarse-

    grained, sheet-like sandstone with horizontal, trough

    cross-, ripple- and planar stratication. The single-

    storey, 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,lat-

    eral continuity, amalgamation and maximum grain s ize ofthe conglomerate and sand bodies increase and the per-

    centage of overbank nes decreases. The palaeodrainage

    pattern shows a northeast-to- southeast-directed ow

    (Fig. 4).

    Uba etal. (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 mega-

    fan (Horton & DeCelles, 2001; Uba etal., 2005). The sharp

    scour surfaces may represent discretechannels or an amal-

    gamation of scour as a result of avulsion events and chan-

    nel abandonment. The Emborozu Formation overlies the

    r 2006 The Authors. Journal compilat ion r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145 170154

    C. E. Ubaet al.

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    11/26

    Guandacay Formation with a well- dened regional angu-

    lar 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

    2.1 0.2 Ma (tu; Abapo section; Ar^Ar on mica) estimate

    by Hulka (2005) agrees relatively well with 1.8Ma docu-

    mented by Echavarria et al. (2003) for correlative strata in

    Argentina. Consequently, 2.1 0.2Ma is used herein as

    the basal age of the formation.

    SEISMIC STRATIGRAPHY OF THE LATECENOZOIC 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 se-quentially N1 to N5. These are delineated by discontinu-

    ities that coincide with changes in seismic facies and that

    can be correlated with wire-line logs. Seismic facies attri-

    butes include prominent reectors, termination geometry

    (onlap, toplap, downlap and truncation), reection cong-

    uration, 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 cor-

    responding lithofacies and depositional environments are

    calibrated by well data.

    Package N1

    The base of N1is a prominent, readily traceable reector of

    high amplitude, medium frequencyand medium continu-

    ity across most of the study area, marking the contact be-

    tween the late Cenozoic foreland basin and underlying

    Mesozoic strata (Fig. 8a). In some seismic sections, the un-

    derlying Mesozoic strata show diuse toplap and trunca-tions with low angular geometry. In wire-line logs, this

    contact shows an abrupt increase from $100 to 170 Om

    in ILD and an immediate drop from $90 to 40 in DTre-

    sponses (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, N1displays internally lateral extensive, low- to med-

    ium-amplitude, subparallel, discontinuous, low-fre-

    quency 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 Om 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-

    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.

    r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170 155

    Evolutionof thelateCenozoicChaco foreland basin

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    12/26

    formity. It could also be a condensed or non- deposition

    surface with minor erosion (e.g. Mitchum etal.,1977; Sher-

    i & Geldart, 1995). However, the quality of the seismic

    data and the overall low thickness ofN1do not allowa clear

    dierentiation between toplap and erosional truncation.

    Distal onlap and the overall wedge form of N1on the Alto

    de Izozog High indicate basin progradation. We interpret

    package N1 as a sand- dominated aggradational uvial sys-

    tem.This interpretation is based on seismic facies charac-

    teristic (variable-amplitude, discontinuous and low-

    frequency reection) combined with low GR and DTand

    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 over-

    lying mudstone- dominated N2 and the underlying, sand-

    stone-dominated Mesozoic rocks also suggests a change

    in lithology and probably a high degree of cementation or

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

    located approximately 40 km farther south.

    r 2006 The Authors. Journal compilat ion r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145 170156

    C. E. Ubaet al.

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    13/26

    pedogenesis. We interpret the strong seismic facies and

    sharp wire-line log relations at the contact of N1 and N2

    to reect an unc onformity. We correlate N1 based on theseismic facies and wire-line log characteristics mentioned

    above as the subsurface equivalent of the Petaca Forma-

    tion.

    Package N2

    The top of N2 is a laterally continuous, high-to-moder-

    ate-amplitude reector (Fig. 8a). The GR, ILD and DT

    logs do not show a sharp dierence but rather a gradual re-

    sponse at the contact to N3 (Fig. 8a). N2 is overall wedge-

    shaped, with a maximum thickness of more than $450m

    (0.20 ms) in the west.To the east, N2 terminates with onlapgeometries 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, discontin-

    uous, low-amplitude, semi transparent and chaotic reec-

    tions,coupledwith low acoustic impedance,occur (Figs 8a

    and 10). Wire-line logs show $60^120 API in GR, 95^85

    Om in ILD and 65^85 in DTvalues. However, GR and

    DTvalues 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 DTvalues increase upsection. GR

    logs show an irregular or serrated response (Cant, 1992;

    Emery & Myers, 1996) and small-s cale variability in values

    as indicated by numerous thin cycles with ning-upward

    trends (Fig.11).

    The variable-amplitude, discontinuous, semi-trans-

    parent, internal structure, higher GR and lower DTvalues

    are typical of a poorly stratied mudstone-dominated sys-

    tem,deposited mainlyby suspension settling and subordi-

    nate channel settings (Cant, 1992; Alves etal., 2003). Based

    on the seismic facies and well-log characteristics, we can

    interpret the N2 package as deposition in varied settingssuch as shallow marine, lacustrine and uvial environ-

    ments (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 in-

    terpretation 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 fre-

    quency (Sangree & Widmier,1977; Badley, 1985; Cant, 1992;

    Sheri & Geldart, 1995). As N2 thickens westward towards

    its depocentre, it develops varied-amplitude and low-

    continuity reections and dened-clinoforms. The

    small-scale variability observed in the GR, DT and ILD

    Fig. 10. Segment of a W-E uninterpreted and interpreted migrated seismic line along approximately 201450S showing the ve late

    Cenozoic sequences and thrusting and folding of the foreland sequences.The location of the line is shown in Fig. 2.

    r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170 157

    Evolutionof thelateCenozoicChaco foreland basin

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    14/26

    proles largely represents variation in depositional energy

    associated with high-frequency cyclicity (Cant, 1992; Em-

    ery & Myers, 1996). The gradual GR and DTresponses at

    N2^N3 contact suggest a fairly steady change of deposi-

    tional environment between thes e two packages.The e ast-

    ward 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-ampli-

    tude reector; the GR and DTshow decrease and decrease

    in values, respectively (Figs 8 and 9).The laterally continu-

    ous, moderate- to high-amplitude reector marks the top

    of this package and a transitional contact to N4. In wire-

    line logs, this contact is marked by a relatively sharp low

    GR and high ILD and DTresponses (Figs 8b and 9). N3

    shows a maximum thickness of $1500m ( $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, discon-

    tinuous, subparallel, low-frequency, semi-transparent

    and hummocky reectors (Figs 10 and 12). However, the

    seismic facies changes upward to more moderate-low con-

    tinuous, varied-amplitude, less chaotic and less hum-

    mocky reectors and wedge-sheet external forms. The

    proportion of clinoform, hummocky, low-frequency re-

    ectors increases to the east. The clinoforms show east-

    oriented downlap onto, and appear to coalesce with, med-

    ium-amplitude reectors. The log character of the N3

    package is distinguished from the underlying N2 package

    because it contains relatively lower GR (30^90API),higher

    DT (70^100) and higher ILD (95^160) responses. In addi-

    tion, it shows a thicker and larger percent of a low GR re-

    sponse compared with the N2 package, with the percent

    and thickness of low GR and DTresponses increasing sig-

    nicantly 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, moder-

    ately to discontinuous seismic facies, coupledwith low GR

    Fig.11. Segment of a W-E uninterpreted and interpreted migrated seismic line along approximately 191300S showing pinch out of the

    late Cenozoic sequences.The location of the line shown is in Fig. 2.

    r 2006 The Authors. Journal compilat ion r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145 170158

    C. E. Ubaet al.

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    15/26

    and high DT response of package N3, as alternations of

    sandstone and mudstone deposited by s andstone-domi-

    nated uvial deposition (e.g. Sangree & Widmier, 1977;

    Cant, 1992; Sheri & Geldart, 1995; Emery & Myers,

    1996). The low to high amplitude, moderate^low continu -

    ity, lens- sheet external forms, and serrated- bell shape GR

    and DTresponses suggest channels aggrading into ood-

    plains (Cant, 1992). This interpretation is further sup-

    ported by the upsection change in seismic facies (e.g.

    chaotic, hummocky, semi-transparent, combined withup -

    section decrease in the GR values and thickness of lowGR

    response) that indicate an upsection increase in the pro-

    portion of sandbodies and bed thickness. The serrated-

    and bell-shaped GR responses suggest multiple ning-

    upward trends and variable depositional energy. The

    semi-transparent and chaotic seismic features represent

    lack of stratication. The westward-thickening wedge-

    shaped 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 basin-

    ward shift in facies and depocentre location. We assign

    the N3 package to theTariquia Formation because of inter-

    pretation 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 termina-

    tions of N4 reectors on N5 (Figs 8d and 13, inset photo),

    accompanied by abrupt breaks on GR, DTand ILD logs

    (Figs 9, 12 and14). Figures 8d and13 show that this contact

    is a well-dened angular unconformity.The base of N4 is

    Fig.12. W-Ewire-line logs at about 211S illustrating aspects of sequence boundaries.The depositional sequences identied can b e

    correlated to near outcrops. See Fig. 2 for location.

    r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170 159

    Evolutionof thelateCenozoicChaco foreland basin

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    16/26

    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 ap-

    proximately1500 m ( $750 ms) thinning to the eastto 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-am-

    plitude and -frequency, and moderate^low continuity re-ectors (Figs 8c, 9, 10 and 13). Package N4 shows a wedge-

    shaped, 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 LaVer-

    tiente 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 s errated shape.

    The high GR and DT values, cylindrical shape and

    thickness, combined with parallel-subparallel, varied-

    amplitude, 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 in-

    tercalated-thin-serrated GR and DTresponse as alternat-

    ing 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 in-

    crease in thickness indicate a thickening- and coarsen-

    ing-upward trend. The seismic facies attributes and wire-

    line log characteristics of package N4 described above are

    analogous to the sandstone-conglomerate-dominated

    Guandacay Formation.The top of N4 marks a local angu -

    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 lo cation of the line is shown in Fig. 2.

    r 2006 The Authors. Journal compilat ion r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145 170160

    C. E. Ubaet al.

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    17/26

    lar unconformity.The westward thickening,wedge geome-

    try 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 ex-

    isted 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 re-

    ector implies a strong acoustic impedance contrast and

    likely represents the erosional surface or an angular regio-

    nal unconformity (Fig.12, inset; Dunn etal., 1995; Moretti

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

    2003).

    Package N5

    The base ofN5 is dened bya pronouncedthick high-am-

    plitude, 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 variable-

    amplitude and moderately continuous, high frequency

    Fig.14. W-Ewire-line logs at about191S illustrating aspects of sequence boundaries.The depositional s equences identied can be

    correlated to near outcrops. See Fig. 2 for location.

    r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170 161

    Evolutionof thelateCenozoicChaco foreland basin

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    18/26

    reectors. However, well-log data are not available for

    the top of N5. This package shows a maximum thickness

    of $2000m ($1.0ms) 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 ofvariable amplitude,low continuity and

    low frequency (Figs 8d, 9 and 12). The external forms pre-sent are sheet andwedge geometries.The seismic sections

    show a large vertical spacing of reectors (Figs 8d and 13).

    Growth structures above the tips of major fault-propaga-

    tion folds display characteristic fan- shaped reectors with

    upward-decreasing dip. Only one of these faults (Man-

    deyepecuaFault) can be mapped throughoutthe study area

    (Figs 10 and13). No wire-line data are available to ascertain

    the N5 lithological and sedimentological characteristics.

    Notwithstanding the absence of wire-line logs, we in-

    terpret the N5 package to consist of an alternation of con -

    glomerate and sandstone deposited in a conglomerate-

    dominated uvial setting (e.g. Cant, 1992), probably a large

    alluvial fan. This interpretation is supported by the sub -

    parallel-to-parallel, variable-amplitude, and low continu-

    ous, sheet- to wedge-shaped, and high vertical spacing of

    reectors, which probably suggest thick conglomerate

    lithofacies. The overall thickening- and coarsening-up-

    ward 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 in-

    terpretation for the N5 package permit its correlationwith

    the surface Emborozu Formation.

    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 bur-

    ial.We constructed isopach maps for ve time periods (oc-

    casionally 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 accommoda-

    tion space to infer that the Chaco foreland basin experi-

    enced 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 inter-

    vals 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 PetacaFormation

    (Fig. 15a) shows a regional and broad area along the Villa-

    montes-Camiri axiswitho50 m of strata, possibly reect-

    ing a structural high. This ridge is anked by up to 250 m

    of strata cratonward and4100 m of strata orogenward, re-

    spectively. 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 for-

    mation (e.g. Gubbels et al., 1993; Moretti et al., 1996). The

    very low available accommodation space during this peri-

    odwas probably a result of a long time interval of basin sta-bility 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 etal. (1993) and Moretti etal. (1996) reported a

    similar lack of thickness variation for the Yecua Formation.

    In contrast, Fig.15b shows a distinctive westward thicken-

    ing, reaching a maximum thickness of $600 m in outcrop

    (e.g. Emborozu section). This westward thickening and in -

    crease in sandstone proportion of the Yecua strata is also

    documented in seismic sections (Figs 6, 7, 10 and13). Dur-

    ing 14 7 Ma, exural foreland basin subsidence as a result

    of thrusting episode that was probably centred in the pre-

    sent-day Interandean- or Subandean Zone (Coudert etal.,

    1995; Moretti etal.,1996; Echavarria etal., 2003) led to crea-

    tion ofaccommodation space in thewestern part of the ba-

    sin. Flemings & Jordan (1989) and Sinclair etal. (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 de-

    positional slope was probably too low to produce large

    coarse-grained deposits or there was a long lag-time be-

    tween 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-thick-

    ening 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 $1mmyear1 corresponding during this time in the

    Subandean Zone (Coudert et al., 1995; Echavarria et al.,

    2003).The high accommodation space and sedimentationrate 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 s ediment load within the ba-

    sin, thereby resulting in an increased basin wavelength

    and an increase in sediment supply, causing migration of

    basin-magin facies.We proposed that the high sedimenta-

    tion 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 etal., 2005), which led to high precipitation and high

    denudation, thus high sediment supply.The high denuda-

    tion in the west and high sedimentation in the central part

    r 2006 The Authors. Journal compilat ion r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145 170162

    C. E. Ubaet al.

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    19/26

    of thebasin, coupledwith a humid climate,might have also

    resulted in the migration of the proximal Guandacay con-

    glomerate 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

    Fig.15. Isopach maps of the ve lateCeno zoic units of the Chaco foreland basin in the study area based on measured surface sections,

    interpre ted 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.

    r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170 163

    Evolutionof thelateCenozoicChaco foreland basin

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    20/26

    to 0 at the Alto de Izozog basement high. In contrast, the

    seismic facies of the Emborozu Formation shows a combi-

    nation of symmetrical and asymmetrical geometries.

    Underfilled and overfilled stages of the Chacoforeland basin

    The sedimentary and seismic interpretations, as summar-

    ized in Fig. 15, show an overall asymmetrically westward-

    thickening 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 sedimen-

    tation 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 con-

    sistent easterly transverse sediment dispersal and sedi-

    mentary 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 mudstone-

    dominated lacustrine and marginal marine facies, indicate

    an underlled stage (e.g. Flemings & Jordan, 1989), resem-

    bling the underlled phase of the Western Taiwan and Ca-

    margo Basins, Bolivia (Covey, 1986; DeCelles & Horton,

    2003). Because the sediment-accumulation rate decreased

    ($600m in $7 Ma) and the uvial pattern was modied

    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 theTariquia, Guandacay and

    Emborozu Formations, the predominantly uvial depos-

    its, coarsening-upward trend, increase in single-intercon-

    nected- channel geometry and high avulsion frequencyindicate that the Chaco foreland basin shifted to an over-

    lled stage. Furthermore, this stage is expressed by a pre-

    dominance 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 ofca. 300 km (Horton

    & DeCelles, 1997). It forms a NNE- SSW-trending struc-

    tural high bordering the e astern 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 etal., 1995; Moretti etal., 1996; Horton & De-

    Celles, 1997; DeCelles & Horton, 2003). However, the in-

    terpreted 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 km1 and a heat ow of more than

    100 mWm 2 at the Alto de Izozog.These values are extre-

    mely high compared with the gradient and heat ow values

    of 26 1C km1 and 52mWm 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, onlyca.

    70km separate the exposed basement rocks from the to -

    pographic front of the Subandean ranges, implying thatthe combined width of the wedge-top and foredeep depo -

    centre reaches barely100 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

    of460 km (Stewart & Watts, 1997; Tassara, 2005) is ob -

    served in the southern Central Andes. The low cross-

    strike width between assumed forebulge location and de-

    formation front strongly disagrees with values of ca.

    4300 km for most other foreland basin systems (e.g. De-

    Celles & 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

    r 2006 The Authors. Journal compilat ion r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145 170164

    C. E. Ubaet al.

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    21/26

    althoughCoudert etal. (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 An-

    tler orogeny of the western United States (Giles & Dickin-

    son, 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 sys-

    tem is known below the Chaco foreland basin.We therefore

    consider the Alto de Izozog an unlikely candidate for a re-

    cent 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 Villa-

    montes-Camiri structural high and the axis itself, with

    only ca. 50m thickness of thePetaca Formation,as thefore-

    bulge (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

    Fig.16. Structural cross- sections (modied after Dunn etal., 1995; Moretti etal., 1996; Baby etal., 1997; Kley etal., 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.

    r 2006 The Authors. Journal compilation r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145^170 165

    Evolutionof thelateCenozoicChaco foreland basin

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    22/26

    and10 Ma (e.g. Crampton & Allen, 1995;White etal., 2002),

    similar to the forebulge migration of the Camargo basin

    (Horton & DeCelles, 1997; DeCelles & Horton, 2003). The

    presence of a forebulge-backbulge depocentre during Pe-

    taca time ( $27 14 Ma) is also supported by the westward-

    directed 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 etal.,1996; Echavarria

    etal., 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, Cou-

    dert et al. (1995) and Echavarria et al. (2003) documented a

    subsidence rate of 1m Ma1. Exhumation and structural

    data indicate that the Subandean Zone began to be de-

    formed and exhumed in this interval (Kley et al., 1996;

    Moretti etal., 1996; Echavarria et al., 2003; Ege, 2004). The

    underlled stage of the basin is due to the several-million-

    years time lag between loading to the west of the wideningbasin 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 con-

    trasts with its interpretation as a backbulge depocentre by

    Marshall etal. (1993), but agrees with the interpretation of

    DeCelles & Horton (2003). The distal foredeep develop-

    ment 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 For-

    mation represents a medial-foredeep depocentre ll (Fig.

    16c). This interpretation is suppor ted by the long-lasting

    and substantial creation of accommodation space and a

    high accumulation rate (e.g. Echavarria et al., 2003), a re-

    sulting westward increase in large-scale sandstone-domi-

    nated 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 ar rived

    in the study area ( Fig. 16d). Westward thickening, consis -

    tent eastward-directed palaeocurrents, and a westward in-

    crease in the proportion of conglomerates provide further

    evidence for the prese nce of the proximal foredeep. Strik-

    ingly similar facies and geometries of foredeep depocen-

    tres have bee n documented by Flemings & Jordan (1989),Sinclair & Allen (1992), DeCelles & Horton (2003) and Ca-

    tuneanu (2004) for other foreland basins worldwide.

    We interpret the Emborozu Format ion as representing

    the wedge-top depocentre (Fig. 16e). These thickening-

    and coarsening- upward strata are apparently regionally

    restricted, related to specic thrusts, and show wedge-

    shaped, high-amplitude reectors and growth structures

    above active blind thrusts (e.g. Fig. 12). The contact be-

    tween 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 com-

    pilation of total-shortening and shortening rates in the

    Subandean (Echavarria etal., 2003; Elgeretal., 2005; Onck-

    en 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

    Fig. 17. Diagram illustrating (a) timing and total shortening rate compiled from Echavarria etal. (2003) and Oncken etal. (in press); and

    the possible major thrusting episode. (b) The rate and propagation of exhumation front of the Central Andes (after Ege, 2004).

    r 2006 The Authors. Journal compilat ion r 2006 Blackwell Publishing Ltd, Basin Research, 18, 145 170166

    C. E. Ubaet al.

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    23/26

    Tupiza (Eastern Cordillera)and Villamontes (western Cha-

    co plain) corresponding to eastward propagation of the de-

    formation 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 re-

    late it to Andean tectonics.We speculate that the succes-

    sion of palaeosols may be older than the Andean orogeny

    (Cretaceous? or Eocene?). The subsequent reworking of

    the palaeosols and the sand-mudstone deposition may re-

    present 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 expressedby

    the onset of thrusting to the west in the Tupiza region in

    the Eastern Cordillera (He rail etal., 1996; Kley etal., 1997)

    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 re-

    sulted 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 devel-

    opment 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 sedi-

    mentary basins of the Tupiza region are associated with

    fold^thrust deformation, whereas apatite ssion track ages

    from the Eastern Cordillera show a decrease in coolingages from $38 to 17 Ma (Ege, 2004). During the deposi -

    tion of the Petaca Formation, the structural, sedimentolo -

    gical and thermochronologic data indicate major

    structural growth and crustal thickening within the East-

    ern 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-angle-

    basement thrusting and the arrival of the deformation

    front in the westernmost part of the Subandean Zone (e.g.

    Gubbels etal., 1993; Coudert etal., 1995; Moretti etal.1996;

    Kley et al., 1997; Echavarria et al., 2003; Ege, 2004). Yecuastrata record basement-imbricated uplift in the Interan-

    dean and eastern propagation of the fold^thrust system in

    the eastern Interandean or Subandean Zone. Coudert etal.

    (1995) and Echavarria et al. (2003) suggest that the sedi-

    mentation rate increased rapidly during this time. How-

    ever, 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 theTariquia and Guan-

    dacay Formations to exural response due to thehigh sedi-

    mentation as a result of high erosion of shortened and

    uplifted regions in the Interandean and Subandean ranges

    during tectonic quiescence, notwithstanding the occur-

    rence of minor thrusting (e.g. Echavarria et al., 2003) in

    the basin, such as the development of the local, 6^2.1 Ma-

    old La Vertiente structure (Moretti et al., 1996; Fig. 12). It

    suggests that the 14^7Ma 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 increasedaccommodation space and deposition of more than 3500-

    m-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

    etal., 1996; Kley etal., 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 etal. (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 etal. (2003)postulated a young-

    er age of approximately 2.5 Ma for the in-s equence thrust,

    with out-of-sequence reactivation of older structures in

    the west at 2^2.2Ma. Our study agrees with the results of

    Echavarria et al. (2003) that 2.1Ma (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 re-

    present the reactivation of the older Nogalitos range (Fig.

    2), thus forming an out-of sequence intermontane basin

    (e.g. Echavarria etal., 2003).

    CONCLUSIONS

    The combination of seismic stratigraphy and outcrop fa-

    cies 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 depocen-

    tres, driven primarily by crustal shortening and tectonic

    loading. Signicant exural subsidence developed since

    the late Miocene. Variably created accommodation spacewas 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 wedge-

    top deposition of the Emborozu Formation.

    Three major tectonic episodes are expressed in the fore-

    land strata:(1)the late Oligocene uplift of theEasternCor-

    dillera initiated the forelanddevelopment and is expressed

    as relatively steep to low-angle basement thrusts; (2) late

    Miocene formation of the Intrandean/Subandean fold-

    and-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

    Evolutionof thelateCenozoicChaco foreland basin

  • 8/3/2019 Evolutia Bazinului de Foreland Chaco

    24/26

    and (3) late Pliocene shortening generated the overlled,

    coarse clastic wedges of the Emborozu Formation. Fore-

    land 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 thick-

    ening-upward trend and stratigraphic architectures docu-

    ment a propagating fold^thrust belt and correspondingforeland basin depocentres (DeCelles & Giles, 1996). Si-

    milar migration of depocentre with time and foreland ba-

    sin architecture has been recorded for numerous other

    basins worldwide, such as the Karoo foreland basins (Ca-

    tuneanu et al., 1999), Taiwan (Covey, 1986) and North Al-

    pine (Schlunegger etal., 1997; Pner etal., 2002).

    ACKNOWLEDGEMENTS

    This paper is part of a PhD thesis by the rst author at the

    Freie Universitt Berlin, Germany. The authors were sup-

    ported nancially by the DFG through the Sonder-

    forschungsbereich (SFB) 267 and logistically by Chaco

    S.A., SantaCruz, Bolivia.We are indebtedto Oscar Aranibar,

    Fernando Alegria and Nigel Robinson of Chaco S.A. for

    their assistance.Thanks are also due to David Tuno 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 geody-

    namics. We are grateful for comments provided by the re-

    viewers Jonas Kley, Fritz Schlunegger and Patrice Baby,

    which greatly improved the early revisions of this manu-script.

    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. In-

    forme 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 r ift basins: examples from the Lu-

    sitanian Basin of West Iberia. Sediment. Geol., 162, 273^303.

    Baby, P., He rail, G., Salinas, R. & Sempere,T. (1992) Geome-

    try and kinematic evolution of passive roof duplexes deduced

    from cross-section balancing: example from the foreland

    thrust system of the southern Bolivian Subandean Zone. Tec-

    tonics, 11, 523 536.

    Baby, P., Moretti, I., Guillier, B., Limachi, E., Mendez, E.,

    Oller,J. & Specht, M. (1995) Petroleum system of the north-

    ern and central Bolivian Subandean Zone. In: Petroleum Basins

    of South America (Ed. by A.J. Tankard, R. Sua rez Soruco & H.J.

    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.

    Badley, M.E. (1985) Practical Seismic Interpretation. Intern. Hu-

    man 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 o ccurrence in the Subandean

    fold and thrust belt of northern Argentina. In: Petroleum Basins

    of South America (Ed. by A.J. Tankard, R. Sua rez Soruco & H.J.

    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. Bla ckwell, Oxford, 491pp.

    Buatois, L.A., Uba, C.E., Ma' ngano, M.G, Hulka, C. & Heu-

    beck, C. (in press). Deep bioturbation in continent al environ-

    ments: 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 thrust ing in terrestrial foreland basins; appli-

    cations to the northwestern Himalaya. In: New Perspectives in

    Basin Analysis (Ed. by K.L. Kleinspehn & C. Paola) Springer-

    Verlag, NewYork.

    Cant, D.J. (1992) Subsurfacefacies analysis.In : FaciesModels:Re-

    sponse 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