journal of south american earth sciences basin... · puncoviscana and medina formations (bossi,...

18
Tectonic inversion in a segmented foreland basin from extensional to piggy back settings: The Tucumán basin in NW Argentina Diego Nicolas Iaffa a, * , F. Sàbat a , D. Bello b , O. Ferrer a , R. Mon c , A.A. Gutierrez c a GEOMODELS Research Institute, Department de Geodinàmica i Geofísica, Facultat de Geologia, Universitat de Barcelona, C/Martí i Franquès s/n, 08028 Barcelona, Spain b Ecopetrol, D.C. Edicio Principal Cr 13 No. 36 e 24, Bogotá, Colombia c Dept. de Geología, Facultad de Ciencias Naturales e Instituto Miguel Lillo, Universidad Nacional de Tucumán, Miguel Lillo 205 4000 Tucumán, Argentina article info Article history: Received 20 July 2010 Accepted 14 February 2011 Keywords: Central Andes Argentina Cretaceous rift Foreland basin Tectonic inversion Growth strata abstract The Tucumán foreland basin is bounded by: 1) basement cored ranges with elevations over 6000 m in the west; 2) inverted extensional grabens to the north; 3) basement thrust blocks in the south and 4) basement cored small ranges in the east. This foreland basin is located between two geological provinces: the Sierras Pampeanas and the Santa Bárbara system. Cretaceous Salta rifting extended southwards covering the entire eastern part of the province of Tucumán in NW Argentina. Syn-rift and post-rift deposits can be recognized in accordance with their architectural geometries. Foreland basin sediments progressively covered the rift deposits as the Andean orogen propagated towards the east. Despite some early studies, the Tucumán basin is poorly documented. For the present study, 44 sections of 2D seismic surveys amounting to more than 730 km were used to describe the structure and the depositional evolution of the basin. The present structure is the result of a long sequence of events that includes a compressional deformation during the Paleozoic, a rifting stage during the Cretaceous and a foreland stage during the late Cenozoic. Although tectonic inversion, which has played a role during the foreland stage since the Miocene, can be observed in many sectors of the basin, it is more prominent along the margins. Reactivation of old basement discontinuities and inversion of Cretaceous normal faults produced the compartmentalization of the foreland, giving rise to the present shape of the Tucumán basin. This evolution is recorded in the Neogene deposits. Ó 2011 Elsevier Ltd. All rights reserved. 1. Introduction The Tucumán Basin is located at the foot of the Andean Chain, between 26 and 28 south and 65 e66 west (Fig. 1), and straddles the border of different geological provinces (Ramos, 1999). The basin is bounded by the Sierra de Aconquija and the Cumbres Calchaquíes to the west (Fig. 2). These are the northern ranges of the Sierras Pampeanas, which are characterized by basement uplifted blocks that cut the Pampean at (Allmendinger et al., 1983; Costa et al., 1999; González Bonorino, 1950; Jordan et al., 1983). The basin is bounded by the Sierra de San Javier, Sierra de Medina and the Sierra de Ramada to the north. These ranges form the southern units of the Santa Bárbara system due to tectonic inversion of Cretaceous extensional faults and to thrusting of syn-rift deposits over neogene layers (Abascal, 2005; Kley and Monaldi, 2002; Kley et al., 2005; Ramos, 1999). In the proximity of the Tucuman basin towards NNW is the southern tip of the Eastern Cordillera (Carrera et al., 2006)(Fig. 1). To the east of the Tucumán basin lies the Chacoparanaense plain (Fig. 2), which has an average elevation of 200 m above sea level and is separated from the Tucumán Basin by the Sierra de Guasayán. This NeS striking range attains a maximum elevation of 700 m (Fig. 2), constitutes the eastern limit of the Tucuman basin and acted as a structural high during Paleozoic orogenies (Cristallini et al., 2004). The Sierra de Aconquija and the Cumbres Calchaquíes, to the west of the Tucuman basin, form an orographic barrier with peaks over 6000 m (Fig. 2). The step between the Sierra de Aconquija and the Tucumán plain reaches about 5000 m. Further to the west of the Sierra de Aconquija and the Quilmes Range is the southern part of the AltiplanoePuna with a median altitude of 4000 m. This is the highest plateau in the world in an active subduction margin (Allmendinger et al., 1997; Hindle et al., 2005; Isacks, 1988). Further to the south of the Sierra de Aconquija is the Sierra de Ambato, which constitutes the southwestern limit of the Tucumán Basin * Corresponding author. Tel.: þ34 934 035 914; fax: þ34 934 021 340. E-mail address: [email protected] (D.N. Iaffa). Contents lists available at ScienceDirect Journal of South American Earth Sciences journal homepage: www.elsevier.com/locate/jsames 0895-9811/$ e see front matter Ó 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.jsames.2011.02.009 Journal of South American Earth Sciences 31 (2011) 457e474

Upload: duongthien

Post on 09-Sep-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

lable at ScienceDirect

Journal of South American Earth Sciences 31 (2011) 457e474

Contents lists avai

Journal of South American Earth Sciences

journal homepage: www.elsevier .com/locate/ jsames

Tectonic inversion in a segmented foreland basin from extensional to piggyback settings: The Tucumán basin in NW Argentina

Diego Nicolas Iaffa a,*, F. Sàbat a, D. Bello b, O. Ferrer a, R. Mon c, A.A. Gutierrez c

aGEOMODELS Research Institute, Department de Geodinàmica i Geofísica, Facultat de Geologia, Universitat de Barcelona, C/Martí i Franquès s/n, 08028 Barcelona, Spainb Ecopetrol, D.C. Edificio Principal Cr 13 No. 36 e 24, Bogotá, ColombiacDept. de Geología, Facultad de Ciencias Naturales e Instituto Miguel Lillo, Universidad Nacional de Tucumán, Miguel Lillo 205 4000 Tucumán, Argentina

a r t i c l e i n f o

Article history:Received 20 July 2010Accepted 14 February 2011

Keywords:Central AndesArgentinaCretaceous riftForeland basinTectonic inversionGrowth strata

* Corresponding author. Tel.: þ34 934 035 914; faxE-mail address: [email protected] (D.N. Iaffa).

0895-9811/$ e see front matter � 2011 Elsevier Ltd.doi:10.1016/j.jsames.2011.02.009

a b s t r a c t

The Tucumán foreland basin is bounded by: 1) basement cored ranges with elevations over 6000 m inthe west; 2) inverted extensional grabens to the north; 3) basement thrust blocks in the south and4) basement cored small ranges in the east. This foreland basin is located between two geologicalprovinces: the Sierras Pampeanas and the Santa Bárbara system.

Cretaceous Salta rifting extended southwards covering the entire eastern part of the province ofTucumán in NW Argentina. Syn-rift and post-rift deposits can be recognized in accordance with theirarchitectural geometries. Foreland basin sediments progressively covered the rift deposits as the Andeanorogen propagated towards the east.

Despite some early studies, the Tucumán basin is poorly documented. For the present study,44 sections of 2D seismic surveys amounting to more than 730 km were used to describe the structureand the depositional evolution of the basin. The present structure is the result of a long sequence ofevents that includes a compressional deformation during the Paleozoic, a rifting stage during theCretaceous and a foreland stage during the late Cenozoic. Although tectonic inversion, which has playeda role during the foreland stage since the Miocene, can be observed in many sectors of the basin, it ismore prominent along the margins. Reactivation of old basement discontinuities and inversion ofCretaceous normal faults produced the compartmentalization of the foreland, giving rise to the presentshape of the Tucumán basin. This evolution is recorded in the Neogene deposits.

� 2011 Elsevier Ltd. All rights reserved.

1. Introduction

The Tucumán Basin is located at the foot of the Andean Chain,between 26� and 28� south and 65�e66� west (Fig.1), and straddlesthe border of different geological provinces (Ramos, 1999). Thebasin is bounded by the Sierra de Aconquija and the CumbresCalchaquíes to the west (Fig. 2). These are the northern ranges ofthe Sierras Pampeanas, which are characterized by basementuplifted blocks that cut the Pampean flat (Allmendinger et al., 1983;Costa et al., 1999; González Bonorino, 1950; Jordan et al., 1983). Thebasin is bounded by the Sierra de San Javier, Sierra de Medina andthe Sierra de Ramada to the north. These ranges form the southernunits of the Santa Bárbara system due to tectonic inversion ofCretaceous extensional faults and to thrusting of syn-rift depositsover neogene layers (Abascal, 2005; Kley and Monaldi, 2002; Kley

: þ34 934 021 340.

All rights reserved.

et al., 2005; Ramos, 1999). In the proximity of the Tucuman basintowards NNW is the southern tip of the Eastern Cordillera (Carreraet al., 2006) (Fig. 1). To the east of the Tucumán basin lies theChacoparanaense plain (Fig. 2), which has an average elevation of200 m above sea level and is separated from the Tucumán Basin bythe Sierra de Guasayán. This NeS striking range attains a maximumelevation of 700 m (Fig. 2), constitutes the eastern limit of theTucuman basin and acted as a structural high during Paleozoicorogenies (Cristallini et al., 2004).

The Sierra de Aconquija and the Cumbres Calchaquíes, to thewest of the Tucuman basin, form an orographic barrier with peaksover 6000 m (Fig. 2). The step between the Sierra de Aconquija andthe Tucumán plain reaches about 5000m. Further to thewest of theSierra de Aconquija and the Quilmes Range is the southern part ofthe AltiplanoePuna with a median altitude of 4000 m. This is thehighest plateau in the world in an active subduction margin(Allmendinger et al., 1997; Hindle et al., 2005; Isacks, 1988). Furtherto the south of the Sierra de Aconquija is the Sierra de Ambato,which constitutes the southwestern limit of the Tucumán Basin

Fig. 1. Location of the study area in the central Andes in Argentina, Chile and Bolivia with the political borders; the Tucumán province borders are in bold. Main geological provincesare displayed, modified from Ramos (1999) and Hilley and Coutand (2010).

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474458

(Fig. 2). The NNW trending Ambato range is formed by a set ofbasement thrust blocks dipping to the east (Cristallini et al., 2004;Gutiérrez and Mon, 2008; Roy et al., 2006).

The structure of the Tucumán Basin has been studied bya number of authors. Cristallini et al. (2004) reprocessed andinterpreted several seismic sections from different surveys focusingon deep faults and detachments. Pacheco et al. (2000) presenteda sketch-map of the basin in time domains based on the interpre-tation of seismic sections. Other authors used different geophysicalmethods such as gravimetry (Pomposiello et al., 1993) and mag-netotellury (Favetto et al., 2007) to describe the shape and thedepth of the basin.

The aim of the present paper is twofold: 1) to describe theTucumán basin structures, including those resulting from tectonicinversion due to Cretaceous extension and Andean shortening, and2) to elucidate the role played by reactivation and inversion ofearlier faults in the basin evolution.

2. Geological setting

2.1. Stratigraphy

In this sector of the Andean chain, the basement is composed ofPuncoviscana and Medina formations (Bossi, 1969; Turner, 1959)and granitoids of Precambrian to early Cambrian age (Aceñolazaet al., 2002) (Fig. 3). The two previous formations are made up ofgrayish colored metasediments, low to medium grade, bandedschists, which preserved the original lamination. Basement rocksare strongly foliated as a result of the dynamic metamorphismproduced during the accretion of terranes in the Early Paleozoic andsubsequent orogenies (Ramos, 1988). Basement was subsequentlyintruded by Cambrian to Ordovician granitoids (Battaglia, 1982;González Bonorino, 1950; Mon and Hongn, 1991).

Eastwards of the Rosario Fault (Fig. 2), thick series of Paleozoiclayers, belonging to the Chacoparanaense basin, have been

Fig. 2. Geological map of the study area with location of the seismic sections used in this study. Names of ranges and basins are abbreviated (from south to north): P.V. (PipanacoValley), A.B. (Ambato Block), S.Ac. (Sierra de Aconquija), C.A. (Campo Arenal), S.Q. (Sierra de Quilmes), S.M.B. (Santa María Basin), C.C. (Cumbres Calchaquíes), S.S.J. (Sierra de SanJavier), S.L.R. (Sierra de La Ramada), S.M. (Sierra de Medina), S.d.C. (Sierra del Campo), Ch. B. (Choromoro Basin), S.L.C. (Sierra de La Candelaria), S.B. (Sierra del Brete) andAn.B. (Angastaco Basin).

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474 459

described (Fernández Garrasino et al., 2005). They are made up ofSilurian to Devonian clastic rocks (Padula et al., 1967), depositedinitially in an extensional basin and then in a foreland setting(Ramos, 1988). Although these layers have not been identified inthe Tucumán basin (Cristallini et al., 2004) (Fig. 3) or to the west ofthe study area, small outcrops are present in the east of the Sierrade la Candelaria (Fig. 2) (Mángano and Buatois, 2004).

Cretaceous to Paleogene sedimentary rocks unconformablyoverlie basement rocks and have interstratified alkaline volcanics(Galliski and Viramonte, 1988; Turner, 1959). These layers form theSalta Group (Fig. 3), which is attributed to the Salta rifting thatoccurred in an extensional back arc environment (Kley andMonaldi, 2002; Turner, 1959). The basal unit of the Salta Group isthe Pirgüa subgroup, which consists of breccias, conglomerates,sandstones and red beds of continental environments such asalluvial fans, fluvial plains and debris flows (Gómez Omil et al.,1989; Moreno, 1970; Reyes and Salfity, 1973; Salfity andMarquillas, 1981). The Pirgüa subgroup has been interpreted as

Cretaceous syn-rift deposits associated with extensional faults(Turner, 1959). The syn-tectonic character is evidenced by a fangeometry of layers onlapping the basement (Carrera and Muñoz,2008; Comínguez and Ramos, 1995; Cristallini et al., 1997).Starved growing trenches began to be filled with breccias andconglomerates of basement clasts at the base of the sequence.Basaltic lavas and pyroclastic flows attributed to the extensionalstage are intercalated with the basal units (Galliski and Viramonte,1988). Subsequently, these trenches were covered by decreasinggrain sequences of sandstones and shales (Salfity and Marquillas,1994). When the rifting episode ceased, syn-rift deposits werecovered by post-rift layers (Salfity and Marquillas, 1981; Turner,1959). These new sequences overlap the Pirgüa subgroup and thebasement (Boll et al., 1989; Carrera et al., 2006; Cristallini et al.,1997). The Balbuena and Santa Bárbara subgroups are made up ofsandstones, limestones, evaporites and shales deposited in lacus-trine, shallow marine and continental settings (Bonaparte et al.,1977; Marquillas et al., 2003; Moreno, 1970). The Yacoraite

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474460

formation, which is the upper part of the Balbuena subgroup, isformed by limestones and shales (Marquillas and Salfity, 1994). Thisunit is the main oil source in other sub basins of the Salta Rift basinsuch as the Lomas de Olmedo sub-basin (Boll et al., 1989; Turicet al., 1987). The Balbuena and Santa Barbara subgroups filleda sag basin resulting from thermal subsidence after the extension ofthe rifting episode (Bianucci et al., 1981; Comínguez and Ramos,1995; Salfity and Marquillas, 1994). The normal faults of the Saltarifting stage continued to be poorly active in the early stages of thedeposition of the Balbuena subgroup (Bianucci et al., 1982; Kleyet al., 2005). The Balbuena subgroup is equivalent to the Rio Loroformation in the Choromoro Basin (Abascal, 2005). Foreland sedi-ments were progressively covered by post-rift deposits. The tran-sition from post-rift to foreland settings occurred as the Andeandeformation started to propagate towards the east (Carrera et al.,2006; Jordan and Alonso, 1987; Russo and Serraiotto, 1979). Theforeland basin stage can be recognized by the presence of anangular unconformity and a stratigraphic gap of the middle Eocene(Del Papa et al., 2010; Reynolds et al., 2000). Paleogene sandstonesand shales of the Aconquija formation crop out along the easternmargin of the Sierra de Aconquija (González Bonorino, 1950)(Figs. 2 and 3). Paleogene sandstones are overlain by Río Salísandstones and evaporites that are known as the Guasayánformation further to the east (Battaglia, 1982). These layers area key level for correlation and are associated with the middleMiocene Paraná Atlantic marine transgression (Ramos and Alonso,1995; Uliana and Biddle, 1988). The transgressive event was evi-denced by gypsum rich sandstones between continental facies(Battaglia, 1982; Gavriloff and Bossi, 1992). These formations cropout along the periphery of the Tucumán basin in the southern partof the Sierra de Medina, on the western slope of the Sierra deGuasayán and on the eastern slope of the Sierra de Aconquija(Fig. 2). They are also present on the western side of the Sierra deAconquija and in the El Cajón and Santa María valleys (Bossi et al.,2001; Gavriloff and Bossi, 1992; Kleinert and Strecker, 2001;Mortimer et al., 2007; Strecker et al., 1989).

The stratigraphic column of the Tucumán basin culminates inPliocene to Quaternary clastic continental synorogenic sequences(Ramos, 1999) that consist of the following formations: the IndiaMuerta, Acequiones and Ticucho. The India Muerta formation iscomposed of gray sandstones and shales of a fluvial origin (Bossiand Gavriloff, 1998). Overlying the India Muerta formation are theAcequiones and Ticucho formations that consist of coarser cong-lomerates, reflecting the diachronic uplifting of the surroundingranges (González, 2000). These units are grouped in the Las Cañasformation in the Sierra de Gusayán and further to the east of theSierra (Battaglia, 1982) (Fig. 2).

2.2. Regional structure

Tectonic inversionofpreviousextensional faultsplays akey role inthe formation of the Eastern Cordillera (Carrera et al., 2006; Coutandet al., 2001;Grier et al.,1991;Hongnet al., 2007; Kleyet al., 2005) andthe Santa Bárbara System (Bianucci et al.,1982; Cristallini et al.,1997;Kley and Monaldi, 2002,) (Fig. 1).

However, the Sierras Pampeanas, which are composed ofbasement blocks bounded by high angle thrust faults (Jordan et al.,1983), have a deep detachment that resulted from reactivation ofold crustal discontinuities (Allmendiger et al., 1983; GonzálezBonorino, 1950; Jordan et al., 1983). In the southern sector of theSierras Pampeanas, inverted extensional Cretaceous structureshave been documented (Schmidt et al., 1995).

The geological province of the Santa Bárbara system is formed bybasement cored ranges, with syn-rift and post-rift sedimentarylayers cropping out on their flanks (Fig. 1) (Cristallini et al., 1997;

Grier et al., 1991; Kley and Monaldi, 2002). In the study area, theSierradeMedina, Sierrade laCandelaria, Sierrade laRamadaand theSierra del Campo form part of this geological province (Fig. 2). Theserangeswereupliftedbyhighangle thrust faultsdue to reactivationofupper crustal discontinuities that originated in earlier tectoniccycles and to inversion of Cretaceous extensional faults (Kley et al.,1999; Monaldi and Kley, 1997). The Sierra de Medina is boundedalong its southern margin by a thrust fault with a concave shape tothe north, which is very typical of a listric normal fault. This rangeshows syn-rift deposits of the Pirgüa subgroup thrusted over post-rift and foreland layers (Abascal, 2005; Bossi,1969; Iaffa et al., 2008).The Sierra de la Ramada is an NNE trending anticlinewith basementrocks in the core, and is bounded by syn-rift deposits along bothmargins (Fig. 2). This anticline was brought about by compression,inversion and folding of a Cretaceous extensional depocenter(Mángano and Buatois, 2004).

In the study area, the geological province of the Sierras Pam-peanas consists of the Cumbres Calchaquíes and the Sierra deQuilmes, Sierra de San Javier, Sierra de Aconquija, Sierra de Ambatoand the Sierra de Guasayán (Fig. 2). These basement upthrustranges are the northernmost units of this large geological province(Allmendinger et al., 1983; González Bonorino, 1950; Jordan et al.,1983). The Sierras Pampeanas cut the foreland plain and areextensive and have high altitudes (Ramos, 1999). The Sierra de SanJavier is a small NNE trending range, mainly formed by basementrocks (Mon and Suayter, 1973). This range was uplifted by a thrustfault along the eastern margin and shows a normal series of theSalta group as far as the Neogene deposits in its western limb(González, 2000). The Sierra de Aconquija and the Cumbres Cal-chaquíes are uplifted by an active double vergent system formed byhigh angle thrust faults (Cristallini et al., 2004; Drozdzewski andMon, 1999). This type of structure has not reused pre-existingextensional faults, but has reactivated major crustal discontinuities,generating pop-up structures (Sobel and Strecker, 2003). The tworanges show basement and a thin cover thrusted over the Cenozoicsedimentary layers (Fig. 2). The Sierra de Aconquija and the Cum-bres Calchaquíes are separated by the Amaicha northwest trendinglineament, which is probably due to a thrust fault that uplifted theCumbres Calchaquíes over the Sierra de Aconquija (Allmendingeret al., 1983; De Urreiztieta et al., 1996). To the west of Sierra deAconquija is the Santa María basin, which is filled by Neogenesedimentary rocks (Bossi et al., 1997). The uplift of the Sierra deAconquija, Cumbres Calchaquíes and the Sierra de Quilmes hasplayed amajor role in the infilling of the basin (Strecker et al., 1989).The uplifting of these ranges has been attributed to the early Plio-cene given the coarsening up in the sequence sediments and thepresence of gravel size deposits (Bossi et al., 2001; Kleinert andStrecker, 2001). Using the apatite fission-track methodology,Sobel and Strecker (2003) proposed that the uplifting of the Sierrade Aconquija and the Cumbres Calchaquíes started at 6 Ma andcontinues to be active. Neotectonic activity occurred on both sidesof the range, the western flank being the more active one(Cristallini et al., 2004; Strecker et al., 1989). On this slope, Pleis-tocene rock avalanche deposits have been triggered by seismicitynear Campo Arenal (Fauqué and Strecker, 1987) (Fig. 2).

The southeast orogenic front of the Sierra de Aconquija describesa regional rectilinear feature known as the Tucuman Lineament(Mon, 1976) or the Tucumán Transfer Zone (De Urreiztieta et al.,1996). This feature has been interpreted as a dextral strike-slipfault by some authors (De Urreiztieta et al., 1996; Roy et al., 2006).

Further south of the Sierra de Aconquija, the southwestern limitof the Tucumán basin is bounded by the Sierra de Ambato (Fig. 2),a northwest trending range formed by a series of basement thrustblocks dipping to the east (Cristallini et al., 2004; Roy et al., 2006;Toselli et al., 1999).

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474 461

The eastern limit of the Tucuman basin is the Sierra de Guasayán(Fig. 2), which is a small NeS range, with elevations lower than700 m and is uplifted by high angle thrust faults (Cristallini et al.,2004). The Guasayán thrust fault places basement rocks overNeogene layers. According to Cristallini et al. (2004), the MujerMuerta or the Tacanas high is the prolongation of the structureof the Sierra de Guasayán to the north (Fig. 2). As the Sierra deGuasayán started to uplift probably in Pleistocene times, theTucumán basin evolved into a piggy back basin on top of the activeGuasayán thrust fault (Cristallini et al., 2004; Drozdzewski andMon, 1999; Pacheco et al., 2000).

3. Basin description

3.1. Methodology

Forty-four 2D seismic sections amounting to more than 730 kmwere interpreted and their main reflectors were correlated. Theseseismic sections were originally acquired by the Argentine nationaloil companyYPFbetween1989and1991, thenprocessedas stackbutnot migrated. The geological interpretation of the sections wasperformed using Kingdom Suite software. These seismic sectionswere correlated in a 2D environment and the horizons were

Fig. 3. Stratigraphic column and Seismostra

matched. Eight horizons were correlated to reconstruct the geom-etry of the Tucumán basin. The vertical scale of the seismic sectionsis presented in timedomains in two-way time travel (seconds units).The main horizons and structures are outlined in the seismicsections with the prominent seismostratigraphic units (Figs. 3e12).

Subsequently, the seismic interpretations were converted fromtime to depth domains. To this end, the velocity values obtained byCristallini et al. (2004) were used. These authors obtained seismicvelocities from a well drilled in the margin of the basin, Isca Yacux�1. The fact that there were no other wells in the basin obliged usto use data from other regions to compare and adjust the velocityvalues. The Metán sub-basin wells (Cristallini et al., 1997) and theLomas de Olmedo sub-basin structural maps in depth domainswere used as constraints (Disalvo et al., 2002; Masaferro et al.,2003). Thereafter, the basement horizon in depth domains wasused to construct a structural map.

3.2. Seismostratigraphic units

The lateral and vertical variations of the reflection patternsthroughout the seismic sections of the Tucumán basin enabled usto identify four superimposed reflective packages. The lowermostpackage one corresponds to the acoustic basement. It is composed

tigraphic units of the Tucumán basin.

Fig. 4. Seismic section 2542. Location in Fig. 2.

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474462

of low intensity reflections with poor lateral continuity giving riseto a homogeneous and chaotic seismic fàcies. Associated strong anddipping reflections have been interpreted by Cristallini et al. (2004)as basement fabric discontinuities (Fig. 4).

Overlying the basement, syn-rift deposits forming part of thePirgua subgroup can be identified by marked lateral thicknessvariations and fan geometries (Fig. 5). The contact between thispackage and the acoustic basement top occurs through an onlaprelation. The syn-rift deposits are located in half-graben structures,increasing layer thickness towards the faults (Fig. 5).

After the cessation of extension, the geometry of the rift basinchanged as the subsidence and the depositional rate diminished.Post-rift layers were set during thermal subsidence, when theactivity of normal faults became less intense and the post-rift basinacquired a sag geometry. The post-rift sequence shows good lateralcontinuity and intensity in their seismic reflectiveness despite the

existence of some intercalated low amplitude reflectors. An initialpost-rift package onlaps the syn-rift sediments and the basement,and constitutes the Balbuena subgroup (Fig. 3). An upper post-riftpackage, which corresponds to the Santa Bárbara subgroup, isconcordantly located above the initial post-rift package in thecenter of the Tucumán basin but onlaps the basin borders (Fig. 3).

Foreland deposits show more extensive and lateral continuoushorizons with high amplitudes. The lower contact is approximatelyparaconcordant over the post-rift facies. Three episodes may bedistinguished during this stage: Foreland basin I, Foreland basin IIand Foreland basin III. Foreland basin I is constituted by large,continuous reflective seams of different intensities associated withthe Andean uplifting that occurred at a considerable distance to thewest. This sequence coarsened up as the source area approachedthe depositional area. Intercalated in Foreland basin I are strongreflective horizons, which are interpreted as marine deposits of the

Fig. 5. Seismic section 2523. Location in Fig. 2.

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474 463

Atlantic transgressive event that covered the whole basin. Forelandbasin II is related to the start of the tectonic inversion of earlierextensional faults in the area and layers of this packagewere foldedand thrusted. Finally, Foreland basin III results from the mainuplifting of the surrounding ranges. During this stage, tectonicinversion controlled the depositional space and produces erosionalunconformities and growth strata.

3.3. Interpretation of seismic sections

Nine of the forty-four seismic sections of the different sectors ofthe Tucumán Basin will be discussed below in an attempt to char-acterize the shape, structure and evolution of the basin (Fig. 2).

The most suitable section that illustrates the characteristics ofthe seismostratigraphic units of the basin is seismic section 2542

Fig. 6. Seismic section 1559. Location in Fig. 2.

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474464

Fig. 7. Seismic section 1567. Location in Fig. 2.

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474 465

(Fig. 4). This NeS cross-section, which ends close to the town ofSimoca, is parallel to the longest axis of the basin and shows thethickest package of syn-rift, post-rift and foreland deposits. Syn-riftdeposits were controlled by a large graben structure with a highangle normal fault and antithetic faults. Syn-rift layers grow inthickness against the southern fault and thin out to the north. Post-rift layers onlap syn-rift deposits and have lensoidal geometries,which enabled us to interpret a more subsident area in the center ofthe basin during the deposition. At the base of the post-riftsequence are transparent bands that correspond to mud and salt.These transparent bands are covered by strong reflective seams in

the middle and upper parts of the top post-rift sequence. Forelandstage tabular packages were deposited over post-rift units and havea different reflectiveness with respect to the lower units. Forelandbasin I is represented by thick, mainly parallel and tabular units.Lateral continuity is less clear with respect to the post-rift stageexcept for a clear reflective seam located at 1.4 s. This seam could beinterpreted as marine deposits of the Paraná transgressive event.Foreland basin II shows an increase in thickness of up to 0.8 s to thesouth. This could be due to growing accommodational space causedby folding of the older units in a gentle anticline located in thenorthern part of the section. This folding is attributed to gentle

Fig. 8. Seismic section 2503. Location in Fig. 2.

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474466

inversion of the northernmost extensional faults. The upper units,which cover the previous foreland stages, constitute Foreland basinIII. The geometry of these units suggests a break in the localuplifting.

The eastern margin of the Tucumán basin is exemplified byseismic section 2523 (Fig. 5), located northeast of Simoca (Fig. 2). Inthis sector the basement shows strong reflective seams dipping tothe east. The contact between the basement top and the base of thesedimentary fill is located at 2 s in the east and 2.6 in the west. Syn-rift deposits can be recognized by their strong lateral thicknessvariations controlled by normal faults dipping to the west. At least3 half-graben structures are prominent, conditioned by high anglenormal faults. Hangingwall syn-rift layers increase in thicknessagainst the fault. The post-rift deposits are characterized byreflective seams of good lateral continuity and a more intense

reflectiveness at the top. Post-rift deposits can be identified by theironlapping relation over the syn-rift deposits towards the east(Cristallini et al., 2004). The Santa Bárbara layers increase inthickness westwards. The foreland deposits gently dip to the westin a paraconcordant relation over the post-rift deposits. The recentmost deposits of Foreland basin III can be recognized by their onlapto the east that is related to the tectonic uplifting of the east marginof the basin and by a greater thickness in the center of the basin tothe west.

Further north of seismic section 2523 and to the east of the townof Tucumán (Fig. 2) is seismic section 1559 (Fig. 6). In this sector, thebasement top is closer to the surface and the entire sedimentarysequence is thinner. The basement is homogeneous without strongreflections, but chaotic and without continuous reflective seams.The thickness of the syn-rift and post-rift deposits decreases with

Fig. 9. Seismic section 2501. Location in Fig. 2.

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474 467

respect to that of the center of the basin. The Pirgüa deposits aredifficult to be recognized in this section. Basal post-rift layers of theBalbuena subgroup are thinner but with a strong reflectiveness. TheSanta Bárbara layers have a constant thickness. Nearly the wholesequence is folded in an anticline due to the inversion of the mainfault that controlled the syn-rift deposits and the formation ofa footwall short cut. Small and gentle anticlines and synclines affectthe Balbuena deposits in the hangingwall of the main fault and areprobably due to the inversion of the antithetic faults. The depositsof Foreland basin III were sedimented synchronously with tectonicinversion as evidenced by growth strata.

Seismic section 1567 (Fig. 7), which is in the northeast of thestudy area, has an EeWorientation and is located at the foot of theSierra de la Ramada (Fig. 2). This seismic section shows a hanging-wall anticline that is associated with a high angle fault. A shortcutthrust is recognized in the footwall. An antithetic extensional faultdipping to the west was also inverted with minor displacements asdemonstrated by the gentle anticline present in the middle ofa synclinorium. Tectonic inversion along this section was strongerthan in the previous sections. Inversion affected the sequence up tothe layers of Foreland basin I. These layers were folded and trun-cated by erosion before toplap sedimentation of the layers ofForeland basin II. The geometry of uppermost Pliocene to Quater-nary layers (Foreland basin II and III) reveal that tectonic inversionwas slightly reactivated during these periods. The inverted fault isa blind thrust that does not reach the surface.

The structure of the western margin of the Tucumán basin isstrongly influenced by the uplifting of the Sierra de Aconquija and

the Sierra de Ambato. Seismic section 2503 (Fig. 8) is orientedWeEin the western half and NWeSE on its eastern side. This section islocated in the foothills of the Sierra de Aconquija and to the south ofthe town of Concepción (Fig. 2). This seismic section shows tiltedlayers dipping to the southeast, with an anticline syncline pair. Inthe northwest, three high angle faults can be identified. They wereextensional faults during rifting as evidenced by the greater syn-riftsediment thickness in the hangingwall than in the footwall. Two ofthese faults (between shot points 1100 and 900) have been invertedas shown by repetition and folding of the syn-rift, post-rift andForeland basin I layers. The fault located further west has not beeninverted and shows the original extensional geometry. To the east,an antithetic fault dipping to the northwest has not undergoneinversion. Pliocene to Holocene layers (Foreland basins II and III)show growth strata, indicating that the sedimentation of theselayers occurred during inversion and folding.

Seismic section 2501(Fig. 9) is located further south, close to thetownofVillaAlberdi. Themain feature in this section is ahalf-grabenstructure controlled by an extensional fault with a thick syn-riftpackage in the hanging wall. The extensional fault continued to beslightly active during the sedimentation of the lowest post-rift unit(the Balbuena subgroup). This fault, at its highest point, branchesinto two thrust faults that have an associated anticline involving thesequence up to the layers of Foreland basin II. The kinematicconnection of the thrust faults and the extensional fault suggest thatthe latter has been slightly inverted. Towards the east, the layers ofForeland basin III onlap the limb of the anticline that was generatedby the tectonic inversion of the main fault.

Fig. 10. Seismic section 2570. Location in Fig. 2.

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474468

Fig. 11. Seismic section 2491. Location in Fig. 2.

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474 469

In the southern sector of the Tucumán basin, seismic sectionsare more spaced. Seismic section 2570 (Fig. 10) is located in thesouthern part of the basin, with an NNWeSSE orientation. This longseismic section describes a series of half-graben structurescontrolled by extensional faults dipping to the south. Syn-riftdeposits increase in thickness towards the southern extensionaldepocenters. The post-rift and foreland layers are tilted and grow inthickness to the north, i.e. towards the center of the Tucumán basin.Foreland sequences are thinner than in the previous seismicsections. Folding affected the whole stratigraphic sequence. Smallasymmetric anticlines with a southern limb with a greater dip are

located over the main extensional faults in the southern part(Fig. 10). Their origin will be discussed below.

To the east and south, close to the Sierra de Guasayán and to thesouth of the Río Hondo dam is seismic section 2491 (Fig. 11). Thisseismic section is located in the southeastern margin of the basin.The basement shows a series of strong reflective seams dipping tothe east. Almost no syn-rift and few post-rift layers were identifiedin this section. The upper post-rift units (Santa Bárbara subgroup)thicken to the west as do the foreland layers. An anticline can beobserved in the middle of the section associated with a high angleblind thrust. This basement thrust fault constitutes a backthrust of

Fig. 12. Seismic section 2467. Location in Fig. 2.

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474470

the Sierra de Guasayán (Fig. 2) and does not control the thickness ofthe syn-rift layers. The anticline involves the whole sedimentarysequence as far as the layers of Foreland basin II. Moreover, its westlimb is onlapped by the layers of Foreland basin III.

The eastern border of the Tucumán basin is bounded by a highangle thrust fault that is controlled by a deep detachment. Seismicsection 2467 (Fig. 12) crosses an area that underwent no exten-sional activity during the Cretaceous as evidenced by the uniformthickness of the syn-rift deposits. These deposits are much thinnerthan in other sections and represent distal positions with respect tothe extensional depocenter. The layers of Foreland basin I increaseslightly in thickness towards the east. Between shot points 700 and500 is a pop-up that corresponds to the Mujer Muerta high. Thewestern thrust fault of this structure produced a fault propagationfold, which folded the stratigraphic sequence as far as the layers ofForeland basin II. The layers of Foreland basin III onlap the afore-mentioned fault propagation fold. This structure is similar to theone described to the south in seismic section 2491 (Fig. 11). Theeastern thrust fault of the pop-up produces a small displacement in

the whole stratigraphic sequence and reaches the surface. The twothrust faults that bound the Mujer Muerta High are active. Close tothe eastern end of the section are two conjugate normal faults witha small displacement and associated folds.

4. Structural analysis of the Tucumán basin

The basin has a triangular shape in plan view, the eastern limit isNeS, the northwestern limit is NEeSW and the southwestern limitis NNWeSSE (Fig. 13). The basin is asymmetric with a steeperwestern margin and a gentler eastern slope (Porto et al., 1982). Thenorthern sector of the basin is shallower and narrower than thesouthern sector. The geometry of the basin was controlled byextensional faulting and was subsequently shaped by tectonicinversion and thrusting. The main depocenter of the basin wastermed the Leales depocenter by Pacheco et al. (2000) and coin-cides with a Cretaceous graben (Fig. 4) striking NEeSW (Fig. 13). Tothe north, and separated by a small structural high is another smalldepocenter with syn-rift deposits (Fig. 13), which may be termed

Fig. 13. Structural map of the basement top in depth domains and of interpreted subsurface structures. Gray lines correspond to seismic profiles. Basement and Cretaceous syn-rift(Pirgua subgroup) outcrops, and regional faults visible in the surface are also represented.

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474 471

the Famaillá depocenter because of its proximity to this town(Fig. 2).

Cretaceous extensional structures were also described in otherparts of the Tucumán basin. The Villa AlberdieConcepción graben(Pomposiello et al., 1993) (Figs. 2, 8 and 9 and 13) was controlled bya northeast dipping fault in the north and a southwest dipping faultin the south. Another large accumulation of syn-rift deposits islocated to the northeast of the basin (Figs. 2, 6, 7 and 13) in a grabenthat could be the subsurface continuation of the inverted graben thatcrops out in the Sierra de La Ramada or in the Sierra de Medina. Tothe south, a number of half-grabens were identified (Figs. 2 and 10).

The grabens and their normal faults must be oriented EeW sincethey are not recognized in the orthogonal seismic sections to the east(Fig. 11) and to the west.

Folding of the sedimentary cover occurred as a result of differentmechanisms. Some folds have roll-over geometries and are locatedin the hangingwall of Cretaceous extensional faults (Figs. 5 and 9).Other hangingwall anticlines are fault propagation folds producedby inversion of Cretaceous extensional faults (Figs. 6, 7 and 8) or byhigh angle thrust faults (Figs. 11 and 12). By contrast, some anti-clines are located over the extensional faults (Fig. 10) and maybe due to one of three mechanisms: a) tectonic inversion,

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474472

b) extensional reactivation or c) differential subsidence due tolateral change in thickness of the syn-rift layers and to their unevencompactation below the load of the sedimentary column (Cristalliniet al., 2006). The last mechanism is the most probable owing to theposition of the syncline trough over themaximun syn-rift thickness.Differential subsidence is responsible for folding the post-rift andthe foreland layers over the syn-rift deposits and the rigid basement(Cristallini et al., 2006).

5. Discussion

The study area has undergone a number of stages of deformationwithin different tectonic settings. The absence of Paleozoic rocks inthe basin suggests that the area was a structural high during thisperiod. In Cretaceous times, the Salta rifting produced half grabensand a sedimentary infill related sequence (Porto et al., 1982; Salfityand Marquillas, 1981). The syn-rift Pirgüa subgroup increases inthickness towards the Cretaceous extensional faults and thins outfrom these faults. The layers of the Pirgüa subgroup onlap thestructural highs but become progressively thinner towards themargins of the basin, petering out to the southeast (Fig. 11).

The post-rift sag phase, which is due to the cooling of thethermal anomaly associated with the Salta rifting, created accom-modational space by slow subsidence. Post-rift deposits of the Saltagroup are thinner towards the basin margins where the basementtop is closer to the surface. The maximum thickness of the post-riftdeposits coincides with the syn-rift depocenters in the center of thebasin (Fig. 4).The onlap of the post-rift deposits on the syn-riftdeposits provides evidence of the arenal growth of the post-riftbasin and of the location of areas of greater subsidence (Figs. 4,5 and 10). A foreland basin developed over the post-rift basin ina paraconcordant stratigraphic relationship. A hiatus in the strati-graphic sequence has been reported (Reynolds et al., 2000; DelPapa et al., 2010). The subsidence mechanisms changed in theforeland basin stage. Sedimentation space increased owing to thetectonic load as the orogen grew in the west (Carrapa et al., 2006;Carrapa and DeCelles, 2008; Jordan et al., 1983; Jordan and Alonso,1987). The foreland basin associated with the Andean orogenystarted as a regional single unit and then compartmentalized intosub-basins coeval with the uplifting of the Pampean Ranges and theSanta Bárbara system (Carrapa et al., 2005; Coughlin et al., 1998).The deposits of Foreland basin I are tabular layers covering thepost-rift units (Fig. 4). This sequence, which is termed Forelandbasin I, was identified by the onlap of these layers on the basinmargins. Foreland basin I started to develop during the earlyuplifting of the eastern margin of the Puna plateau in the middleEocene (Carrapa et al., 2005; Isacks, 1988; Jordan and Alonso,1987).Foreland basin II resulted from the uplifting of the surroundingranges, and is associated with growth strata geometries (Figs. 7and 8). During this stage, the foreland basin started to be frag-mented into different sub-basins. Subsidence became local withranges uplifting and controlling the sedimentary supply. Short-ening produced tectonic inversion of the Cretaceous extensionalfaults, folding the sedimentary cover. As pointed out by Cristalliniet al. (2004), during sedimentation of the layers of Foreland basinIII, the Tucumán basin was separated from the rest of the forelandbasin by uplifting of the Sierra de Guasayán and the India MuertaHigh, which has caused the Tucumán basin to evolve into a piggyback basin. The western margin uplift generates more subsidencethrough tectonic load (Fig. 8).

According to the age of growth strata, some structures can beconsidered as active or potentially active whereas others cannot.Neotectonic activity has been recorded along both margins of theSierra de Aconquija (Cristallini et al., 2004; Drozdzewski and Mon,1999). Active seismicity has been documented together with surface

evidence of earthquakes and rock avalanches on the western slope(FauquéandStrecker,1987). In seismic section2503 (Fig. 8), thewholesequence is involved in a thrust fault related fold. Growth strataprovide evidence of folding during sedimentation of the layers ofForeland basins II and III. Given that the Sierra deAconquija started touplift 6 Ma ago (Sobel and Strecker, 2003), this age could be assignedto the basal layers of Foreland basin II. Seismic section 2501 (Fig. 9)also shows growth strata in the layers of Foreland basin III but, in thiscase, onlap is in the opposite sense in seismic section 2503 (Fig. 8).

The Sierra de Guasayán and theMujerMuerta High are a pop-upuplifted by a double vergent high angle thrust fault system (Figs. 11and 12), which is similar to the Sierra de Aconquija and the Cum-bres Calchaquíes (Drozdzewski andMon,1999). The double vergentstructure visible in seismic section 2467 (Fig. 12) folded the wholestratigraphic sequence with topographic and subsurface evidenceof recent uplifting.

The Sierra de Medina provides surface evidence of recentinversion of a Cretaceous extensional fault as a thrust fault thatelevates the range in its southern margin (Fig. 2). Nearby, seismicsections show different geometries such as in seismic section 1567(Fig. 7), which has an erosion surface that affects the layers ofForeland basin I. This inversion structure ceased to uplift before thesedimentation of Foreland basin II. A slight increase in layerthickness in the west and folding of the layers of Foreland basins IIand III suggest that the inverted fault could have been reactivatedrecently. Moreover, in seismic section 1559 (Fig. 6), growth stratafurnish evidence that the inverted fault continues to be active.

In the center of the Tucumán Basin, tectonic inversion is not asclear as along the margins. In seismic section 2542 (Fig. 4), thelayers of Foreland basin III cover the whole sequence and presentno folding nor growth strata. This indicates that these faults haveceased to be active.

Strike-slip effects are not easy to identify using 2D seismicsections. De Urreiztieta et al. (1996) described a dextral fault slip ofthe Tucumán transfer zone, but strike-slip structures such astransgressive or transtensive structures have not been documented(Cristallini et al., 2004).

The depth of the Tucumán basin cannot be constrained becauseof the lack of wells across the entire column. The maximum depthof the basement top is located in themiddle of seismic section 2542(Fig. 4) and is about 3.4 s in TWT. The velocity values of Cristalliniet al. (2004) were used to perform a time to depth conversion.Our results reveal that the basin should be about 7000 m deep.These values are comparatively higher than other estimations.Pomposiello et al. (2002) using gravimetry and Favetto et al. (2007)using magnetothelluric obtained depths of 4000e5000 m in thecentral part of the basin. The results from neighbouring basins ofthe Salta rift such as the Choromoro Basin gave a depth of 3000 m(Abascal, 2005).

6. Conclusions

Basement, syn-rift, post-rift and foreland strata may be inter-preted and differentiated in the seismic sections in accordancewiththe geometry of their horizons. The basement discontinuities con-ditioned the structural style and exerted a strong influence on theselective orientation of Cretaceous extensional faults and on that ofblocks uplifted in Neogene times. Inversion of the Cretaceousextensional faults occurred along the northern andwesternmarginsof the Tucumán basin. In the eastern margin it is not easy to detecttectonic inversion despite the fact that it could be interpreted thatsmall extensional steps are slightly inverted. The southern part ofthe basin preserved the Cretaceous extensional geometries withoutbeing inverted. Inversion is clear along the southwestern andnorthwestern margins of the Tucumán basin. Moreover, inversion

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474 473

structures are clearly observed at the surface to the north of theTucuman basin, i.e. in theMedina Range syn-rift deposits are thrustover post-rift and foreland layers.

Folding was produced by reactivation of basement discontinu-ities by means of high angle thrust faults, inversion of earlierextensional faults and by differential subsidence owing to thestratigraphic column load on syn-rift sediments that varied later-ally in thickness. Double vergent high angle thrust faults in theeastern margin of the basin continued to uplift the Guasayán Rangeand the Mujer Muerta High, evolving into a pop-up and initiatingthe transformation of the Tucuman basin into a piggy back basin.The growth of pop-ups in the foreland is a long term process thatproduces the compartmentalization of an early continuous forelandbasin.

Growth strata associatedwith uplifting of the nearby ranges andwith blind thrusts were identified in many sectors of the basin.These features were crucial in identifying the evolution of the basinsedimentary infill and the age and mechanisms of range uplift.

Acknowledgements

This research is supported by the following projects: 2005-00397SGR from the Generalitat de Catalunya and Consolider-Ingenio2010 program (CSD2006-004 “Topo-Iberia”) and CGL2007-66431-C02-01/BTE (Modelización Estructural 4D) from the Ministerio deEducación y Ciencia of the Spanish government. The first authorwas funded by the AlBan scholarships program. We would like tothank Yanina Basile and Tomas Zapata from Repsol-YPF as well asErnesto Cristallini of Universidad de Buenos Aires for facilitatingthe seismic information, Mireia Butille and Joana Mencos from theUniversity of Barcelona for their help in software applications andGeorge von Knorring for reviewing the English of this paper. Weare indebted to Fernando Hong and to an anonymous reviewer fortheir helpful comments on an earlier version that greatly improvedthis manuscript.

The seismic interpretation used The Kingdom Company soft-ware, whichwas generously provided by SeismicMicro-Technologyvia the University Gift Program.

References

Abascal, L., del, V., 2005. Combined thin-skinned and thick-skinned deformation inthe central Andean foreland of northwestern Argentina. Journal of SouthAmerican Earth Sciences 19, 75e81.

Aceñolaza, F.G., Miller, H., Toselli, A.J., 2002. ProterozoiceEarly Paleozoic evolutionin western South America. A discussion. Tectonophysics 354, 121e137.

Allmendinger, R.W., Ramos, V.A., Jordan, T.E., Palma, M., Isacks, B.L., 1983.Paleogeography and Andean structural geometry, northwest Argentina.Tectonics 2, 1e16.

Allmendinger, R.W., Isacks, B.L., Jordan, T.E., Kay, S.M., 1997. The evolution of theAltiplanoePuna plateau of the Central Andes. Annual Reviews of Earth andPlanetary Sciences 25, 139e174.

Battaglia, A.C., 1982. Descripción geológica de las Hojas 13f, Río Hondo, 13g, San-tiago del Estero, 14g, El Alto, 14h, Villa San Martín, 15g, Frías. Provincias deSantiago del Estero, Catamarca y Tucumán. Servicio Geológico Nacional, BuenosAires, Argentina 569 (186), 80.

Bianucci, H.A., Acevedo, O.M., Cerdán, J.J., 1981. Evolución tectosedimentaria delGrupo Salta en la subcuenca Lomas de Olmedo (Provincias de Salta y Formosa):VIII Congreso Geológico Argentino (San Luis). Actas 3, 159e172.

Bianucci, H., Momovc, J.F., Acevedo, O.M., 1982. Inversion Tectónica y Plegamientosresultantes en la Comarca puesto Guardian-Dos Puntitas. Depto. Orán, Provinciade Salta. Primer Congreso Nacional de Hidrocarburos, Buenos Aires. 23e30.

Boll, A., Gómez Omil, R.J., Hernández, R.M., 1989. Síntesis estratigráfica del GrupoSalta. Gerencia general de Exploración Y.P.F. (inedit).

Bonaparte, J., Salfity, J., Bossi, G., Powell, J., 1977. Hallazgo de dinosaurios y avescretácicas en la Formación Lecho de El Brete (Salta), próximo al límite conTucumán. Acta Geológica Lilloana 14, 5e17.

Bossi, G.E., 1969. Geología y estratigrafía del sector sur del Valle de Choromoro. ActaGeológica Lilloana 10, 17e64.

Bossi, G.E., Gavriloff, I.J., 1998. Terciario. Estratigrafía, Bioestratigrafía y Paleo-geografía. In: Gianfrancisco, M., Puchulu, M.E., Durango de Cabrera, J.,

Aceñolaza, G.F. (Eds.), Geología de Tucumán (2a ed.), Publicación Especial,Colegio de Graduados en Ciencias Geológicas de Tucumán, pp. 87e109.

Bossi, G.E., Muruaga, C., Georgieff, S., Ahumada, A.L., Ibañez, L., Vides, M.E., 1997. TheSanta María-Hualfín Neogene Basin of the Pampean Ranges: an example ofmixed tectonic evolution. 1� Congreso Latinoamericano de Sedimentología I,97e104.

Bossi, G.E., Georgieff, S., Gavriloff, I., Ibáñez, L., Muruaga, C., 2001. Cenozoic evolu-tion of the intramontane Santa María basin, Pampean ranges, northwesternArgentina. Journal of South American Earth Sciences 14, 725e734.

Carrapa, B., Adelmann, D., Hilley, G., Mortimer, E., Strecker, M.R., Sobel, E.R., 2005.Oligocene uplift, establishment of internal drainage and development ofplateau morphology in the southern Central Andes. Tectonics 24, TC4011.

Carrapa, B., DeCelles, P.G., 2008. Eocene exhumation and basin development in thePuna of northwestern Argentina. Tectonics 27, TC1015.

Carrapa, B., Sobel, E., Strecker, M.R., 2006. Cenozoic orogenic growth in the CentralAndes: evidence from rock provenance and apatite fission track thermochro-nology along the southernmost Puna Plateau margin (NWArgentina). Earth andPlanetary Science Letters 247, 82e100.

Carrera, N., Muñoz, J.A., Sàbat, F., Mon, R., Roca, E., 2006. The role of inversiontectonics in the structure of the Cordillera oriental (NW Argentinean Andes).Journal of Structural Geology 28, 1921e1932.

Carrera, N., Muñoz, J.A., 2008. Thrusting evolution in the southern CordilleraOriental (Northern Argentine Andes): constraints from growth strata. Tecto-nophysics 459, 107e122.

Comínguez, A.H., Ramos, V.A., 1995. Geometry and seismic expresión of theCretaceous Salta rift of northwestern Argentina. In: Tankard, A.J., Suarez, R.,Welsink, H.J. (Eds.), Petroleum Basins of South America. Memoir, vol. 62.American Association of Petroleum Geologists, pp. 325e340.

Costa, C.H., Giaccardi, A.D., Gonzalez Díaz, E.F., 1999. Palaeolandsurfaces and neo-tectonic analysis in the southern Sierras Pampeanas, Argentina. In: Smith, B.J.,Whalley, W.B., Warke, P.A. (Eds.), En: uplift, erosion and stability; perspectiveson longterm landscape development. Geological Society Special Publications,162, pp. 229e238.

Coughlin, T.J., O’Sullivan, P.B., Kohn, B.P., Holcombe, R.J., 1998. Apatite fission-trackthermochronology of the Sierras Pampeanas, central western Argentina;implications for the mechanism of plateau uplift in the Andes. Geology 26 (11),999e1002.

Coutand, I., Cobbold, P.R., de Urreiztieta, M.de, Gautier, P., Chauvin., A., Gapais, D.,Rossello, E.A., Lopez Gamundi,., O., 2001. Style and history of the Andeandeformation, Puna Plateau, northwestern Argentina. Tectonics 20, 210e234.

Cristallini, E., Bottesi, G., Gavarrino, A., Rodríguez, L., Yomezzoli, R., Cameron, R.,2006. Syn-rift geometry of the Neuquén basin in northeastern Neuquénprovince, Argentina. In: Kay, S.M., Ramos, V.A. (Eds.), Evolution of anAndean Margin: a Tectonic and Magmatic View from the Andes to theNeuquén Basin (35�e39�S Lat), vol. 407. Geological Society of AmericaSpecial Paper, pp. 147e161.

Cristallini, E.O., Comínguez, A., Ramos, V.A., 1997. Deep structure of the Metá-neGuachipas region: tectonic inversion in Norhwestern Argentina. Journal ofSouth American Earth Sciences 10, 403e421.

Cristallini, E.O., Comínguez, A., Ramos, V.A., Mercerat, E.D., 2004. Basement double-wedge thrusting in the northern Sierras Pampeanas of Argentina (27�S) e

constraints from deep seismic reflection. In: McClay, K.R. (Ed.), Thrust Tectonicsand Hydrocarbon Systems. Memoir, vol. 82. AAPG, pp. 1e26.

De Urreiztieta, M., Gapais, G., Le Corre, C., Cobbold, P.R., Rosello, E.A., 1996. Cenozoicdextral transpression and basin development at the southern edge of theAltiplanoePuna, northwestern Argentina. Tectonophysics 254, 17e39.

Del Papa, C., Kirschbaum, A., Powell, J., Brod, A., Hongn, F., Pimentel, M., 2010.Sedimentological, geochemical and paleontological insights applied to conti-nental omission surfaces: a new approach for reconstructing Eocene forelandbasin in NW Argentina. Journal of South American Earth Sciences 29 (2),327e345.

Disalvo, A., Rodríguez Schelotto, M.L., Gómez Omil, R., Jhoffman, C., Benítez, J.,Hurtado, S., 2002. Los reservorios de la formación Yacoraite. V Congreso deexploración y desarrollo de hidrocarburos, Actas, 721e724.

Drozdzewski, G., Mon, R., 1999. Oppositely-verging thrusting structures in the northArgentine Andes compared with the German Variscides. Acta Geológica His-pánica 34 (2e3), 185e196.

Fauqué, L.E., Strecker, M.R., 1987. Rasgos de neotectónica y avalanchas de rocas pro-ducidas por terremotos en la vertiente occidental de los nevados del Aconquija,Provincia de Catamarca, Argentina. In: Aceñolaza, F., Bossi, G., Toselli, A. (Eds.),Décimo Congreso Geológico Argentino. Actas, vol. I, pp. 219e222.

Favetto, A., Pomposiello, M.C., Booker, J., Rossello, E., 2007. Magnetotelluric inver-sion constrained by seismic data in the Tucumán basin (Andean foothills, 27�S,NW Argentina). Journal of Geophysical Research, 112.

Fernández Garrasino, C., Laffitte, G., Villar, H., 2005. Cuenca Chacoparanense. In:Chebli, G.A., Cortiñas, J.S., Spalletti, L.A., Legarretta, L., Vallejo, E.L. (Eds.), Fron-tera Exploratoria de la Argentina. 6� Congreso de Exploración y Desarrollo deHidrocarburos, pp. 97e114.

Galliski, M.A., Viramonte, J.G., 1988. The Cretaceous paleorift in northwesternArgentina: a petrologic approach. Journal of South American Earth Sciences1, 329e342.

Gavriloff, I.J.C., Bossi, G.E., 1992. Revisión general, análisis facial, correlación y edadde las Formaciones San José y Río Salí (Mioceno medio), provincias de Cata-marca, Tucumán y Salta, República Argentina. Acta Geológica Lilloana 17 (2),5e43.

D.N. Iaffa et al. / Journal of South American Earth Sciences 31 (2011) 457e474474

González, O.E., 2000. Hoja Geológica 2766-II San Miguel de Tucumán. SEGEMAR,Buenos Aires, 124 pp.

González Bonorino, F., 1950. Algunos problemas geológicos de las Sierras Pam-peanas. Revista de la Asociación Geológica Argentina 5 (3), 81e110.

Grier, M.E., Salfity, J.A., Allmendinger, R.W., 1991. Andean reactivation of theCretaceous Salta rift, northwestern Argentina. Journal of South American EarthSciences 4 (4), 351e372.

Gutiérrez, A.A., Mon, R., 2008. Macro-indicadores cinemáticas en el bloque Ambato,Provincias de Tucumán y Catamarca. Revista de la Asociación GeológicaArgentina 63 (1), 24e28.

Gómez Omil, R.J., Boll, A., Hernández, R.M., 1989. Cuenca cretácico-terciaria delNoroeste argentino (Grupo Salta). In: Chebli, G.A., Spalleti, L.A. (Eds.), CuencasSedimentárias Argentinas. Serie Correlación Geológica. Universidad Nacional deTucumán, San Miguel de Tucumán, pp. 43e64.

Hilley, G.E., Coutand, I., 2010. Links between topography, erosion, rheologicalheterogeneity and deformation in cotractional settings: Insights from thecentral Andes. Tectonophysics 95, 78e92.

Hindle, D., Kley, J., Oncken, O., Sobolev, S., 2005. Crustal balance and crustal fluxfrom shortening estimates in the Central Andes. Earth and Planetary ScienceLetters 230, 113e124.

Hongn, F., del Papa, C., Powell, J., Petrinovic, I., Mon, R., Deraco, V., 2007. MiddleEocene deformation and sedimentation in the Puna-Eastern Cordillera transi-tion (23�e26� S): control by preexisting heterogeneities on the pattern of initialAndean shortening. Geology 35, 271e274.

Iaffa, D.N., Sábat, F., Ferrer, O., Mon, R., Gutiérrez A.A., 2008. Incipient tectonicinversion in a segmented foreland basin: From extensional to piggy-backsettings. 7th International Symposium on Andean Geodynamics, ISAG 2008,Nice, Extended Abstracts, 261e264.

Isacks, B.L., 1988. Uplift of the Central Andean plateau and Bending of the BolivianOrocline. Journal of Geophysical Research 93 (B4), 3211e3231.

Jordan, T.E., Alonso, R.N., 1987. Cenozoic stratigraphy and basin tectonics of theAndes Mountains, 20�e28� south latitude. The American Association ofPetroleum Geologists Bulletin 71 (1), 49e64.

Jordan, T.E., Isacks, B., Ramos, V.A., Allmendinger, R.W., 1983. Mountain building inthe Central Andes. Episodes 3, 20e26.

Kleinert, K., Strecker, M.R., 2001. Climate change in response to orographic barrieruplift: paleosol and stable isotope evidence from the late Neogene Santa Maríabasin, northwestern Argentina. Geological Society of America Bulletin 113,728e742.

Kley, J., Monaldi, C.R., Salfity, J.A., 1999. Along-strike segmentation of the Andeanforeland: causes and consequences. Tectonophysics 301, 75e94.

Kley, J., Monaldi, C.R., 2002. Tectonic inversion in the Santa Bárbara System of thecentral Andean foreland thrust belt, northwestern Argentina. Tectonics 21 (6),1061.

Kley, J., Rosello, E.A., Monaldi, C.R., Habighorst, B., 2005. Seismic and field evidencefor selective inversion of Cretaceous normal faults, Salta rift, northwestArgentina. Tectonophysics 399 (1/4), 155e172.

Mángano, M.G., Buatois, L.A., 2004. Integración de estratigrafía secuencial, sed-imentología e icnología para un análisis cronoestratigráfico del Paleozoicoinferior del noroeste Argentino. Revista de la Asociación Geológica Argentina59, 273e280.

Marquillas, R.A., Salfity, J.A., 1994. Relaciones estratigráficas regionales de la For-macion Yacoraite (Cretácico Superior), norte de la Argentina. Actas 7 CongresoGeológico Chileno 1, 479e483.

Marquillas, R., del Papa, C., Sabino, I., Heredia, J., 2003. Prospección del límite K/T enla cuenca del Noroeste, Argentina. The K/T boundary prospecting in theNorthwest Basin, Argentina. Revista de la Asociación Geológica Argentina58 (2), 271e274.

Masaferro, J.L., Bulnes, M., Poblet, J., Casson, N., 2003. Kinematic evolution andfracture prediction of the Valle Morado structure inferred form 3-D seismicdata, Salta province, northwest Argentina. AAPG Bulletin 87 (7), 1083e1104.

Mon, R., 1976. La tectónica del borde oriental de los Andes en las provincias de Salta,Tucumán y Catamarca, Republica Argentina. Revista Asociación GeológicaArgentina 31, 65e72.

Mon, R., Hogn, F.D., 1991. The structure of Precambrian and Lower Paleozoic base-ment of the Central Andes between 22� and 32� S Lat. Geologische Rundschau80 (3), 745e758.

Mon, R., Suayter, L., 1973. Geología de la sierra de San Javier (provincia de Tucumán,República Argentina). Acta Geológica Lilloana 12, 155e168.

Monaldi, C.R., Kley, J., 1997. Balanced cross sections of the northern Santa Barbarasystem and Sierra de Zapla, Northwestern Argentina. In: VIII: Congreso Geo-logico Chileno. Chile, Antofagasta 180e184.

Moreno, J.A., 1970. Estratigrafía y paleogeografía del Cretácico Superior en la cuencadel Noroeste Argentino, con especial mención de los Subgrupos Balbuena ySanta Bárbara, vol. 24. Revista Asociación Geológica, Argentina. 9e44.

Mortimer, E., Carrapa, B., Coutand, I., Schoenbohm, L., Sobel, E.R., Gomez, J.S.,Strecker, M.R., 2007. Fragmentation of a foreland basin in response to out-of-sequence basement uplifts and structural reactivation: El Cajón-Campo del Are-nal basin: NWArgentina. Geological Society of America Bulletin 119, 637e653.

Pacheco, M.M., Mansilla, N.Y., Mon, R., Sosa, J., Piccioni, J.Y.L., 2000. The TucumánBasin as a part of the Cretaceous Continental Rift of South America. XVIILateinamerikan-Kollokium.

Padula, E., Rolleri, E.O., Mingramm, A.R., Criado Roque, P., Flores, M.A., Baldis, B.A.,1967. Devonian of Argentina. In: International Symposium on the DevonianSystem. Memoir, vol. 2. Alberta Society of Petroleum Geologists, pp. 165e199.

Pomposiello, M.C., Mon, R., Díaz, T., 1993. The gravity field of the Tucumán plain andits implications in the structural geology. Géodynamique 6, 3e8.

Pomposiello, M.C., Favetto, A., Sainato, C., Booker, J., Li, S., 2002. Imaging the sedi-mentary basin of the Tucumán plain in the northern 619 Pampean rangesArgentina. Journal of Applied Geophysics 49, 47e58.

Porto, J., Danieli, C., Ruíz Huidobro, O., 1982. El grupo Salta en la provincia deTucumán, Argentina. 58 Congreso Latinoamericano de Geología (Buenos Aires)4, 253e264.

Ramos, V.A., 1988. Tectonics of the Late Proterozoice Early Paleozoic: a collisionalhistory of southern South America. Episodes 11, 168e174.

Ramos, V.A., 1999. Las provincias geológicas del territorio argentino. In: Caminos, R.(Ed.), Geología Argentina. Anales, vol. 29. Instituto de Geología y RecursosMinerales, pp. 41e96 (3).

Ramos, V.A., Alonso, R.N., 1995. El Mar Paranense en la Provincia de Jujuy. RevistaGeológica de Jujuy 10, 73e80.

Reyes, F.C., Salfity, J.A., 1973. Consideraciones sobre la estratigrafía del Cretácico(subgrupo Pirgüa) del noroeste argentino. V Congreso Geológico Argentino,Carlos Paz (1972). Actas 3, 15e36.

Reynolds, J.H., Galli, C.I., Hernandez, R.M., Idleman, B.D., Kotila, J.M., Hilliard, R.V.,Naeser, C.W., 2000. Middle Miocene tectonic development of the Transition Zone,Salta Province, northwestern Argentina: Magnetic stratigraphy from the MetánSubgroup, SierradeGonzález. Geological SocietyofAmerica Bulletin112,1736e1751.

Roy, R., Cassard, D., Cobbold, P.R., Rossello, E.A., Billa, M., Bailly, L., Lips, A.L.W., 2006.Predictive mapping for copperegold magmatic-hydrothermal systems in NWArgentina: use of a regional-scale GIS, application of an expert-guided data-driven approach, and comparison with results from a continental-scale GIS. OreGeology Reviews 29, 260e286.

Russo, A., Serraiotto, A., 1979. Contribución al conocimiento de la estratigrafía ter-ciaria en el noroeste argentino. VII Congreso Geológico Argentino, Neuquén(1978). Actas 1, 715e730.

Salfity, J.A., Marquillas, R.A., 1981. Las unidades estratigráficas cretácicas del Nortede la Argentina. In: Volkheimer, W.Y., Musacchio, E. (Eds.), Cuencas Sed-imentarias del Jurásico y Cretácico de América del Sur, vol. 1, pp. 303e317.

Salfity, J.A., Marquillas, R.M., 1994. Tectonic and sedimentary evolution of the Cre-taceouseEocene Salta group basin, Argentina. In: Salfity, J.A. (Ed.), CretaceousTectonics of the Andes. Braunschweig/Wiesbaden, Earth Evolution Sciences.Friedr. Vieweg and Sohn, pp. 266e315.

Schmidt, C.J., Astini, R.A., Costa, C.H., Gardini, C.E., Kraemer, P.E., 1995. Cretaceousrifting, alluvial fan sedimentation and Neogene inversion, Soutern SierrasPampeanas, Argentina. In: Tankard, A.J., Suárez, R., Welsink, H.J. (Eds.), Petro-leum Basins of South America. Memoir, vol. 62. AAPG, pp. 341e358.

Sobel, E., Strecker, M.R., 2003. Uplift, exhumation and precipitation: tectonic andclimatic control of late Cenozoic landscape evolution in the northern SierrasPampeanas, Argentina. Basin Research 15 (4), 431e451.

Strecker, M.R., Cerveny, P., Bloom, A.L., Malizia, D., 1989. Late Cenozoic tectonismand landscape development in the foreland of the Andes: Northern SierrasPampeanas (26�e28�S), Argentina. Tectonics 8 (3), 517e534.

Toselli, J.A., Lopez, J.P., Sardi, F.G., 1999. El basamento metamórfico en CumbresCalchaquíes noroccidentales, Aconquija, Ambato y Ancasti: Sierras Pampeanas.In: Relatorio XIV Congreso Geológico Argentino: Geología del NoroesteArgentino, 73e79.

Turic, M., Aramayo Flores, F., Gómez Omil, R., Pombo, R., Sciutto, J., Robles, D.,Caceres, A.,1987. Geología de las cuencas petroleras de laArgentina. In: Evaluaciónde las Formaciones en la Argentina. Schlumberger, Buenos Aires, pp. I/1eI/41.

Turner, J.C.M., 1959. Estratigrafía del cordón de Escaya y de la sierra de Rinconada(Jujuy). Revista de la Asociación Geológica Argentina 13, 15e39.

Uliana, M.A., Biddle, K.T., 1988. MesozoiceCenozoic paleogeographic and geo-dynamic evolution of southern South America. Revista Brasileira de Geociencias18 (2), 72e190.