Sei sulla pagina 1di 17

Silurian–Jurassic Stratigraphy and Basin Evolution of

Northwestern Argentina

Daniel Starck
Consultant
La Plata, Argentina

Abstract

N orthwestern Argentina has undergone a complex tectonosedimentary history. From Silurian through
Late Jurassic time, sedimentary sequences were deposited in intracontinental basins. The beginning of
this stage is linked to a first-order orogenic event, the Ocloyic phase. Above the Ocloyic unconformity, 6 km of
sedimentary rocks were deposited during the Silurian–Jurassic interval. This record consists of two major
unconformity-bounded sequences separated by the Lower Carboniferous Chanic unconformity. The first
sequence, spanning the Silurian–Devonian, contains predominantly marine facies arranged in cycles of
different hierarchies that were deposited in an epeiric flexural basin related to the Ocloyic orogen. The less
intense Chanic orogeny changed the subsidence style and combined flexural and thermal causes. The low
average sea level and variable climatic conditions during the late Paleozoic and early Mesozoic controlled the
environmental characteristics of this second sequence, which is represented by predominantly continental
facies of glacial and arid origin. This stage of intracontinental evolution ended with rupture and fragmentation
of the basins in response to Cretaceous rifting, which marked the Gondwana break-up.
Several of the units deposited during the studied interval produce oil or gas that were generated in the
Upper Devonian marine shales. The complex burial history led to the variable thermal maturity of these source
rocks. One of the most important control factors of commercial accumulations here is source rock availability
because the Devonian rocks were affected by erosive cycles of different ages.

Resumen

E l extremo noroeste de Argentina mantuvo durante el Fanerozoico una compleja historia tectosedimen-
taria. En particular, durante el intervalo temporal que abarca desde el Silúrico al Jurásico albergó
secuencias sedimentarias depositadas en cuencas intra-continentales. El inicio de esta etapa tectosedimentaria
está vinculada a un evento orogénico de primera magnitud acaecido a fines del Ordovícico, la fase diastrófica
Oclóyica. Sobre la discordancia generada por este acontecimiento se depositaron durante el intervalo Silúrico-
Jurásico unos 6 km de rocas sedimentarias ordenadas en dos secuencias mayores separadas por la discor-
dancia Chánica del Carbónico Inferior. La primera secuencia abarca el Silúrico y el Devónico y se caracteriza
por el predominio de facies marinas arregladas en ciclos de distintas jerarquías depositadas en una cuenca
epírica de naturaleza flexural y relacionada al orógeno Oclóyico. La fase diastrófica Chánica, de menor inten-
sidad que la Oclóyica, provocó cambios en el estilo de subsidencia de la región, la que con menor intensidad
parece haber combinado causas flexurales y termales. A su vez el bajo nivel promedio del mar y las variadas
condiciones climáticas presentes durante el Neopaleozoico y Eomesozoico condicionaron las características
ambientales de la segunda secuencia, representada por facies mayoritariamente continentales de origen glacial
y de ambientes áridos.
Varias de las unidades depositadas en el intervalo estudiado son productoras de petróleo o gas, los que han
sido generados a partir de las pelitas marinas del Devónico Superior. La compleja historia de soterramiento
que sufrieron estas rocas provocó su madurez térmica. Uno de los factores más importantes para la ocurrencia
de acumulaciones comerciales es la disponibilidad de roca madre, ya que la distribución de las rocas devónicas
fué afectada por ciclos erosivos de distinta edad.

Starck, D., 1995, Silurian–Jurassic stratigraphy and basin evolution of northwestern 251
Argentina, in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South
America: AAPG Memoir 62, p. 251–267.
252 Starck

INTRODUCTION
From Silurian through Late Jurassic time, north-
western Argentina was subjected to intracratonic tectonic
processes. The stratigraphic record has distinct
Gondwana affinities. This paper addresses this strati-
graphy as a basis for reconstructing the tectonic
evolution of the region.
Northwestern Argentina borders Bolivia and shares a
common geology with it. This region has long been
known as the “Northwestern basin,” ignoring the fact
that it is a complex tectonic province consisting of several
basins each of which preserves a unique record of timing
and style of deformation. Oil and gas fields have been
exploited since 1925. Hydrocarbon source rocks and
reservoir rocks accumulated during the Paleozoic and
Mesozoic. In contrast, maturation, migration, and
formation of structural traps are dated to the Neogene
Andean phase of deformation. This commercial interest
is reflected in several thousand kilometers of seismic
surveys and more than 100 exploration wells. This study
supplements these subsurface data with surface geology.

Figure 1—Silurian–Jurassic outcrop distribution. Localities


GEOLOGIC SETTING mentioned in text: AB, Aguas Blancas well; B, Balapuca;
Bjo, Bermejo well; C, Caspalá; CPR, Cinco Picachos
The study area (Figure 1) is located in northwestern Range; H, Humahuaca; P, Angosto del Pescado; PPR, Las
Argentina and straddles the Andean mountain belt from Pavas–Pescada Range; SV, Santa Victoria; ZR, Zapla
4000 m above sea level to the undeformed foreland Range. Geologic provinces: I, Eastern Cordillera; II,
plains of the “Chaco Salteño.” Three distinct geologic Subandean ranges; III, Chaco–Paraná basin (= Chaco
provinces are recognized: the Eastern Cordillera, the sub- Salteño Plain); V, Santa Bárbara System; IV, undefined
Andean ranges, and the Chacopampeana Plain. Each is a region with features similiar to other provinces.
major morphostructural unit that formed in the last 45
m.y. during the Andean orogeny. The geographic distri- Devonian time and had typical foreland basin
bution and boundaries of these provinces are controlled affinities. Considerable thicknesses of clastic
by lateral variations in the sedimentary wedges involved sediments were deposited in shallow marine envi-
in the Andean deformation. The pre-Cenozoic structural ronments. The polarity of this basin was opposite
anisotropy influenced and modified the style of the to that of the pre-Silurian stage of basin develop-
Andean deformation. ment. This phase of basin formation was termi-
Four major unconformity-bounded sequences make nated by an episode of folding and inversion
up the Phanerozoic succession (Figure 2). They are during the Late Devonian and Early Carbonif-
believed to be linked to the behavior and organization at erous, known as the Chanic orogeny.
plate boundaries. These stages are as follows: The second interval spanned the rest of the
Paleozoic, the Triassic, and much of the Jurassic.
1. Cambrian–Ordovician—The region behaved as a Continental sedimentation dominated the over-
back-arc passive margin open to the west. Thick filled Tarija basin. This was the start of the
clastic shelf sequences (> 3000 m) were deposited Gondwana succession. Late Jurassic–Early Creta-
in the Eastern Cordillera region. Farther west, ceous extension closed this episode as the old
contemporaneous turbidites were deposited in the depocenters were dissected into a suite of rift
Puna trough. This sedimentary episode ended basins and intervening highs (horsts and marginal
abruptly at the end of the Ordovician when the plateaus).
basin was inverted. Ramos (1986, 1988) attributed 3. Cretaceous–Early Tertiary—This period was char-
this to the collision of the Arequipa-Antofalla acterized by widespread development and filling
craton with the Pampean terrane, an event known of extensional basins. Fault-controlled synrift subsi-
as the Ocloyic orogeny. dence in the Early Cretaceous was superseded by
2. Silurian–Jurassic—Following terrane accretion, an Late Cretaceous–early Tertiary postrift subsidence.
intracratonic phase of basin development is Continental facies dominated both phases. The
recorded. Two major intervals separated by an transition to a compressional regime in the Andean
unconformity are recognized. The first phase of orogeny and inversion of old rift structures marked
basin subsidence spanned much of Silurian– the end of this extensional period.
Silurian–Jurassic Stratigraphy and Basin Evolution of Northwestern Argentina 253

Eastern Cordillera, it reaches an angular relationship


(Amengual and Zanettinni, 1973). The upper boundary
of the succession is also unconformable.
The Silurian–Devonian succession consists of three
stacked, kilometer-scale, coarsening-upward packages
attributed to supersequences (Starck et al., 1992a, 1993a).
Each supersequence consists of a shaley basal section
that grades upward through interbedded lithologies to
sandstone-dominated upper sections. The sharp contacts
are the basis for defining the supersequences. This
stratigraphy reflects a history of relative sea level
changes. The abrupt contacts are interpreted as first-
order flooding surfaces and are used for correlation
purposes.

Cinco Picachos Supersequence


The Cinco Picachos supersequence has a widespread
distribution. It is exposed in the Cinco Picachos Range
along the boundary between the Eastern Cordillera and
sub-Andean ranges, where it is more than 2000 m thick,
and in the Eastern Cordillera near Santa Victoria and the
Abra de Zenta-Caspalá region (Figure 1). In the north-
western corner of Argentina, the Cinco Picachos superse-
quence also crops out in the Zapla Range and in various
structural units of the Santa Bárbara system. In the San
Antonio and Aguaragüe ranges of the sub-Andean
foothills and in the Chaco Salteño, this supersequence
has been identified in exploration wells.
Table 1 summarizes the stratigraphic nomenclature of
the Cinco Picachos supersequence and regional relation-
ships. The basal Zapla Formation is generally thinner
than 50 m and consists of massive diamictites with clasts
up to 1 m in diameter (Antelo, 1978). In Bolivia, the Zapla
Figure 2—Schematic Phanerozoic chronostratigraphic equivalent is the Cancañiri Formation, which is locally
chart of the area. thicker than 1000 m. The upper part of the formation is
more stratified and contains interbedded sandstones and
shales.
4. Early Tertiary–Recent—Growth of the Andean In the Abra de Zenta–Caspalá region, the Zapla is
orogenic belt marked the final stage of basin devel- absent and the supersequence begins with sandstones
opment dominated by sequences attributed to and conglomerates of the Caspalá Formation, which
foreland basin filling. The sedimentary environ- varies in thickness from a few centimeters to 30 m
ments reflect dominant continental deposition. (Figure 3). The Zapla and Caspalá formations are
overlain by a thick coarsening-upward cycle composed
This paper addresses the Silurian–Jurassic phase of of the Lipeón, Baritú, and Porongal formations (Table 1).
basin formation and sedimentation, consisting of two This is also typical of regional counterparts in the Chaco
unconformity-bounded sequences. The lower sequence Salteño and in Bolivia (Table 1).
has a well-developed cyclicity that encourages a In the Cinco Picachos Range and the Eastern
sequence stratigraphic approach (Starck et al., 1992a). Cordillera, the Lipeón Formation is composed of a
monotonous succession of shales, siltstones, and biotur-
bated argillaceous sandstones. Fossils include abundant
SILURIAN–DEVONIAN STRATIGRAPHY Zoophycus traces and rare brachiopods. Accounting for
deformation, thickness is estimated to be about 700 m.
The Silurian–Devonian succession is represented by The Lipeón Formation grades upward through
shale and sandstone facies deposited mainly in a shallow interbedded mudstones and sandstones into the more
marine environment. Less common are littoral and conti- arenaceous Baritú Formation. Within this transition zone,
nental facies. This succession is separated from the pre- however, a disconformable relationship marked by an
Silurian by a prominent unconformity that has a clear erosive contact is present. The lower Baritú is character-
regional expression. Seismic mapping and stratigraphic ized by a variety of shallow-marine, coarsening-upward
correlations show the nature of this unconformity. In the facies sequences that show a marked cyclicity.
254 Starck

Table 1—Comparative Stratigraphy of the Study Area

The top of the Baritú Formation is more arenaceous top of the progradational supersequence.
and characterized by medium-scale cross bedding. Tran- In this context, the Zapla Formation was restricted to
sition to the Porongal Formation is via alternating the more actively subsiding depocenters. This basin
mudstone and sandstone facies. The sandstones facies setting and the chaotic nature of some of its sedimenta-
generally have erosive bases and some form fining- tion suggest that the Zapla Formation originated as a
upward units with plant debris in places. A fluvial depo- lowstand.
sitional setting is inferred.
The Porongal Formation consists of conglomeratic Las Pavas Supersequence
sandstones and fine, mature conglomerates with
interbedded mudstones. Trough cross bedding is The Las Pavas supersequence is exposed in the Cinco
common, especially in the sandstone facies. The contacts Picachos Range and forms much of the core of the Las
between depositional units have layers of quartzitic Pavas–Pescado Range (Figure 1). In the Eastern
conglomerates suggestive of winnowing processes. Cordillera, it also crops out in the Abra de Zenta–Caspalá
The progradational supersequence was deposited area before finally disappearing at Caspalá latitude
over transgressive sediments that onlapped the Ordovi- where it was eroded to form the pre-Carboniferous
cian substrate in an expanding epeiric basin. The Caspalá unconformity (Starck et al., 1992a, 1993a). It is probable
Formation is thus a sandy transgressive body that was that the Las Pavas also occurs in the Zapla Range and
deposited in a back-stepping fashion as the coastal facies Santa Bárbara system because fossils characteristic of this
equivalent of the Lipeón shelf mudstones. Littoral and supersequence have been reported there (Feruglio, 1929).
continental facies of the Porongal and upper Baritú This stratigraphy has been extensively drilled in the sub-
formations were deposited above shallow marine facies Andean ranges where the Icla and Huamampampa
of the lower Baritú and Lipeón formations. However, formations are recorded. Their counterparts in the Chaco
progradation was not uniform. Stratigraphic relation- Salteño are the Rincón and Michicola formations (Table 1).
ships show sudden lithologic changes and facies shifts The Las Pavas, with thicknesses on the order of 900 m,
typical of third-order variations. The Cinco Picachos is separated from the underlying Cinco Picachos by a
supersequence reflects progradational stacking of several prominent flooding surface. Five depositional sequences
depositional sequences. Palynologic studies confirm an are progradationally stacked over this flooding surface.
increasing component of continental material toward the Each sequence displays thickening- and coarsening-
Silurian–Jurassic Stratigraphy and Basin Evolution of Northwestern Argentina 255

quences have extensively bioturbated topsets and are


laterally continuous. They are easily correlatable between
the Las Pavas–Pescado Range and wells located to the
east (Figure 4).
Malvinokaffric invertebrate assemblages are common
in the toesets of the lower three sequences; brachiopods
dominate these assemblages. Fossil plants occur in the
upper sequences.

Aguaragüe Supersequence
A major flooding surface separates the Aguaragüe
and Las Pavas supersequences. The Aguaragüe consists
of the Los Monos and Iquiri formations (Table 1). It also
typically coarsens upward due to the superimposition of
the sandy Iquiri Formation over the Los Monos
mudstones. In Argentina, these units are best developed
in the subsurface. The pre-Carboniferous erosion has
largely stripped the Devonian section in the outcropping
part of the basin, thus explaining why Los Monos
outcrops are rare in Argentina (Starck et al., 1993a). In the
Eastern Cordillera this supersequence is entirely eroded.
The thickest surface sections are less than 200 m thick
compared to the nearly 1000-m thickness in the sub-
Andean foothills and Chaco Salteño.
The Los Monos Formation comprises black laminated
shales with centimeter-scale sandstone partings. Biotur-
bation is generally moderate and macrofossils are rare.
The Argentinian outcrops contain some crinoids and
plant fossils. Haplostigma furquei is the most common
fossil in the Bolivian section.
A widespread sandstone body with a gradational
base is present in outcrops and wells near the base of the
Figure 3—The basal unconformity of the Silurian–Devonian supersequence. It represents the upper part of the
succession near Caspalá, Eastern Cordillera. lowermost of the several depositional sequences that
make up the Aguaragüe supersequence.
upward trends. The sequences become more argillaceous
distally (eastward). The lithostratigraphic boundary Interpretation of the Silurian–Devonian
between the Icla and Huamampampa formations climbs Succession
through these sequences in a progradational direction.
On a regional scale, the Icla–Huamampampa boundary Stratigraphic analysis implies an overall relative sea
is diachronous. The Icla contains variable thicknesses of level or base level control on sedimentation and cyclicity.
the finer grained lower parts; the Huamampampa This sequence stratigraphic approach distinguishes
provides a topset cap of sandstone facies. several genetically related sequences of variable thick-
The lowest depositional sequence is the thickest, nesses within the Silurian–Devonian succession. There
locally exceeding 300 m; its base forms the Cerro Piedras are likewise different scales of investigation and correla-
shales. Overlying sequences average about 100 m in tion. On the smallest scale, parasequences within the
thickness. Each sequence typically begins with a shale- upper Las Pavas supersequence are correlatable for more
draped flooding surface, above which distal shales are than 30 km (Figure 4). Depositional sequences of the Las
replaced vertically by proximal sandstones. Continental Pavas can be traced, albeit with difficulty, from the Cinco
facies are only locally preserved. Flooding surfaces rather Picachos Range to the wells drilled in the sub-Andean
than unconformities form sequence boundaries. ranges 100 km to the east. On the largest scale, the super-
However, several sandstone bodies in wells in the sub- sequences are major stratigraphic bodies that span the
Andean ranges have sharp, erosive bases. These are entire basin.
interpreted as lowstand system tracts. Similar relation- This hierarchy of northwestern Argentinian sequences
ships are observed in Bolivian outcrops. is believed to have global significance. Comparison of
Coarsening-upward cycles are well developed in southwestern Gondwana sequences with global sea level
some sequences, especially the fifth depositional curves has been attempted, for example, South Africa
sequence in the Las Pavas–Pescado Range. This fifth (Cooper, 1986) and Brazil (Melo, 1989). There are strong
sequence contains a stack of parasequences resembling similarities between the Devonian basins of south-
those of the lower Baritú Formation. These parase- western Gondwana. The flooding surface that marks the
256 Starck

Figure 4—Stratigraphic
correlation of the uppermost
depositional sequence of
the Las Pavas superse-
quence. Location of wells
and surface sections are
shown in Figure 1. Note the
lateral continuity of parase-
quences and parasequence
sets.

beginning of the Las Pavas supersequence, and coincides Tankard and Barwis (1982) attributed Bokkeveld
with the Malvinokaffric faunal paleogeography, is wide- cyclicity to variable subsidence rates superimposed on an
spread. This event is believed to mark the base of the approximately constant eustatic rise. In contrast, Cooper
Ponta Grossa Formation in the Paraná basin of Brazil, the (1986) related Bokkeveld facies shifts entirely to sea level
base of the Lolén Formation in the Ventania basin of changes. The proposed correlation of the Bokkeveld
Argentina, the base of the Fox Bay Formation in the Group with the Las Pavas supersequence and their rela-
Malvinas Islands, and the base of the Bokkeveld Group tionship to the Johnson et al. (1985) eustatic curve is
in South Africa. The flooding surface at the base of the shown in Figure 5. The comparison of the Las Pavas
Aguaragüe supersequence may also have had a wide- supersequence and the eustatic curve is based on correla-
spread distribution, but it was eroded in several places tion of flooding surfaces that bound the supersequence.
prior to Carboniferous deposition. The micropaleontologic data support a late Emsian–early
The well-preserved stratigraphic succession in the Givetian age. In this sense, the Las Pavas spans the Ic to
Cape basin of South Africa bears comparison with the If transgressive–regressive cycles of Johnson et al. (1985).
Silurian–Devonian basin of northwestern Argentina and The Cinco Picachos supersequence spans approxi-
Bolivia. The South African units are the Table Mountain mately 45 m.y. from the uppermost Ordovician (Monaldi
Group, the upper part of which correlates with the Cinco and Boso, 1986) with the Zapla diamictites to the Emsian.
Picachos supersequence; the Bokkeveld Group, which is The top of the supersequence is equated to the Ib trans-
equivalent to the Las Pavas supersequence; and the gressive–regressive cycle of Johnson et al. (1985). In
Witteberg Group, which is a counterpart of the contrast, the Las Pavas supersequence is believed to have
Aguaragüe supersequence. The Bokkeveld Group is had a 13-m.y. span from the late Emsian to middle
subdivided into several formations, each consisting of an Givetian (Ic to If cycles of Johnson et al., 1985) (see Figure
upward-coarsening unit as much as 100 m thick 5). The Aguaragüe supersequence lasted 20 m.y. from
(Tankard and Barwis, 1982; Theron and Loock, 1989). the middle Givetian (IIa cycle) to possibly the Early
These cycles match the depositional sequences recog- Carboniferous. Internally, the duration of each sequence
nized in the Las Pavas supersequence. The similarities within the Las Pavas had a periodicity of 2–4 m.y., equiv-
include lithostratigraphy as well as faunal associations in alent to the third-order cycles of Vail et al. (1977). These
the Bokkeveld Group (Tankard and Barwis, 1982; Hiller interpretations are tenuous and require further biostrati-
and Theron, 1989). graphic testing.
Silurian–Jurassic Stratigraphy and Basin Evolution of Northwestern Argentina 257

Figure 5—Comparison of Middle Devonian records of South Africa with northwestern Argentina and global sea level trends.
(a) Devonian eustatic curve (Johnson et al., 1985). (b) South African section (modified from Tankard and Barwis, 1982). (b)
Northwestern Argentinian section; I–V are depositional sequences.

This study has focused on a Paleozoic intracontinental graphy is expressed in angular relationships (approxi-
basin setting in northwestern Argentina where mately 1˚). This unconformity is most pronounced in the
sequences accumulated on a ramp-type passive margin southwestern part of the area, where the Silurian–
(see Van Wagoner et al., 1990). Progressive flexural Devonian section in parts of the Eastern Cordillera is
subsidence has limited the development of widespread completely eroded (Figure 6). There the Carboniferous
unconformities. Instead, sequence boundaries are marked overlies Ordovician rocks. A network of deeply eroded
by maximum flooding surfaces. Lowstand system tracts paleovalleys is superimposed on this regional unconfor-
are poorly developed and are difficult to separate from mity, incising different levels of the Devonian (Starck et
the preceding sequence. Possible lowstand deposits may al., 1992a, 1993a). The top of the Carboniferous–Jurassic
be the sharp-based sandstone bodies in the upper Las succession is marked by another major unconformity.
Pavas supersequence, the widely dispersed Porongal This unconformity is the most conspicuous in northern
conglomerates, and the sharp-based lower Baritú Argentina (Figure 2), where the pre-Cretaceous record
Formation sandstones. Posamentier et al. (1992) have has been stripped over large areas.
attributed similar associations to forced regressions. The succession is subdivided into four groups that are
genetically distinct and separated by unconformities: the
Macharetí, Mandiyutí, Cuevo, and Tacurú superse-
CARBONIFEROUS–JURASSIC quences.
STRATIGRAPHY
Macharetí Supersequence
The Carboniferous–Jurassic stratigraphic interval is
characterized by a succession of depositional sequences Filling of the Tarija basin began with deposition of the
deposited in the Tarija basin. Each sequence is separated Macharetí Group above an unconformity that had
from the next by an unconformity. Terrestrial deposi- beveled the Silurian–Devonian substrate. Three of the six
tional environments in an overfilled basin setting are formations making up this group in Bolivia occur in
ubiquitous. northern Argentina. These are the Tupambi, Itacuami,
The base of the succession overlies a major regional and Tarija formations (Table 1). The Itacuami and Tarija
unconformity that is locally modified by deep erosion. formations appear to be genetically related and will be
Gentle regional tilting of the pre-Carboniferous strati- discussed together.
258 Starck

Figure 6—Stratigraphic relationships at the southwestern border of the basin (Eastern Cordillera). See Figure 1 for location.
(a) Southern end of Cianzo syncline; (b) Caspala Creek; (c) Pluma Verde Creek.

Tupambi Formation are not obviously related to the margins of the basin.
The Tupambi Formation consists of whitish sand- Starck et al. (1992b, 1993b) have documented the
stones that range from a few meters to several hundred glacial origin of these diamictites. The evidence includes
meters thick. These sediments were deposited in paleo- striated pavements with northwest-southeast oriented
valleys. The valley fills are irregularly stacked, coars- striations, dropstones, and true tillite characteristics of
ening-upward sequences characterized by synsedimen- the diamictites. The tillites are believed to have origi-
tary deformation and slumping in their lower half. nated as lodgement tills and locally as glaciomarine
Diamictites are present at the toes of some cycles. deposits. Ice flow direction was toward the northwest.
Tupambi paleogeography included braided fluvial The Chorro sandstone facies in the upper Tarija are
deposits and deltaic mouth bars, mainly confined to the interpreted as periglacial fluvial sediments that were fed
paleovalley network. High sedimentation rates and into the basin through valleys incised along the edge of
possible basin tectonism are reflected in the widespread the basin, there forming small sandy deltas.
synsedimentary deformation. The diamictites are attrib-
uted to distal glacial influences. Mandiyutí Supersequence
Tarija Formation The Mandiyutí Group is separated from the
The Tarija Formation is a blanket of dark-colored Macharetí by a prominent unconformity. Along the
diamictites that are locally more than 600 m thick. In southwestern margin of the basin in the Eastern
some parts of the Eastern Cordillera, it overlies a striated Cordillera, the Mandiyutí Group oversteps the Macharetí
pavement. Conglomerates, sandstones, and mudstones and directly overlies pre-Carboniferous (Ordovician)
are interbedded with the diamictites. One 70-m-thick stratigraphy (Figure 6). The basal unconformity is deeply
varved or laminated mudstone at the base of the Tarija eroded by paleovalleys. Seismic reflection and well data
(the Itacuami Formation of some authors) contains drop- show that these valleys are widespread in the Chaco
stones. Salteño and in the Bolivian Chaco (Salinas et al., 1978;
The Tarija diamictites are generally homogeneous, Tankard et al., 1995). Two contrasting lithologic units are
have practically no stratification, and contain clasts up to recognized: the Las Peñas and San Telmo formations.
1 m in diameter. Faceted, polished, and striated clasts are
common. The conglomerate and sandstone intercalations Las Peñas Formation
are common, especially near the top of the Tarija The Las Peñas Formation and its Bolivian counterpart,
Formation. (In Bolivia, this sandstone-dominated upper the Escarpment Formation, comprise stacked channel
Tarija has been given the name Chorro Formation.) sandstone units occupying deeply incised paleovalleys.
These sandstone units have an irregular distribution and Thickness varies up to 500 m. In outcrop this is a major
Silurian–Jurassic Stratigraphy and Basin Evolution of Northwestern Argentina 259

Cangapi Formation
The Cangapi Formation is a homogeneous, white
sandstone overlying a thin basal conglomerate, for
example, in the Cinco Picachos Range. Interbedded
siliceous and calcareous layers near the top mark the
transition to the overlying Vitiacua Formation. The 100-
m-thick Cangapi is attributed to eolian processes.

Vitiacua Formation
The Cangapi grades vertically into the Vitiacua
Formation, a 30- to 90-m-thick interval of calcareous
mudstones and limestones with sandstone interbeds.
Conspicuous meter-scale limestone cycles with chert
nodule caps are recognized. Green and purple
mudstones are several meters thick. The sandstone units
are cross bedded and locally herringbone cross bedded.
Sempere (1995) suggests a middle Permian–Early
Triassic age. The calcareous mudstones resemble the Iratí
Formation of the Paraná basin.
The calcareous mudstones and limestones with fossils
and herringbone cross stratification are indicative of
deposition in a carbonate shelf setting (Beltan et al., 1987;
Sempere et al., 1992). The meter-scale limestones form
shallowing-upward cycles in which the crests are brec-
ciated, karstic, and preserve evidence of subaerial
exposure. Flooding surfaces mark the beginning of each
cycle. Eolian dune fields surrounded the basin, as
evidenced by Cangapi sandstones and eolian sandstone
Figure 7—The Tarija basin succession in the Eastern tongues within the Vitiacua.
Cordillera (section c of Figure 6). The Las Peñas Formation
onlaps a paleovalley and pinches out to the right (north). Ipaguazu Formation
Overlying an unconformable base of karstic and brec-
ciated limestones, the Ipaguazu Formation consists of
red siltstones with interbedded mudstones and very fine
cliff-forming unit (Figure 7). Like the Tupambi
grained sandstones in a sequence 60–90 m thick. The
Formation, the Las Peñas displays widespread deforma-
most common sedimentary structures are current and
tion. Deposition is attributed to sandy fluvial systems
wave ripples and less common hummocky cross stratifi-
and overstepping, short-headed delta systems confined
cation. In the northern part of the study area gypsum
by the valley margins.
evaporites occur. This formation was probably deposited
San Telmo Formation in an arid lacustrine environment. In Bolivia the evapor-
ites are more abundant and consist of gypsum and halite.
In contrast, the succeeding San Telmo Formation is
composed of alternating beds of very fine sandstones,
mudstones, and bright red diamictites. The diamictites
Tacurú Supersequence
contain cross-bedded sandstone units. The thickness of The Tacurú Group is extensively preserved in Bolivia
this formation is up to 400 m in the sub-Andean ranges, and in the Argentinian sub-Andean ranges that border
and less than 150 m in the Eastern Cordillera. Deposition Bolivia. Starck et al. (1992b, 1993b) have documented
is believed to have occurred in glacial, fluvial, and lacus- Tacurú outliers even farther south in the Eastern
trine environments. Cordillera. The pre-Tacurú unconformity erodes deep
into the underlying units. However, it is uncertain
Cuevo Supersequence whether this unconformity reflects structuring or a drop
of base level.
Preserved remnants of the Cuevo Group are restricted This group is composed mainly of reddish sandstones
to isolated structures in the sub-Andean ranges in the with large-scale cross bedding. In northern outcrop
northern part of the study area. Starck et al. (1992b, areas, thinly bedded mudstones and fine-grained
1993b) have described Cuevo outcrops in the Rio Cañas conglomerates intertongue with the sandstones. There is
area. This group consists of three formations: Cangapi, also a basin margin facies in the Eastern Cordillera;
Vitiacua, and Ipaguazu (Table 1). Sempere (1995) assigns substantial thicknesses of breccia crop out in the Cianzo
a Pennsylvanian–Early Triassic age to the Cuevo super- zone and along the eastern edge of the Humahuaca
sequence. Valley. Although Amengual and Zanettinni (1973) assign
260 Starck

Carboniferous–Early Permian, following a well-docu-


mented eustatic sea level drop (Veevers and Powell,
1987; Ross and Ross, 1988). These supersequences are
correlated with the Sachayoj, Charata, and Chacabuco
formations of the Chaco-Paraná basin, the Itararé Group
of the Paraná basin, the Sauce Grande Formation of
Ventania, the Malvinas Lafonia Formation, and the
Dwyka Group of the South African Karoo basin.
All of these glacially derived units unconformably
overlie Devonian strata and to a great extent are the
record of a massive continental ice cap. Glacial striations
(Salinas et al., 1978; Starck et al., 1992b, 1993b) indicate a
northwest-directed ice flow in the Tarija basin, consistent
with a radial advance from a paleopole located east of
South Africa. The petrology and ice flow direction in the
Tarija basin suggest a Brazilian shield provenance (also
see López Gamundi, 1986).
Figure 8—Angular unconformity between Tacurú and Salta The unconformity between the Macharetí and
groups in Eastern Cordillera. In this position, the Tacurú Mandiyutí groups is also attributed to a eustatic drop. In
(Jurassic?) is the floor of a half-graben filled by more than both groups, lowering of base level induced deep valley
5 km of Salta Group (Cretaceous–early Tertiary) sedimen- incision. For this reason, both supersequences start with
tary rocks. sand-prone deposition on a highly irregular surface
(e.g., Tupambi and Las Peñas formations) and are
blanketed by widespread glacial or periglacial deposits
these breccias to the younger Pirgua Subgroup, they are
(e.g., Tarija and San Telmo formations). Elsewhere, the
laterally related to cross-bedded Tacurú sandstones and
Itararé Group of the Paraná basin contains three genetic
have an angular unconformable relationship with the
units of arenaceous deposits and diamictites (França and
overlying Pirgua Subgroup (Figure 8).
Potter, 1991; Eyles et al., 1993), not unlike the Tarija
Tacurú paleogeography consisted mainly of eolian
basin stratigraphy.
dune fields and small lakes that deposited the intrafor-
The Cuevo supersequence was deposited on an
mational mudstones. In Bolivia, these lacustrine
unconformity of regional scale, and internally contains a
sediments are more extensive, warranting a unique
secondary unconformity that separates the Cangapi-
description as the Castellón Formation. The breccias in
Vitiacua pair from the Ipaguazu Formation (Table 1).
the Eastern Cordillera with angular meter-size boulders
This group has traditionally been assigned a Late Triassic
reflect proximal alluvial fan environments adjacent to
(Norian) age (Beltan et al., 1987). However, more recent
fault scarps.
studies (Starck et al., 1992b, 1993b; Sempere et al., 1992;
Sempere, 1995) suggest a Late Carboniferous–Early
Interpretation of the Carboniferous– Triassic age for the Cuevo. Deposition of the Cuevo
Jurassic Succession (Tarija basin), Passa Dois (Paraná basin), and Ecca (Karoo
basin) groups occurred after the Gondwana glaciation.
The Carboniferous–Jurassic succession was con- The Iratí and Whitehill transgression (Visser, 1991),
structed by a suite of unconformity-bounded deposi- warm climatic conditions, and reptile and palynomorph
tional sequences. Although tenuous, their chronology is fossils indicate a Permian age.
derived from regional correlations; unfortunately, too The overlying Ipaguazu Formation is probably of
little detailed biostratigraphic information is available. Triassic age. The Ipaguazu may be part of the Sereré
The Carboniferous stratigraphy is separated from the supersequence (Oller and Sempere, 1990; Sempere, 1995)
Devonian by a major unconformity. Starck et al. (1992a, for which a Middle Triassic to possibly Middle Jurassic
1993a) have correlated this unconformity to the Chanic age is inferred.
event as well as to global sea level history (see Ross and The approximate age of the Tacurú supersequence is
Ross, 1988; Veevers and Powell, 1987), suggesting a based on regional correlations and structural considera-
hiatus of approximately 50 m.y. tions. This group is part (together with the previous
It was on this erosional platform that the Macharetí stratigraphy) of the prerift succession of the Cretaceous
and Mandiyutí supersequences accumulated. Microfos- extensional stage. Its age is thus pre-Cretaceous. It could
sils in the thick glacial deposits indicate a Middle–Late also be argued that the Tacurú Group is a counterpart of
Carboniferous age (Azcuy and Laffitte, 1981; Azcuy et the Tacuarembó and Botocatú eolianites of the Chaco-
al., 1984). Recent palynologic analysis suggests a West- Paraná and Paraná basins, respectively, in which case it
phalian or younger age for these supersequences or could date to the Early Jurassic. The implication is that
groups. These ages match those of other glacial during latest Triassic–Jurassic time, extremely arid condi-
sequences throughout southwestern Gondwana tions prevailed over much of southwestern Gondwana,
(Tankard et al., 1995). Consequently, the Macharetí and resulting in large deserts of which the Tacurú was a local
Mandiyutí groups were deposited during the Late expression.
Silurian–Jurassic Stratigraphy and Basin Evolution of Northwestern Argentina 261

IMPLICATIONS FOR BASIN EVOLUTION Unconformities are deeply incised and form wedges
toward the western highland. For example, the Ocloyic
The dynamics of basin evolution in northwestern and Chanic unconformities are markedly angular in the
Argentina are poorly understood. Mesozoic and Tertiary Caspalá area, as is the basal Tacurú unconformity.
tectonism have largely reactivated old basin-forming Faulting in the Eastern Cordillera was contemporaneous
structures and deformed the stratigraphic column. These with Tacurú deposition.
effects, combined with erosion, have resulted in a poorly The western margin of the basins is represented by an
preserved record of early basin architecture. Neverthe- orogenic belt. This belt, located in the Eastern Cordillera,
less, throughout much of the Silurian–Jurassic interval, developed as a highland in the middle Paleozoic, sepa-
the basins appear to have had intracratonic affinities with rating the Paleozoic Puna basins from those of the sub-
an overall north-south continuity along the Pacific Andean realm (Salfity et al., 1975; Vistalli, 1989). Ramos
margin of Gondwana. There also appears to have been a (1986, 1988) attributed this uplift to the effects of a
broadscale continuity in Paleozoic time of the north- collision orogen formed by the accretion of the Antofalla-
western Argentinian, Chaco, Chaco-Paraná, and Paraná Belen-Arequipa terrane. The Puna basins are interpreted
depocenters. For most of the Paleozoic, these depocen- as peripheral basins (Ramos, 1986; Palma and Irigoyen,
ters were yoked together. Only in Mesozoic time did 1987) separated from the sub-Andean flexural foreland
extension dissect this landscape into isolated rift basins. basins. The Silurian–Devonian succession also overlies
As incomplete as it is, the stratigraphic record and the an angular unconformity in the Puna realm (Donato and
distribution of its unconformities preserve some of the Vergani, 1985).
tectonic history of these basins. The Cinco Picachos, Las Pavas, and Aguaragüe super-
The Silurian–Devonian cycle had a depositional sequences fill a wide basin in which they display a
border or basin margin in the west. Facies trends show a pronounced facies asymmetry as a result of sediment
western provenance. Transgressive relationships within supply from the west. This basin is believed to have had
the Lipeon Formation confirm this interpretation. similarities to the Laramide foreland basin of North
The Macharetí and Mandiyutí supersequences also America (see Cross, 1986). A flanking foldbelt in the west
record the basin geometry in the Eastern Cordillera. supports this interpretation. This foldbelt is believed to
Whereas the Macharetí supersequence onlaps and have been formed by the Ocloyic orogeny.
pinches out against pre-Carboniferous strata, the Ramos et al. (1984) have interpreted the Early
Mandiyutí oversteps even farther to the west beyond Carboniferous Chanic orogeny in southern latitudes as a
outcrop control. The fluvial paleovalley fills of the collisional event in plate tectonic terms. Although this
Tupambi Formation have eastward-directed paleocur- orogeny was not intense in the study area, the dias-
rents (Starck et al., 1992b, 1993b). trophism and hiatus represent the onset of a new
The Cuevo supersequence has a tabular or saucer- tectonostratigraphic cycle and formation of the Tarija
shaped geometry on a regional scale, suggesting low rates basin. These various stages of basin development are
of regional subsidence. In outcrops of the Cinco Picachos summarized in Figure 9 as calculated subsidence curves
Range, the Cangapi Formation has a southward- for Caspalá, Cinco Picachos Range, and Chaco Salteño.
onlapping internal geometry. The Vitiacua Formation The increasing angularity of the Lower Carboniferous
also suggests low rates of subsidence with its dark Chanic unconformity toward the west reflects the farfield
mudstones and thin but laterally continuous limestone effects of the compressive Pacific–Gondwana plate
cycles. margin. These effects can also explain the subsidence
The Ipaguazu Formation unconformably overlies the along the western fringe of the basin during Carbonif-
Vitiacua with a depocenter in the Entre Rios area of erous–Permian time. Eyles and Eyles (1993) have noted
southern Bolivia. Oller and Sempere (1990) have inferred the farfield effects of the Pacific plate margin as far east
a period of extensional subsidence for the Ipaguazu. as the Paraná basin. Figure 9 shows that during Tacurú
However, I have been unable to find any evidence of time, the Tarija basin followed a course of subsidence
fault-controlled subsidence in northwestern Argentina different from earlier episodes. Although the base of this
that would separate the Vitiacua and Ipaguazu styles of interval is a regional unconformity, there were no
subsidence. The unconformable relationship is believed obvious tectonic climaxes. Faulting along the margins of
to reflect a drop of base level and restriction of Ipaguazu the basin were extensional in nature and probably
sediments to the more actively subsiding parts of the presaged the late Mesozoic rifting.
basin, such as Entre Rios, where thick evaporites occur. Finally, Table 2 summarizes this tectonostratigraphic
The upper Ipaguazu marks expansion of the basin and interpretation for northwestern Argentina. The pre-
onlapping of its margins. In contrast, thick breccias Silurian part of the geology not covered in this study
stacked about the margin of the Tacurú basin imply includes igneous and sedimentary rock assemblages
fault-controlled or extensional subsidence in the west. attributed to a Late Proterozoic–early Paleozoic passive
The stratigraphic evidence is overwhelming that, from margin (González Bonorino and González Bonorino,
Silurian time on, northwestern Argentinian and Bolivian 1991). An Ordovician magmatic arc or Puna eruptive belt
basins were bordered to the west by a highland area created a fore-arc basin and flysch depocenters (Méndez
large enough to supply enormous amounts of relatively et al., 1972; Coira et al., 1982; Spalletti et al., 1989). The
coarse sediments. Between this orogenic belt and the Silurian–Jurassic stratigraphy of the present study
Brazilian shield, the basins subsided intermittently. explores the subsequent evolution of this orogenic belt.
262 Starck

Figure 9—Subsidence curves calculated at different basin positions. Note the different Silurian–Devonian, Carboniferous–
Middle Triassic, and Jurassic evolutive stages. Note also the increasing intensity of the unconformities toward the Eastern
Cordillera (Caspalá).

DISCUSSION The Cuevo supersequence closed the late Paleozoic–


Triassic phase of basin formation with the lowest rates of
At the end of the Ordovician, collision of the subsidence (Figure 9). Cuevo subsidence appears to have
Arequipa-Antofalla craton produced a fold and thrust been restricted to the center of the basin. By the end of
belt along the trend of the Eastern Cordillera. This belt this episode, however, a new diastrophism resulted in a
continued southward into the Pampean ranges. A widespread unconformity and marked a change in
foreland basin subsided in front (east) of this compres- regional stress fields as extensional processes started to
sive load. Due to the high sea level during the Silurian–- partition the Paleozoic landscape in Tacurú time. Exten-
Devonian, the various southwestern Gondwana basins sional subsidence was not yet intense. Tacurú sedimenta-
were generally yoked together as underfilled shallow tion appears to have blanketed this incipient and
marine basins (see published maps, e.g., Melo, 1989; restricted extensional topography.
Barret and Isaacson, 1989; Isaacson and Sablock, 1990). In northwestern Argentina, a fundamental change in
An overall fining of the succession may imply a gradual the style of subsidence occurred with the Late Jurassic
lowering of topographic relief. (Oxfordian–Kimmeridgian) Araucanian diastrophism,
At the end of the Devonian, or more likely at the which effectively terminated the Tarija basin (the
beginning of the Carboniferous, another diastrophic Bolivian counterpart is known as the Condo event).
event occurred. This event was probably related to a Extension dissected the old epeiric basin into a suite of
minor reorganization of the Pacific plate edge relative to fault-controlled half-grabens and intervening highs. The
Gondwana. This event is the Chanic orogeny, which Cretaceous–lower Tertiary Salta Group developed as a
induced gentle tilting of the pre-Carboniferous succes- rift basin fill. The prerift successions of this study were
sion and development of a regional angular unconfor- systematically eroded from the rift margins as a result of
mity. This unconformity increases in angularity toward rift flank uplift (Figure 10). For example, the southern
the west. Seismic sections show that this unconformity edge of the Tarija basin was uplifted along a basin
persists through the Chaco-Paraná basin (especially in margin arch known as the Michicola high, south of
the Alhuampa subbasin). The Chanic diastrophism and a which the Lomas de Olmedo half-graben formed (see
marked eustatic drop in the Middle Carboniferous Cominguez and Ramos, 1995). Figure 10 shows the struc-
combined to inhibit sedimentation in the region until the tural asymmetry of the Lomas de Olmedo rift. Its
Late Carboniferous. northern margin consists of a narrow zone of south-
In spite of the middle Paleozoic diastrophic events, the dipping listric normal faults, interpreted here as a major
late Paleozoic Tarija basin had a distribution similar to basin-forming fault system. The southern part of the
the earlier Silurian–Devonian basins, probably because basin is a fault-terraced ramp. Reflection seismic profiles
they reused older structural fabrics. Onlap relationships show the essential structural architecture of this basin
(Figure 6) show that the Tarija basin was more restricted (Figure 11). Some of the deeper reflection patterns
than its predecessors. Rates of subsidence were also suggest that Paleozoic strata may occur in fault blocks
lower (Figure 9), possibly due to thermal behavior of the beneath the synrift Salta succession (Starck et al., 1992b,
lithosphere. These lower rates of subsidence emphasized 1993b; Cominguez and Ramos, 1995). Silurian–Jurassic
the unconformities cut by glacioeustatic oscillations that outcrops in the Abra de Zenta–Caspalá region also owe
characterized the late Paleozoic ice age. their preservation to extensional processes.
Silurian–Jurassic Stratigraphy and Basin Evolution of Northwestern Argentina 263

Table 2—Tectonic Evolution of Northwestern Argentina

Cretaceous–early Tertiary extension and subsequent Figure 10—Paleogeologic map of the early Salta Group
Andean compressive deformation reactivated old crustal (Cretaceous–early Tertiary) synrift fill. Note the increasing
fabrics, many of them of Eocambrian origin. The impor- stripping of prerift sequences from the north toward the
tance of these rift basins, their rift flank uplifts, and Lomas de Olmedo rift margin (= Michicola high) and the
normal fault system. Symbols: Ag, Aguaragüe superse-
arches (e.g., the Michicola and Pampean arches of
quence; CP, Cinco Picachos supersequence; Cue, Cuevo
Argentina and the Izozog arch of Bolivia) is that they Group; LP, Las Pavas supersequence; Mch, Macharetí
effectively dismembered the Silurian–Jurassic epeiric Group; My, Mandiyutí Group; Ord, Ordovician; Tac, Tacurú
basin. The Chaco, Tarija, and Chaco-Paraná system is Group.
believed to have been a continent-wide basin complex,
consisting of several depocenters that were generally absorbed most of the shortening. The normal faults of the
yoked together (Starck et al., 1991b, 1992b). The strati- Lomas de Olmedo basin were little inverted because
graphic record of northwestern Argentina reflects the their orientations paralleled the direction of maximum
broadscale history of southwestern Gondwana. The compressional stress of the Andean deformation.
Paleozoic succession records the processes of terrane Farther south, the Santa Bárbara system is a geologic
accretion, continent building, and compressive deforma- province (Rolleri, 1976) consisting of short, broad,
tion that are emphasized in the Ocloyic and Chanic basement-involved anticlines. High-angle reverse faults
events. In contrast, the Mesozoic was dominated by display varying strikes and vergences where they have
extensional processes that culminated in continental frag- inverted normal faults. These Santa Bárbara extensional
mentation and opening of the Atlantic Ocean. The faults were more favorably oriented relative to the
earliest extensional record in northwestern Argentina compressive stress field than those of Lomas de Olmedo
can be dated at the base of the Tacurú Group, whereas and thus were more susceptible to inversion. Much of
the Late Jurassic Araucanian diastrophism marks the the Eastern Cordillera was subjected to inversion. For
climax of widespread extensional stresses. Elsewhere in example, Abra de Zenta–Caspalá outcrops are inverted
Patagonia, extension is dated as far back as the Late parts of the prerift substrate of a Salta Group half-graben.
Triassic (Uliana et al., 1989). The pre-Tertiary tectonostratigraphic framework
The study area straddles four geologic provinces (see affected the style of Andean deformation in two ways.
Turner, 1979). Provinces of the Andean mountains First, the stratigraphy determined the way sedimentary
include the Puna, the Eastern Cordillera, and the sub- wedges were involved in deformation. Second, the orien-
Andean ranges. To the east is the undeformed foreland tation and shape of Salta-age extensional faults influ-
of the Chacopampean Plain or Chaco Salteño. The sub- enced their reactivation. The present-day distribution of
Andean ranges are a thin-skinned fold and thrust belt the Silurian–Jurassic succession reflects erosion related to
with prominent detachments in the Lipeón and Los Cretaceous extension, especially the Late Jurassic Arau-
Monos formations. The erosion of the Los Monos canian diastrophism.
Formation represented by the Araucanian unconformity The commercial consequences of this geologic history
along the Michicola arch controlled the southern termi- have long been understood, particularly the Devonian
nation of the sub-Andean ranges. At the latitudes of the source rocks, the mainly Carboniferous reservoirs, and
Lomas de Olmedo rift system, the Eastern Cordillera Tertiary trap formation. Dynamic studies of the strati-
264 Starck

Figure 11—Seismic section across Lomas de Olmedo rift. See Figure 10 for location. Note the asymmetry of major basin-
forming faults along the northern flank and minor faults on the opposite side. The northern flank corresponds to the
Michicola high, where the entire stratigraphic section discussed in the text has been eroded. Symbols: Pr, prerift sequence
(Paleozoic–Jurassic); Sr, synrift sequence (Early–Late Cretaceous); Po, postrift sequence (Maastrichtian–Oligocene?); To,
distal Andean foreland basin infill. (From Cominguez and Ramos, 1995.)
Silurian–Jurassic Stratigraphy and Basin Evolution of Northwestern Argentina 265

Hydrocarbon generation began several million years


after deposition of the Los Monos shales. The complex
geologic evolution of this region, including several major
unconformities, makes maturation and expulsion
modeling difficult (e.g., Lopatin, 1971). Nevertheless, the
beginning of maturation is believed to have occurred in
the Jurassic. Because of the Late Jurassic deformation,
regional variations in erosion, and thermal history, the
Los Monos source rock followed different maturation
paths in different parts of the basin. Much of the Los
Monos still appears to be within the generative window.
The expulsion and migration are believed to be related to
Andean deformation.

Aknowledgments This study is the result of my work at


YPF. I gratefully thank all the YPF personnel, especially those
who accompainied me in the 4th Geologic Commision. I also
thank the editors and referees (H. Belotti and M. Mozetic) for
Figure 12—Oil and gas field distribution. Fields related to
the Paleozoic are in black; fields belonging to the Creta-
their constructive comments and suggestions.
ceous–lower Tertiary Salta Group basin are in gray. All
Paleozoic fields have developed where the Los Monos
Formation is present. Reservoir ages: D, Devonian; C,
Carboniferous; T, Tertiary.
REFERENCES CITED
Amengual, R., and J. Zanettinni, 1973, Geología de la Comarca
de Cianzo y Caspalá: Revista Asociación Geológica
graphy and basin evolution enable us to understand Argentina, Buenos Aires, Argentina, v 28, p. 341–352.
these parameters better. Antelo, B., 1978, Formaciones de edad silúrica en el noroeste
Source rock geochemistry confirms that the Los argentino (Provincias de Jujuy y Salta): Revista Asociación
Monos Formation of the Aguaragüe supersequence is Geológica Argentina, Buenos Aires, Argentina, v. 33,
the major source rock interval. Although the TOC p. 1–16.
content of this formation generally does not exeed 1%, its Azcuy, C. , and G. Laffitte, 1981, Palinología de la cuenca del
great thickness (approximately 1000 m of organic-rich Noroeste. Argentina. I Características de las asociaciones
shales) is able to generate large volumes of hydrocar- carbónicas: problemas e interpretación: Octavo Congreso
Geológico Argentino, San Luis, Argentina, v. 4, p. 823–838.
bons. There is a strong relationship between this unit and Azcuy, C., G. Laffitte, and L. Rodrigo, 1984, El límite
the reservoired hydrocarbons. Outside the erosive edge Carbónico–Pérmico en la Cuenca Tarija–Titicaca: Tercer
of the Los Monos Formation (Figure 12), there are no Congreso Argentino de Paleontología y Bioestratigrafía, v.
commercial fields, in spite of the broader distribution of 1, p. 39–44.
the Lipeón and Icla mudstones. This is also reflected in Barret, S. , and P. Isaacson, 1989, Devonian paleogeography of
the organic content and degree of bioturbation. The South America, in N. J. McMillan, A. Embry, and D. Glass,
Kirusillas and Icla formations are generally more biotur- eds., Devonian of the World: Canadian Society of
bated and have a greater abundance of benthonic fauna Petroleum Geology Memoir 14, v. I, p. 655–669.
than the Los Monos Formation. This biotic activity has Beltan, L., S. Freneix, P. Janvier, and O. Lopez-Paulsen, 1987,
prevented the preservation of the organic matter in the La fauna triasique de la formation de Vitiacua dans la
région de Villamontes (Départament de Chuquisaca,
Kirusillas and Icla sediments. The general increase in Bolivie): N. Jb. Geol. Paläont. Mh. 1987, h. 2, p. 99–115.
preserved organic matter toward the end of the Coira, B., J. Dadvison, C. Mpodozis, and V. Ramos, 1982,
Devonian is a phenomenon that occurs worldwide and is Tectonics and magmatic evolution of the Andes of
attributed to a global anoxic event (Johnson et al., 1985). northern Argentina and Chile: Earth Science Reviews, v.
Reservoirs and traps span the entire stratigraphic 18, p. 303–332.
column where there are several levels of primary Cooper, M., 1986, Facies shifts, sea-level changes and event
porosity, as well as areas of secondary porosity formed stratigraphy in the Devonian of South Africa: South Africa
by fracturing. Although potential reservoirs and traps Journal of Sciences, v. 82, p. 255–288.
span the entire Silurian–Tertiary section, the best fields Comínguez, A., and V. Ramos, 1995, Geometry and seismic
are developed mainly in the Huamampampa and expression of the Cretaceous Salta rift system of north-
western Argentina, in A. J. Tankard, R. Suarez, and H. J.
Tupambi formations, both of which are in stratigraphic Welsink, Petroleum basins of South America: AAPG
contact with the Los Monos source rocks. The largest Memoir 62, this volume.
accumulations are related to north-south anticlines in the Cross, T., 1986, Tectonic controls of foreland basin subsidence
sub-Andean belt. The network of Tupambi paleovalleys and Laramide style deformation, western United States, in
forms stratigraphic traps in the Chaco Salteño. P. Allen and P. Homewood, eds., Foreland basins: Interna-
266 Starck

tional Association of Sedimentology Special Publication 8, Noroeste Argentino: Revista Instituto Geología y Minería,
p.15–39. Universidad Nacional de Jujuy, Jujuy, Argentina, v. 6,
Donato, E., and G. Vergani, 1985, Geología del Devónico y p. 13–28.
Neopaleozoico de la zona sur del Cerro Rincón, provincia Ramos, V., 1988, Late Paleozoic–early Paleozoic of South
de Salta, Argentina: Cuarto Congreso Geológico Chileno, America, a collisional history: Episodes, v. 2, p. 168–173.
Antofagasta, Chile, v. 1, p. 262–284. Ramos, V., T. Jordan, R. Allmendinger, S. Kay, J. Cortés, and
Eyles N., and C. H. Eyles, 1993, Glacial geologic confirmation M. Palma, 1984, Chilenia: un terreno alóctono en la
of an intraplate boundary in the Paraná basin of Brazil: evolución paleozoica de los Andes centrales: Noveno
Geology, v. 21, p. 459, 462. Congreso Geológico Argentino, v. 2, p. 84–106.
Eyles C., N. Eyles, and A. França, 1993, Glaciation and Rolleri, E., 1976, Sistema de Santa Bárbara: una nueva
tectonics in an active intracratonic basin: the Late Paleozoic provincia geológica argentina: Sexto Congreso Geológico
Itararé Group, Paraná basin, Brazil: Sedimentology, v. 40, Argentino, v. 1, p. 239–255.
p. 1–25. Ross, C., and J. Ross, 1988, Late Paleozoic transgressive–
Feruglio, E., 1929, Fósiles Devonicos del Quemado (San Pedro regressive cycles, in C. K. Wilgus, B. Hastings, H. Posa-
de Jujuy) en la region subandina del norte: Boletín de mentier, C. Ross, and C. Kendall, Sea-level changes—an
Informaciones Petroleras, Buenos Aires, Argentina, v. 15, integrated approach: SEPM Special Publication 42,
p. 65–73. p. 227–247.
França, A., and P. Potter, 1991, Stratigraphy and reservoir Salfity, J., R. Omarini, B. Baldis, and W. Gutierrez, 1975,
potential of glacial deposits of the Itararé Group (Carbonif- Consideraciones sobre la evolución geológica del Precám-
erous–Permian), Paraná basin, Brazil: AAPG Bulletin, v. brico y Paleozoico del norte argentino: Segundo Congreso
75, p. 62–85. Iberoamericano de Geología Económica, Buenos Aires,
González Bonorino, G., and F. González Bonorino, 1991, Argentina, v. 4, p. 341–361.
Precordillera de Cuyo y Cordillera Frontal en el Paleo- Salinas, E., J. Oblitas, and F. Vargas, 1978, Exploración del
zoico: terrenos “bajo sospecha” de ser autóctonos: Revista Sistema Carbonífero en la cuenca oriental de Bolivia:
Geológica de Chile, v. 18, p. 97–107. Revista Técnica Yacimientos Petrolíferos Fiscales Boli-
Hiller N., and J. Theron, 1989, Benthic communities in the vianos, La Paz, Bolivia, v. 7, p. 5–49.
south African Devonian, in N. J. McMillan, A. Embry, and Sempere, T., 1995, Phanerozoic evolution of Bolivia and
D. Glass, eds., Devonian of the World: Canadian Society of adjacent regions, in A. J. Tankard, R. Suarez, and H. J.
Petroleum Geology Memoir 14, v. 1, p. 229–242. Welsink, Petroleum basins of South America: AAPG
Isaacson P., and P. Sablock, 1990, Devonian palaeogeography Memoir 62, this volume.
and palaeobiography of central Andes, in W. McKerrow Sempere, T., E. Aguilera, J. Doubinger, P. Janvier, J. Lobo, J.
and C. Scotese, eds., Paleozoic Palaeogeography and Oller, and S. Wenz., 1992, La Formation de Vitiacua
Biogeography: Geologic Society Memoir 12, p. 431–435. (Permien moyen a superieur–Trias? inferieur, Bolivie du
Johnson, J., G. Klapper, and C. Sandberg, 1985, Devonian Sud): stratigraphie, palynologie et paléontologie: N. Jb.
eustatic fluctuations in Euroamerica: Geologic Society of Geol. Paläont. Abh. 185, h. 2, p. 239–253.
America Bulletin, v. 96, p. 567–587. Spalletti, L., C. Cingolani, R. Varela, and A. Cuerda, 1989,
Lopatín, N. V., 1971, Temperature and geologic time as factors Sediment gravity flow deposits of an Ordovician deep-sea
in coalification (in Russian): Izv. Akad. Nauk. SSSR, Seriya fan system (western Precordillera, Argentina): Sedimen-
Geologicheskaya 3, p. 95–106. tary Geology, v. 61, p. 287–301.
Lopez Gamundi, O., 1986, Sedimentología de la Formación Starck, D., E. Gallardo, and A. Schultz, 1992a, La discordancia
Tarija, Carbonífero de la Sierra de Aguaragüe, Provincia de precarbónica en la porción argentina de la Cuenca de
Salta: Revista Asociación Geológica Argentina, Buenos Tarija: Boletín de Informaciones petroleras, Buenos Aires,
Aires, Argentina, v. 41, p. 334–355. Argentina, Tercera Epoca, v. 29, p. 2–11.
Melo, J., 1989, The Malvinokaffric realm in the Devonian of Starck, D., E. Gallardo, and A. Schulz, 1992b, La Cuenca de
Brazil, in N. J. McMillan, A. Embry, and D. Glass, eds., Tarija: Estratigrafía de la porción argentina: Boletín de
Devonian of the World: Canadian Society of Petroleum Informaciones Petroleras, Buenos Aires, Argentina, v. 30,
Geology Memoir 14, v. 1, p. 669–704. p. 2–14.
Méndez, V., A. Navarini, D. Plaza, and V. Viera, 1972, faja Starck, D., E. Gallardo, and A. Schultz, 1993a, The pre-
Eruptiva de la Puna oriental: Quinto Congreso Geológico Carboniferous unconformity in the Argentine portion of
Argentino, v. 4, p. 147–158. the Tarija Basin: XII International Congress on Carbonif-
Monaldi, C., and M. Boso, 1986, Hallazgo de Dalmanitina erous–Permian, Buenos Aires, v. 2, p. 373–384.
(trilobita) en la Formación Zapla del noroeste argentino. Starck, D., E. Gallardo, and A. Schultz, 1993b, Neopaleozoic
Implicancia cronológica: Cuarto Congreso Argentino de stratigraphy of the Sierras Subandinas and Cordillera
Paleontología y Bioestratigrafía, Argentina, v. 1, p. 99–101. Oriental, Argentina. With comments on the southern
Oller, J., and T. Sempere, 1990, A fluvio-eolian sequence of borders of the Tarija Basin: XII International Congress on
probable Middle Triassic–Jurassic age in both Andean and Carboniferous–Permian, Buenos Aires, v. 2, p. 353–372.
Subandean Bolivia (abs.): International Symposium of Tankard, A. J., and J. H. Barwis, 1982, Wave-dominated
Andean Geodynamics, p. 237–240. deltaic sedimentation in the Devonian Bokkeveld basin of
Palma, M., and M. Irigoyen, 1987, Los estratos de Botijuela en South Africa: Journal of Sedimentary Petrology, v. 52,
la puna Catamarqueña: Décimo Congreso Geológico p. 959–974.
Argentino, Tucumán, Argentina, v. 2, p. 139–142. Tankard, A. J., M. A. Uliana, H. J. Welsink et al., 1995, Tectonic
Posamentier, H., G. Allen, D. James, and M. Tesson, 1992, controls of basin evolution in southwestern Gondwana,in
Forced regressions in a sequence stratigraphy framework: A. J. Tankard, R. Suarez, and H. J. Welsink, Petroleum
concepts, examples, and exploration significance: AAPG basins of South America: AAPG Memoir 62, this volume.
Bulletin, v. 76, p. 1678–1709. Theron, J., and J. Loock, 1989, Devonian deltas of the Cape
Ramos, V., 1986, El diastrofismo Oclóyico: un ejemplo de Supergroup, South Africa, in N. J. McMillan, A. Embry,
tectónica de colisión durante el Eopaleozoico en el and D. Glass, eds., Devonian of the World: Canadian
Silurian–Jurassic Stratigraphy and Basin Evolution of Northwestern Argentina 267

Society of Petroleum Geology Memoir 14, v. 1, p. 729–741. Author’s Mailing Address


Turner, J., 1979, Geología Regional Argentina: Academia
Nacional de Ciencias, Córdoba, Argentina, 1717 p. Daniel Starck
Uliana, M., K. Biddle, and J. Cerdán, 1989, Mesozoic extension Consultant
and the formation of Argentine sedimentary basins, in A. J. Avenida 53 # 1776 – (1900) La Plata
Tankard and H. Balkwill, eds., Extensional tectonics and Argentina
stratigraphy of the north Atlantic margins: AAPG Memoir
46, p. 599–614.
Vail, P., R. Mitchum, R. Todd, J. Widmier, S. Thompson, J.
Sangree, J. Bubb, and W. Hatleid, 1977, Seismic stratig-
raphy and global changes in sea level, in C. Payton, ed.,
Seismic stratigraphy—applications to hydrocarbon explo-
ration: AAPG Memoir 26, p. 49–212.
Van Wagoner, J., R. Mitchum, K, Campion, and V.
Rahmanian, 1990, Siliciclastic sequence stratigraphy in well
logs, cores, and outcrops: concepts for high-resolution
correlation of time and facies: AAPG Methods in Explo-
ration Series 7, 55p.
Veevers, J., and C. Powell, 1987, Late Paleozoic glacial
episodes in Gondwanaland reflected in
transgressive–regressive depositional sequences in
Euramerica: GSA Bulletin, v. 98, p. 475–478.
Visser, J., 1991, Geography and climatology of the Late
Carboniferous to Jurassic Karoo basin in south-western
Gondwana: Ann. South Africa Mus., v. 99, p. 415–431.
Vistalli, M., 1989, Cuenca Siluro Devónica del Noroeste, in
Cuencas sedimentarias argentinas: Publicación Especial
del Instituto Lillo, Tucumán, Argentina, p. 19–45.

Potrebbero piacerti anche