Download as pdf or txt
Download as pdf or txt
You are on page 1of 48

Structural and Tectonic Controls of Basin Evolution in

Southwestern Gondwana During the Phanerozoic

A. J. Tankard J. Skarmeta
Tankard Enterprises ENAP
Calgary, Alberta, Canada Santiago, Chile

M. A. Uliana H. Santa Ana


ASTRA ANCAP
Buenos Aires, Argentina Montevideo, Uruguay

H. J. Welsink F. Wiens
Perez Companc Geo Consultores
Neuquén, Argentina Asunción, Paraguay

V. A. Ramos M. Cirbián
Universidad de Buenos Aires
Buenos Aires, Argentina O. López Paulsen
YPFB
Santa Cruz, Bolivia
M. Turic
YPF
Buenos Aires, Argentina G. J. B. Germs
JCI Research Unit
Randfontein, South Africa
A. B. França

E. J. Milani M. J. De Wit
Petrobras/Nexpar University of Cape Town
Curitiba, Brazil Rondebosch, South Africa

B. B. de Brito Neves T. Machacha


Universidade de S. Paulo Geological Survey Department
São Paulo, Brazil Lobatse, Botswana

N. Eyles R. McG. Miller


Department of Geology NAMCOR
University of Toronto Windhoek, Namibia
Scarborough, Ontario, Canada

Abstract

T he continental lithosphere of southwestern Gondwana, comprising the southern part of South America
and southern Africa, was largely assembled before the end of the Proterozoic. Geologic studies indicate
that the basement anisotropy that controlled the development of Phanerozoic basins was established by
Neoproterozoic–Early Cambrian tectonism. This tectonism reactivated the older terrane boundaries or cut
across them. The backbone linking the system of Pan-African and Brasiliano basins was a system of northeast-
trending structures. There were four areas of pronounced Neoproterozoic–Early Cambrian basin subsidence in
the study area: the Chiquitanas trough in Bolivia, the Puncoviscana basin in Argentina, the Dom Feliciano-
Ribeira basins in Brazil, and the Damara-Nama basin complex of southern Africa.

Tankard, A. J., M. A. Uliana, H. J. Welsink, V. A. Ramos, M. Turic, A. B. França, E. J. Milani, B. B. de Brito 5


Neves, N. Eyles, J. Skarmeta, H. Santa Ana et al., 1995, Tectonic controls of basin evolution in southwestern
Gondwana, in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South America: AAPG
Memoir 62, p. 5–52.
6 Tankard et al.

The Phanerozoic basins developed largely along these Pan-Gondwana trends, their stratigraphy recording
periodic changes in tectonic behavior. Large volumes of terrigenous clastic sediments accumulated in the
Paleozoic Chaco-Tarija, Paraná, and Cape-Karoo basins. Major depositional hiatuses at the end of the Ordovi-
cian and the end of the Devonian are attributed to reorganization of stress fields and basin inversion. In
Argentina and Bolivia, these events are called the Ocloyic and Chañic orogenies, respectively. The succeeding
Permian–Carboniferous basins had an extensional component of subsidence. Tectonically formed relief is
believed to have initiated the various centers of Gondwana glaciation. Widespread inversion of Paleozoic
depocenters in the Late Permian and Early Triassic is expressed in the Cape-Ventana fold belt and the western
Pampean ranges. Inversion marks the transition to Mesozoic extension. There were two styles of Mesozoic
extension. In the first, a Basin and Range type of extension resulted from orogenic collapse during the
Triassic–Early Jurassic. Major offsets of the extensional tracts were accommodated by the preexisting northeast-
trending structures. In contrast, Late Jurassic–Cretaceous extension was controlled by northwestward-
diverging shear zones related to fragmentation of Gondwana, opening of the Atlantic Ocean, and trench roll-
back along the Pacific margin of South America. The climax of the Andean orogeny in the late Cenozoic
overprinted and reactivated these earlier trends.
The petroleum prospectivity of the South American basins is due largely to their composite form, multiple
intervals of Paleozoic and Mesozoic source rocks, and the modification or formation of structural traps in the
Tertiary. Clastic reservoirs and inversion structures are common.

Resumen

L a litósfera continental del extremo sudoeste de Gondwana que comprende la parte meridional de
Sudamérica y el sur de Africa fue ensamblada mayormente antes del fin del Neoproterozoico. Diversos
estudios geológicos señalan que las heterogeneidades del basamento que controlaron el desarrollo de las
cuencas fanerozoicas de dicha región fueron impuestas por el tectonismo del Proterozoico tardío y el Cámbrico
temprano. Ese tectonismo reactivó antiguas zonas de sutura o eventualmente se transmitió através de ellas. El
esqueleto que vinculó a este sistema de cuencas del Pan-Africano y Brasiliano se relaciona con una serie de
estructuras de orientación noreste. Dentro de la región en estudio hay cuatro áreas que registran pronunciada
subsidencia durante el Neoproterozoico tardio-Cámbrico temprano: la fosa Chiquitanas en Bolivia, la cuenca
Puncoviscana en Argentina, las cuencas Dom Feliciano-Ribeira en Brasil, y el complejo de cuencas Damara-
Nama de Africa meridional.
Las cuencas fanerozoicas se desarrollaron principalmente a lo largo de los lineamientos pan-gondwánicos, y
su estratigrafía registra cambios periódicos en el comportamiento tectónico. En el Paleozoico las cuencas Chaco-
Tarija, Paraná, y Cabo-Karoo acumularon grandes volúmenes de sedimentos clásticos. El desarrollo de hiatos
importantes hacia el fin del Ordovícico y del Devónico es atribuido a fenómenos de reorientación del campo de
esfuerzos y a inversión de las cuencas. En Argentina y Bolivia esos hiatos son referidos a las orogenias Oclóyica
y Cháñica, respectivamente. Las cuencas permo-carboníferas desarrolladas a continuación, registran un compo-
nente de subsidencia extensional. Se estima que varios de los focos de glaciación gondwánica iniciaron su
actividad a partir de altos topográficos inducidos tectónicamente. La inversión regional de los depocentros del
Paleozoico durante el Pérmico tardío y Triásico temprano es ostensible en el cinturón del Cabo-Ventana y en las
Sierras Pampeanas Occidentales. Dicha inversión marcó una etapa de transición a las condiciones de extensión
que dominaron al Mesozoico. Es posible reconocer dos estilos de extensión mesozoica. Al principio la extensión
asumió un estilo Basin-and-Range, derivado del colapso orogénico que tuvo lugar durante el Triásico y Jurásico
temprano. Variaciones en la ubicación de las áreas sometidas a extensión fueron compensadas a lo largo de
estructuras preexistentes de orientación noreste. De manera contrastante, la extensión jurásico-cretácica fue
controlada por zonas de cizalla divergentes hacia el noroeste, relacionadas a la fragmentación de Gondwana, a
la apertura del Océano atlántico, y a subducción retrocedente sobre el margen pacífico de Sudamérica. En el
Cenozoico tardío el clímax de la orogenia Andina se sobreimpuso y reactivo a las estructuras precedentes.
La prospectividad petrolera de las cuencas de Sudamérica se relaciona principalmente con su carácter
compuesto, con intervalos múltiples de rocas generadoras paleozoicas y mesozoicas, y con la modificación o
generación de trampas estructurales en el Cenozoico. Los reservorios clásticos y las estructuras de inversión son
rasgos muy frecuentes.
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 7

Figure 1—Southwestern
Gondwana showing the distribu-
tion of Neoproterozoic-Phanero-
zoic basins. The Paraná, Chaco,
and Chaco-Paraná basins origi-
nally formed a continuous basin
complex that was dismembered
by Mesozoic extension and
erosion. Regional cross sections
and seismic lines are identified
with their figure numbers. BH,
Boomerang Hills; CH, Chiqui-
tanas; DB, Damara belt; DF,
Dom Feliciano; GB, Gariep belt;
GC, Ghanzi-Chobe belt; KL,
Kalahari Line; PB, Paganzo
basin; 1, Otuquis well; 2, Masetl-
heng Pan-1 well. (Data from
Miller and Schalk, 1980;
Schobbenhaus et al., 1981;
Visser, 1984.)

INTRODUCTION This paper examines basin evolution in southwestern


Gondwana and attempts to demonstrate the linked
Southwestern Gondwana consisted of a collage of
nature of basin development. Interpretation is based on a
fault-bounded terranes that had accumulated around the
literature review, as well as on a large amount of previ-
Archean cratons, mostly before the end of the Protero-
ously unpublished surface and subsurface data,
zoic (Hartnady et al., 1985; De Wit et al., 1988; Unrug,
including wells and reflection seismic. Space does not
1992). The continental lithosphere was then deformed
allow an exhaustive survey of the literature. Instead, we
intermittently for at least 200 m.y. in the Neoproterozoic
hope that the diverse authorship will contribute a
and early Paleozoic. These Pan-African–Brasiliano events
breadth generally not available in the literature.
are expressed in a clastic- and carbonate-filled suite of
This is a first attempt to synthesize an enormous area.
sedimentary basins and associated intrusive rocks that
We realize that detailed paleostress measurements reveal
stretch across southern Africa, Brazil, Bolivia, Paraguay,
large-scale variations across a continent (e.g., Zoback and
Uruguay, and northern Argentina (Figure 1). Field
Zoback, 1980). This is probably inherent in Gondwana as
studies and exploration reflection seismic data show that
well, but there is insufficient paleostress data for rigorous
the Phanerozoic basins have repeatedly exploited preex-
analysis. We have taken the approach of estimating the
isting Neoproterozoic–Early Cambrian lineaments. For
effects of mean stress directions.
example, the Paleozoic Chaco-Tarija, Paraná, and Cape-
Azimuths given in this paper refer to a palinspastic
Karoo basins seem to have been tectonically linked
arrangement in which Africa is rotated relative to South
(Figure 1). They contain similar records of deposition
America. Present-day north arrows are shown for both
and subsidence and similarly exploited older crustal
continents.
zones of weakness.
8 Tankard et al.

Figure 2—Crustal architecture of


southwestern Gondwana. The
distribution and stratigraphy of
the Paleozoic basins (Visser,
1984; França et al., 1995; YPF,
proprietary reports) suggest that
this continental plate was largely
assembled by the earliest
Paleozoic. The rheology of the
terranes and their mechanical
boundaries influenced the distrib-
ution of Phanerozoic basins
(compare with Figure 3). The
composite Namaqua province
consists of the Bushmanland,
Gordonia, and Rehoboth terranes.
AC, Aconquija lineament; AMC,
Amazonia craton; CC, Congo
craton; KL, Kalahari line; LA, Luis
Alves craton; LB, Las Breñas;
NNMB, Namaqua-Natal mobile
belt; PB, Puncoviscana belt;
SCCB, southern Cape conduc-
tivity belt; SFC, São Francisco
craton; TL, Tandilia lineament;
TVL-PL, Tantalite valley
line–Pofadder lineament zone; ZF,
Zoetfontein fault. (Data from
Hartnady et al., 1985; Joubert,
1986; Ramos, 1988; Petrobras,
proprietary reports.)

PROTEROZOIC CRUSTAL EVOLUTION Mesoproterozoic basement is preserved in the Alto


Paraguay, Rio de la Plata, and Namaqua tectonic
The crustal evolution of southwestern Gondwana is
provinces (Figure 2). The Namaqua province is a high-
expressed in a mosaic of tectonic provinces and accre-
grade region that abuts the Kaapvaal craton (Tankard et
tionary terranes arranged around the Archean cratons
al., 1982) along the Kalahari line, which has a
(Figure 2). Several authors have interpreted the distribu-
pronounced gravity and magnetic signature (Hutchins
tion of mafic and ultramafic rocks and intracratonic
and Reeves, 1980; Corner, 1991). Abrupt crustal and
orogenic belts as the sutures that weld together the
lithospheric thickness changes occur across the Kalahari
various lithospheric blocks (Almeida et al., 1976; Cordani
line (De Wit et al., 1992) that may have influenced later
and Brito Neves, 1982; Hartnady et al., 1985; Joubert,
basin development. The Namaqua province comprises a
1986; Ramos, 1988; Unrug, 1992). The cratonic areas
diverse assemblage of terranes separated by mafic rocks
consist of several discrete crustal blocks of Archean–
and tectonized belts (Figure 3) (Thomas et al., 1993). This
Mesoproterozoic age that were probably allochthonous.
Bushmanland-Richtersveld terrane was accreted to the
These include the Amazonian, São Francisco, Luis Alves,
Kaapvaal craton about 1300 Ma (Joubert, 1986). The
Rio de la Plata, Congo, and Kalahari cratons. The
suture between the Gordonia and Bushmanland terranes
Archean Amazonian, São Francisco, and Kaapvaal
is a line of mafic rocks at least 1200 Ma that forms the
cratons have well-developed tectonic boundaries
Tantalite Valley line (TVL) along the Pofadder shear
(Hartnady et al., 1985; Ramos, 1988; De Wit et al., 1992).
zone. However, the number of terranes and their
Elsewhere the contacts are obscured by intracratonic
detailed thermotectonic history is still unresolved and
cover sequences such as the Chaco, Paraná, and Karoo
subject to much debate (Harris, 1992). The end of this
basins. Figure 3 shows the relationship among Phanero-
Mesoproterozoic orogenesis and stabilization of the
zoic basins, major structural trends, and potential field
Namaqua terranes enlarged the Kaapvaal craton into the
data. In South America, steep gradients are attributed to
Kalahari craton.
late Phanerozoic reactivation of terrane boundaries.
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 9

Figure 3—Potential field data sets and major


structural trends that define the basement
framework of southwestern Gondwana. South
American contours are Bouguer gravity;
Botswana data are magnetic. Steep gradients
are attributed to younger reactivation of old
terrane boundaries which are based on field
studies (compare with Figure 2). Note the
following: (1) The Kalahari magnetic anomaly
(KL) probably marks the edge of the Archean
Kaapvaal craton. (2) Steep Bouguer gradients
separate the Alto Paraguay and Rio de la Plata
cratons. The Paraná basin is associated with
negative Bouguer values, whereas the Chaco-
Paraná basin is a condensed cover above
normal crustal thicknesses without significant
variations (Wiens, 1995). Pronounced
northeast trends characterize the Paraná
basin. (3) The Chaco basin of Bolivia overlies
thinned crust. The Boomerang-Chiquitanas
transcurrent fault coincides with a distinct
gravity boundary, to the north of which the
Bouguer anomalies are typical of shield areas
(E. Ebner, 1993, personal communication). (4)
Near the Andes, the Bouguer gravity drops
significantly because there is no isostatic
correction. AC, Aconquija lineament; AP, Alto
Paraguay craton; BF, Boomerang fault; CH,
Chiquitanas front; Gord, Gordonia terrane; KL,
Kalahari line; LB, Las Breñas fault, LO, Lomas
de Olmedo structural trend; NC, Ncojane
basin; NO, Nosop basin; RP, Rio de la Plata
craton; SCCB, southern Cape conductivity
belt; TB, Trans-Brasiliano structure. (Data from
Hutchins and Reeves, 1980; Miller and Schalk,
1980; Visser, 1984; Hartnady et al., 1985;
Salfity, 1985; Joubert, 1986; Pitts et al., 1992;
Petrobras, proprietary reports.)

The Alto Paraguay terrane and Rio de la Plata craton Magnetic, gravity, and magnetotelluric data record
may have been tectonically linked to the Namaqua the possible southern margin of the Namaqua province
province in the Mesoproterozoic (Figure 2). Figure 3 beneath the Cape fold belt, where the southern Cape
compares the Bouguer gravity characteristics of the Rio conductivity belt (SCCB) coincides with the Beattie
de la Plata and Alto Paraguay terranes. The latter is a magnetic anomaly (De Beer, 1983; Pitts et al., 1992). This
broad gravity low where the Phanerozoic Paraná basin is margin is modeled as a southward-dipping structure
uncompensated at a subcrustal level. The northwest- containing dense material such as serpentinized mafic
trending transition is interpreted as a terrane boundary and ultramafic rocks attributed to ophiolite accretion.
or shear zone (Figure 2) (Ramos, 1988). Along the coast of The Saldania complex of metasedimentary and
eastern Uruguay, this lineament is closely associated magmatic rocks (Malmesbury, Kango, and Kaaimans
with mafic rocks (Brito Neves and Cordani, 1991). groups) abutting the SCCB margin of the Namaqua
Toward the northwest, the Rio de la Plata–Alto Paraguay province has been attributed to Neoproterozoic accretion
shear zone bifurcates around the Guaporé craton of (Hartnady et al., 1985). The SCCB-Beattie lineament may
Bolivia, a tectonic extension of the Amazonian craton have a counterpart on the conjugate Argentinian margin
(Almeida et al., 1976). The southern margin of the where it is broadly colinear with the Tandilia trend and
Guaporé craton was affected by two Mesoproterozoic its associated mafic rocks (Figure 2) (Ramos, 1988; Brito
orogenies: the San Ignacio orogeny at about 1300 Ma and Neves and Cordani, 1991).
the Sunsas-Aguapei mobile belt at 950 Ma (Litherland The Pampeanas terrane was joined to the Rio de la
and Bloomfield, 1981; Litherland et al., 1989; Unrug, Plata terrane in the Neoproterozoic along a westward-
1992). These events were broadly contemporaneous with dipping suture marked by mylonitized ultramafic bodies
Bushmanland-Kaapvaal coupling (Joubert, 1986), with magmatic arc affinities (Ramos, 1988). In Bolivia,
suggesting tectonic linkage. Outliers of Proterozoic this suture is believed to form the eastern margin of the
basement are exposed in the Andean belt as the Izozog arch (Figure 2), implying that Pampean basement
Arequipa-Antofalla massif (as old as 1918 Ma) (Ramos, underlies the Chaco basin. Geobarometry of metamor-
1988; Unrug, 1992). phic parageneses suggests as much as 15 km of colli-
10 Tankard et al.

sional uplift in the eastern Pampeanas terrane (Gordillo,


1984). Collision of the pre-Cordillera terrane with the N
gua ia
Pampeanas resulted in delamination and imbrication of -
Ar a
B
the crust (Cominguez and Ramos, 1991). The Sierra de

P a r a g u ay
TB
Valle Fértil fault system at the contact is a major crustal Tu c
P l aa v a c
structure that was repeatedly reactivated throughout the tf o a
rm

Phanerozoic (Baldis et al., 1982; Fielding and Jordan,


1988). The structure of the Pampeanas and pre-Cordillera na

a
is controlled by northeast- and northwest-striking faults

sc
Pu n c o vi
(Baldis et al., 1984; Salfity, 1985). Emplacement of the JF
Chilenia-Patagonia terrane complex largely completed
the construction of southwestern Gondwana. Ramos LF RB

(1988) and Dalla Salda et al. (1993) have commented on


Mesoproterozoic basement forming the Patagonia D
F

terrane.
When the Pampeanas, pre-Cordillera, and Chilenia-
Patagonia terranes were joined is contentious. For

fts
Ri
example, De Wit and Ransome (1992) believe that part of sib
the Patagonia terrane may have been sutured during the No
Cape orogeny in the middle Paleozoic. The eastern OL
MA
Sierras Pampeanas preserves a 1200-km-long belt of Damara
magmatic rocks with a 700–600 Ma age range (Ramos, Gariep
1988). Blanketing stratigraphic relationships suggest that
the pre-Cordillera and Pampeanas were already in KL
Saldania

N
contact by the earliest Paleozoic, although large amounts
of right-lateral displacement may have affected these
blocks in the Paleozoic (Aceñolaza and Toselli, 1988). The
stratigraphic and deformational records (e.g., Ocloyic
and Chañic events) provide no evidence for distin-
R
guishing between initial accretion of the Chilenia terrane
and transpressional inversion of an amalgamated exten- R’

sional complex. Likewise, a middle Paleozoic age for


suturing of the Patagonia terrane has been inferred from
the Somuncura batholith. However, Rapela and Kay
(1988) argue that there is no evidence of a suture and that
the siliceous character of the Somuncura batholith is
atypical of subduction. Riccardi and Rolleri (1980) Figure 4—Distribution of major outcrop belts of Neopro-
describe the Bahia La Lancha suture of southern terozoic (Riphean–Vendian) rocks. In Namibia, early
Patagonia, which they attribute to a Carboniferous-age Damara extension created the east-west oriented Nosib
collision. rifts. The Matchless belt is a 350-km-long volcanic-plutonic
The present study examines the stratigraphic relation- belt attributed to Nosib extension which culminated in
ships, paleogeographies, and tectonic framework of separation and formation of oceanic-like crust (Breitkopf
southwestern Gondwana and suggests that these major and Maiden, 1987; Stanistreet et al., 1991). In eastern Brazil,
mafic and granitic intrusions in the Dom Feliciano and
terranes were largely joined by the end of the Protero-
Ribeira belts are associated with northeast-trending struc-
zoic. This interpretation does not preclude significant tures (e. g., Braun et al., 1991). Counterparts of these rocks
reorganizations throughout the Phanerozoic. in Bolivia form the Tucavaca platform. Extensional
In summary, when results from either side of the processes were ubiquitous. B, Boomerang; DF, Dom
Atlantic are combined, they demonstrate a unity in the Feliciano; JF, Jacutinga fault; LF, Lancinha-Cubatão fault
pattern and timing of basement construction. The zone; MA, Matchless amphibolite; OL, Okahandja
systematic pattern of terrane accretion imposed a first- lineament; RB, Ribeira belt; TB, Trans-Brasiliano
order northwest-trending structural grain, as well as a lineament.
secondary intraterrane structural fabric (e.g., Harris,
1992). Northwest and northeast trends are common in
the Gordonia-Rehoboth, Alto Paraguay, and Pampeanas
terranes. Neoproterozoic–Phanerozoic basin develop- northeast-striking strike-slip faults or accommodation
ment reflects the relative strengths of the lithosphere. The zones that cut across terrane boundaries was a major
Gordonia-Rehoboth-Alto Paraguay lithosphere formed control of basin development. Only in the Late Jurassic
major depocenters, as did the northern Pampeanas and Cretaceous did the Proterozoic sutures directly
terrane. The Archean cratonic nucleii, including the control tectonism. It is as if fragmentation of south-
Guaporé, Rio de la Plata, and Kaapvaal cratons, were western Gondwana in the late Mesozoic attempted to
generally positive elements. A conspicuous set of dismantle the Proterozoic construction.
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 11

PAN-GONDWANA TECTONIC Riphean Extension


FRAMEWORK Extensional subsidence was ubiquitous in the Riphean
(Figure 4). The Damara succession of Namibia was
The Pan-African and Brasiliano belts form a network initially deposited in a suite of disconnected rift basins
of Neoproterozoic–Cambrian basins and orogenic between the Congo and Kalahari cratons. Alkaline
systems which span the African and South American magmatism dates this intracratonic rifting at 830–730 Ma
continents (Kröner, 1977; Almeida et al., 1981; Miller, (Miller, 1983). These Nosib rift basins were almost 7 km
1983; Brito Neves and Cordani, 1991; De Wit and deep in places (SACS, 1980). By late Nosib and Swakop
Ransome, 1992). They record a complex structural and time (Table 1), partial breakup of the Khomas trough is
metamorphic history and several episodes of deformation believed to have created a narrow band of oceanic floor
that surround the cratons. This Pan-Gondwana episode reflected in the 350-km-long strip of Matchless serpenti-
was initiated about 960 Ma ago and locally persisted nite, gabbro, and metabasite (Figure 4) with its MORB-
through Early Cambrian time (Miller, 1983; Brito Neves type geochemical affinities (Finnemore, 1978; Miller,
and Cordani, 1991). Table 1 summarizes the Riphean– 1983; Breitkopf and Maiden, 1987), pillow lavas (Sawyer,
Cambrian stratigraphy and its distribution. Age and 1981), and pelagic sedimentary rocks (Miller, 1983;
stratigraphic relationships, however, are generally poorly Stanistreet et al., 1991). Other conspicuous rock types
constrained, and basin development was probably include the Chuos (Nosib) and Numees diamictites.
diachronous. Gresse and Scheepers (1993) review the Postrift thermal subsidence created a broad epeiric
Pan-African geochronology of southern Africa in detail. basin above the previous rifts, resulting in a blanket of
The principal Pan-African–Brasiliano depocenters are carbonate platform and argillaceous sediments. A
located along prominent intracratonic structural belts. postrift Swakop succession up to 15 km thick accumu-
Forming the core is a system of northeast-trending linea- lated above the northern Damara rift. The loci of upper
ments and shear zones associated with the São Francisco, crustal extension (rift stage Khomas trough) and lower
Ribeira, Dom Feliciano, and Gariep belts (Figures 1, 4). crustal extension (postrift stage northern Damara
The Trans-Brasiliano lineament and northeast-directed Swakop cover) were offset, suggesting that extension
Bouguer gravity trends emphasize this fabric in the São may have been accommodated by intracrustal detach-
Francisco basin (Braun et al., 1991; Brito Neves and ment (see Kusznir and Egan, 1989; Henry et al., 1990).
Cordani, 1991). In Bolivia, the Chiquitanas front is a The Gariep belt (Figure 4) is essentially a coastal
shear zone that diverges along the edge of the Guaporé continuation of the Damara and preserves a similar
craton (Litherland et al., 1989). Likewise, the inland geologic history, including an extensional phase
branch of the Damara belt occurs between the Congo dominated by alluvial fans. In the west, allochthonous
and Kalahari cratons (Figure 4). The Puncoviscana belt of ophiolite terranes are juxtaposed at a major fault
northwestern Argentina lies along the western edge of (Hartnady et al., 1985; Von Veh, 1990). The Gariep has
Neoproterozoic Gondwana. affinities with the highly tectonized Dom Feliciano–
An aeromagnetic survey in Botswana (Figure 3) shows Ribeira conjugate margin of Brazil (Almeida et al., 1981;
a considerable depth to magnetic basement west of the Porada, 1989; Braun et al., 1991; Brito Neves and
Kalahari anomaly (Hutchins and Reeves, 1980). This Cordani, 1991; Stanistreet et al., 1991; Unrug, 1992).
geology was tested with acquisition of 20-sec reflection The São Francisco and Paraná basins of Brazil appear
seismic data in the Nosop-Ncojane basin (Geologic to have followed a similar pattern of basin evolution
Survey Department, 1987–1988, proprietary seismic data). between about 730 and 560 Ma (Braun et al., 1991). (In
A very thick succession of well-defined and persistent this paper, we use the term Paraná basin to refer to its pre-
reflectors (Figures 5, 6) confirm basin depths in excess of Ordovician genetic precursors as well as its Paleozoic
15 km. The Masetlheng Pan exploration well (Figure 1) depocenters.) The São Francisco basin contains the
recovered late Riphean palynomorphs at 2 sec (~4000 m) Bambui Group, a variable assemblage of terrigenous
beneath a Nama section (A. Knoll, 1994, personal commu- clastics and calcareous rocks (Table 1) (Dardenne, 1978).
nication). Hall et al. (1990) have shown that the basin- Initial sedimentation of coarse alluvial clastics occurred
forming structures have a long history of repeated reacti- along northeast-trending structures (e.g., Sete Lagoas
vation, and they interpret the Kalahari magnetic anomaly Formation) (Braun et al., 1991; Brito Neves and Cordani,
as an old fault system. Hall et al. (1990) estimate crustal 1991). Similar northeastward structural trends were also
thickness to be about 40 km. Refraction seismic data important in the Riphean Paraná basin (Brito Neves and
record a similar crustal thickness beneath the northern Cordani, 1991; Petrobras proprietary reports).
Damara cover (42 km) (Green, 1983). In Bolivia, the Chiquitanas front forms the wrench
The various Pan-African and Brasiliano basins shared margin of the Guaporé craton. It formed by reactivation
broadly similar patterns and timing of basin develop- of the older San Ignacio (1300 Ma) and Sunsas (950 Ma)
ment and deformation, reflecting tectonic linkage. The belts (Litherland et al., 1989). The stratigraphic column of
initial stages of basin evolution were extensional or the Chiquitanas margin consists of three groups (Table
transtensional and locally culminated in creation of 1): (1) the Boqui extensional margin, on which the
oceanic-type crust. After a period of rift abandonment, Tucavaca platform of (2) Corumbá carbonates and (3)
widespread contraction, shearing, and associated Tucavaca fine-grained terrigenous clastics were
magmatism closed this Pan-Gondwana cycle. deposited. The Boqui consists of as much as 2000 m of
12 Tankard et al.

Table 1—Comparative Vendian–Cambrian Stratigraphy

(Data from SACS, 1980; Almeida et al., 1981; Germs, 1983; Miller, 1983; Litherland et al., 1989; Braun et al., 1991; Brito Neves and Cordani, 1991; J. Salfity, 1993, personal
communication, França et al., 1995.)

channelized arkoses attributed to alluvial deposition. (Braun et al., 1991; Unrug, 1992). The Matchless metaba-
About 1700 m of limestones with clastic interbeds were site in the inland Damara extensional tract is interpreted
drilled in the Otuquis well (Figure 1) (YPFB, proprietary as a sliver of oceanic crust (Kukla and Stanistreet, 1991),
report). Outcrop geology along the Sierras Chiquitanas although paleomagnetic data would appear to preclude
and seismic reflection data show that pre-Chaco sedi- formation of a wide ocean basin (McWilliams and
mentation occurred in rift basins (Figure 7). Seismic data Kröner, 1981) such as the “Adamastor ocean” of
show hundreds of meters of throw on the rift-bounding Hartnady et al. (1985).
faults, some of which had a component of oblique slip
movement. (foldout, opposite)
Glacial diamictites in the Bolivian Boqui Formation Figure 5—GSD deep seismic reflection profile 9493. Unin-
(Lopez, 1982) and the Brazilian Puga and Jequitaí forma- terpreted (top) and interpreted (bottom) sections showing
tions may be counterparts of the Chuos of Namibia Ghanzi thrusts rooted in pre-Nama basement. The shallow
reflectors above 4 sec are horizontal, implying little flexural
(Alvarenga and Trompete, 1992).
deformation in front of the Ghanzi-Chobe thrust belt.
In summary, Riphean extension was widespread and P, inferred base of Proterozoic; B, reflective lower crust.
lasted 180 m.y. or more, during which the thickest See Figure 10 for location.
Neoproterozoic sediments accumulated. Coarse terrige-
nous sediments marked the climax of rifting, but were (foldout, reverse side)
superseded by finer grained clastics and carbonates. The Figure 6—GSD deep seismic reflection profile 9492. Unin-
Chiquitanas and inland branch of the Damara notwith- terpreted (top) and interpreted (bottom) sections showing
standing, a pervasive northeast-trending structural fabric a considerable thickness of layered rocks, normal faults,
was a major influence on basin development. This is and apparent structural inversion which has produced
shown by field studies as well as Bouguer gravity trends reverse offsets of reflectors. Masetlheng Pan-1 is projected
onto this profile. The well was drilled to 2 sec. Nama sedi-
in South America (Braun et al., 1991). Interior belts of
mentary rocks occur at 0. 87–1. 34 sec, with upper Riphean
flysch, volcaniclastic materials, and bimodal suites of below that (A. H. Knoll, 1994, personal communication). N,
volcanics and calcalkaline granitoids (850–500 Ma) are base of Nama; P, inferred base of Proterozoic; B, reflective
associated with these lineaments and shear zones: Trans- lower crust (reflector “e” of Hall et al., 1990). See Figure 1
Brasiliano lineament, Dom Feliciano, Ribeira, and Gariep for location.
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 13
14 Tankard et al.
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 15

Transtensional processes appear to have been Kalahari sands in Botswana. The Nama spans about 50
important, locally resulting in considerable crustal atten- m.y. from 580 to 530 Ma and is thus synorogenic
uation and even creation of small pull-apart type oceanic (Kaufman et al., 1993). Crystallinity measurements on
basins (e.g., Miller, 1983; Breitkopf and Maiden, 1987; white micas suggest that temperatures in the upper
Brito Neves and Cordani, 1991; Stanistreet et al., 1991). Nama may have exceeded 200°C.
There is no evidence for a larger scale Wilson cycle. The classic Nama section in Namibia is divided into
However, this remains a lively debate (see Miller, 1983). three packages (Germs, 1983; Gresse and Germs, 1993).
A possible analog for these tectonic associations is the The Kuibis and Schwarzrand subgroups consist of mixed
west Peruvian trough (McCourt, 1981) or the Cariaco limestones and terrigenous clastics. Arkosic sandstones
trench of Venezuela. and mudstones of the Fish River Subgroup cap this
succession. Phycodes pedum near the top of the
Vendian–Early Cambrian Orogenesis Schwarzrand is believed to mark the Proterozoic–
Cambrian transition (Germs, 1974, 1983) at approxi-
By late Riphean time, the basins were subjected to mately 540 Ma (Kaufman et al., 1993).
widespread contraction, except for the Puncoviscana belt The geometry of the Nama foreland basin resulted
at the edge of the continent. Two structural trends, north- from flexural adjustment to the load applied by the
eastward and approximately northward, accommodated Damara mountain belt and reactivation of foreland struc-
much of the contraction in Brazil and Namibia. Together tures. Figures 8 and 9 show the relationship of Nama
these created linear tracts of transtensional basins. Reacti- stratigraphy to basement-involved faults. The principal
vation of older structures by compression influenced faults were derived from 1:1,000,000 geologic maps
sedimentation in the foreland basins. Mixed clastic and (Miller and Schalk, 1980). Northeast- and northwest-
carbonate deposition characterized this interlude, while trending faults predominate. Although many of these
magmatism persisted along many of the structures. faults were active during the middle Paleozoic (Münch,
Deposition of the Mulden sediments started at about 1974), considerable evidence suggests a Damara-age
650 Ma (Table 1) (Miller, 1983). These deposits were origin (Miller, 1980). On a regional scale, the geometries
deformed at 600 Ma (Lombard et al., 1986) and are thus of unconformity-bounded sequences and the distribu-
older than the lower Nama Group. tions of stratigraphic units reflect basement structure
The Nama Group is a synorogenic cover sequence (Figure 9). For example, onlapping Kuibis stratigraphy
that is relatively undeformed except along its margins, was folded and deeply eroded during the inversion of
notably in the Naukluft klippe complex. It is covered by the Koedoelaagte arch in earliest Nomtsas time (~590
16 Tankard et al.

Ma) (Gresse and Germs, 1993, their figure 3). Unconfor-


mities within the succession were eroded deeply along
the basin margins or across prominent basement highs
such as the Osis ridge or Koedoelaagte arch. Lower Fish
River isopachs and paleocurrent azimuths (Germs, 1983)
match the structural framework (Figure 8), suggesting
that these faults were active during sedimentation.
Basement-controlled faulting has also localized patch
reef buildups along the margin of the Witputs
depocenter (Figure 10A) (Germs, 1983).
In the Nosop-Ncojane basin of Botswana, deep
seismic data show insignificant flexural deformation at
Nama levels (0.9–1.3 sec) (Figure 5). In contrast, the
strike-oriented seismic profile (Figure 6) shows reactiva-
tion and gentle inversion of pre-Nama normal faults.
Masetlheng Pan was drilled on a structural culmination
created by inversion. There is a prominent intra-
Schwarzrand unconformity (GSD proprietary report)
that broadly coincides with the uplift across the Osis and
Koedoelaagte arches in basal Nomtsas time (Gresse and
Germs, 1993). Hall et al. (1990) have interpreted the
Kalahari magnetic anomaly along the eastern margin of
the Nosop-Ncojane basin as an ancient fault system.
Figure 10A shows the distribution of probable Nama
counterparts in the southwestern part of South Africa.
Germs and Gresse (1991) have equated the Vanrhyns-
dorp Group with the Nama; folds and thrusts verge
northeastward. Farther south, the Franschhoek and
Klipheuwel formations accumulated in northwest-
striking fault-bounded rift basins. Both have thick accu-
mulations of angular conglomerates, sandstones,
mudstones, and calc-alkaline detritus (De Villiers et al.,
1964) attributed to alluvial fan, debris flow, and lahar
deposition. The conglomerates are abundant near highly
Figure 8—Outcrop of Vendian–Cambrian Nama strati- sheared and mylonitized granites. These facies associa-
graphy and major faults. Isopach geometry and paleocur- tions and mylonitized faults suggest a transtensional
rent azimuths for the Stockdale Formation (Fish River pull-apart basin setting. Similar structural styles charac-
Subgroup) broadly match the present structural terize the Vanrhynsdorp basin (Gresse, 1986).
framework, suggesting sedimentation and structural Damara orogenesis closed the Khomas trough, culmi-
linkage in these Pan-African basins. (Modified after Miller
and Schalk, 1980; Germs, 1983.) See Figure 1 for location.
nating in the Nama foreland basin and Naukluft nappe
complex (Hartnady, 1978, measured 78 km of nappe

Figure 9—Nama basin fill. The


Osis ridge is attributed to
contractional uplift of a
basement horst. See Figure 8 for
location. (Modified after Germs
and Gresse, 1991.) OR, Osis
ridge; KA, Koedoelaagte arch.
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 17

N
NA MIB I A BO TS WANA

94
GC

93

KL
MA
MP 2 ZF
NN 4
9
V 9

Z NO S O P
NA M A
BA S IN BA S IN
OR
T

W
K
SO U T H AFRI CA

KF

R’
VAN R H Y N SD O R P
R
Nam a O ut cr op
BA S IN
20 s e c Se is mi c
Pa tch R ee f
Exp lo ra tio n We ll 0 km 500
KH
Oi l Se e p
WF CF
0 km 500
M

A B

Figure 10—Pan-African geology and interpretation. (A) Map showing principal faults and basins of Neoproterozoic–
Cambrian age (after Miller and Schalk, 1980; Visser, 1984). Nama depocenters: K, Karas, W, Witputs; Z, Zaris. G-C, Ghanzi-
Chobe thrust belt; KF, Kheis fault zone; KL, Kalahari line; KH, Klipheuwel basins; M, Malmesbury; MA, Matchless amphibo-
lite; MP, Masetlheng Pan-1 well; NN, Naukluft nappe complex; OR, Osis ridge; T, Aquitaine Tses-1 well; V, Artnell Vreda-281
well; ZF, Zoetfontein fault. (B) Tectonic interpretation of the late Pan-African basins. CF, Kango fault; WF, Worcester fault;
large arrows, maximum horizontal paleostress inferred from fault slips.

transport). Deep seismic data show little flexural defor- principal compressive stress σ1 was oriented toward the
mation in front of the Ghanzi-Chobe belt, although reac- northwest (see Hoffmann, 1984). The northwest-trending
tivation of northwest-oriented structures did occur faults were either parallel or at a low angle to the axis of
(Figures 5, 6). maximum compressive stress and were thus not suscep-
On the basis of these interpreted stratigraphic and tible to large amounts of reactivation (e.g., gentle Masetl-
structural relationships, Figure 10B attempts to recon- heng inversion). The orientation with respect to σ1 of the
struct the late Pan-African tectonic setting for the western intrabasinal Osis ridge suggests that it was contractional
part of southern Africa. Geologic maps (Miller and in origin, possibly as an uplifted basement horst.
Schalk, 1980) show that the inferred northeast-trending Figure 11 shows the distribution of upper Brasiliano
accommodation zones were not necessarily through- stratigraphy and its structural context in the southern
going but rather formed en echelon strands in which part of South America. In the São Francisco and Paraná
displacement was relayed from one zone to the next. basins, northeast-oriented structural trends were ubiqui-
Reactivation of basement structures is typical of foreland tous. Important examples are the Trans-Brasiliano,
basin subsidence. Balkwill et al. (1995) and Mathalone Jacutinga, Lancinha-Cubatão, and Blumenau-Soledade
and Montoya (1995) show the seismic expression of such fault zones (Zalán et al., 1991). Associated with these
structures in the Andean foreland basin. The interpreta- fault zones are a secondary set of north-northeast
tion shown in Figure 10B implies that the maximum trending faults. (Note that the palinspastic arrangement
18 Tankard et al.

Figure 11—Neoproterozoic–Early
Cambrian tectonic reconstruc-
VENDI AN - CAMBR IA N tion in which contraction and
structural inversion modified
N earlier Pan-African–Brasiliano
BF
TB paleogeography (see Figure 4).
CH
The conjugate Alto Paraguay and
Rehoboth terranes (Figure 2)
TON
IC supported similar patterns of

GF
CA CRA O RE
C subsidence and alluvial sedimen-
ET tation. The northeast-trending
OF
structures cross-cut the older
PF JF terrane boundaries and were a
major control of Phanerozoic
RI B E I RA
AC LB
TC LF basin development. AC,
LA
Aconquija lineament; BF,
Boomerang fault; BS, Blumenau-
BS
Soledade fault; CA, Apa craton;
DOM NO
FEL
ICIA CH, Chiquitanas front; DB,
Damara belt; ET, El Toro
lineament; GB, Gariep belt; GF,
Goiânia belt; JF, Jacutinga fault;
DB R
KL, Kalahari line; LA, Luis Alves
M
A R’
craton; LB, Las Breñas; LF,
Lancinha-Cubatão fault; MA,
Matchless amphibolite; NB,

N
NB

GB Nama basin complex; OF,


GH
AN Ocloyic fault; PF, Puna fault; TB,
ZI
BE - C H Trans-Brasiliano lineament; TC,
VB LT O B
E Tebicuary craton; VB, Vanrhyns-
KL
dorp basin; ZF, Zoetfontein fault.
ZF (Data from Miller and Schalk,
1980; Schobbenhaus et al., 1981;
U. Ve n dia n - L. C am br ia n
Cla stic s Brito Neves et al., 1984; Visser,
Rip he a n - Ven d ian Cla s tic s 1984; Salfity, 1985; Chebli et al.,
an d C ar b o na te s 1989; Pezzi and Mozetic, 1989;
Ve nd ia n - C amb r ia n P lu ton s Willner et al., 1987; Litherland et
an d B a sic Ig ne o us Ro c k s al, 1989; Braun et al., 1991; Brito
Pe tr ol eu m E xp lo r atio n Neves, 1991; Zalán et al., 1991;
Well s
França et al., 1995; Wiens, 1995;
Oil S ee p
Petrobras, proprietary reports.)

in Figure 11 shows Africa rotated with respect to South releasing bends along the shear systems, while contem-
America, so that the Paraná north-northeast trends are poraneous Brasiliano folding occurred at restraining
parallel to the northwest extensional azimuths of the bends. The Trans-Brasiliano and Ribeira–Dom Feliciano
Nama basin in Figure 10.) These fault zones and the magmatic provinces are associated with this left-lateral
cross-cutting trends encapsulate the Ribeira and Dom shear regime. These lineaments consist of anastomosing
Feliciano belts and their syntectonic calc-alkaline deep shear systems containing mylonitic belts up to 2 km
magmatic rocks (Bernasconi, 1987; Braun et al., 1991; wide (Almeida and Hasui, 1984).
Brito Neves and Cordani, 1991). These regional stress fields reacted with the Chiqui-
Late Brasiliano rifts developed along these cross- tanas-Boomerang suture in a right-lateral transpressional
cutting structural trends in the Paraná basin. They sense. Deposition was restricted to a thin blanket of
appear to be associated mainly with Alto Paraguay Tucavaca argillites. Nevertheless, the Chiquitanas shear
basement (compare Figures 2, 3, 11), possibly a counter- zone (the San José wrench fault of Sempere, 1995)
part of the Rehoboth terrane which underlies the Nama relayed strain between the Brasiliano-age belt of the
depocenters (Figure 2). The Itajai cycle is represented by Peruvian Andes (Dalmayrac et al., 1980), the Puncovis-
half-grabens and transpressional foreland basins filled cana belt of the Puna and eastern Cordillera of
with conglomerates, sandstones, and rhyolites (Macedo Argentina, and the Paraná basin.
et al., 1984; Rostirolla et al., 1992). Brito Neves and The Pampeanas and pre-Cordillera terranes of north-
Cordani (1991) interpret the rifts as pull-apart basins that western Argentina were dissected by northwest- and
they relate to northeast-oriented wrench tectonics. northeast-trending faults (Baldis et al., 1984; Salfity, 1985)
The development of these pull-apart basins with felsic that were repeatedly reactivated throughout the
to intermediate magmatism (545 Ma) is attributed to Phanerozoic. Here the Puncoviscana belt accumulated
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 19

several kilometers of flysch sediments in a set of exten- farm Heigums No. 105 (Figure 11). These oils are in close
sional basins (Figures 1, 11) (Omarini and Baldis, 1984; association with black Kuibis limestones. The oil is
Jezek and Miller, 1987). Aceñolaza and Durand (1986) dominated by acyclic isoprenoids and is slightly biode-
date these rocks as Vendian to mid-Cambrian. Facies graded. The predominance of the C29 steranes and their
relationships show that the Ocloyic and Puna faults abundance relative to the C28 steranes, absence of C30
behaved in a normal sense during deposition of the 4-desmethylsteranes, and the presence of methyl
Lerma Group and were dissected by the El Toro and steranes are characteristic of Neoproterozoic–early
Aconquija lineaments (J. Salfity, 1993, personal commu- Paleozoic oils. Furthermore, the low ratio of pristane to
nication) (Figure 11). Syntectonic and posttectonic grani- phytane, the high abundance of long-chain acyclic
toids (489–530 Ma) were emplaced in the closing orogeny isoprenoids and gammacerane, and the presence of
when the passive margin sediments were folded (Willner 8,14-secohopanes are indicative of carbonate source rock
et al., 1987). Puncoviscana counterparts occur in the deposited under highly anoxic conditions (M. Fowler,
southern sub-Andean foothills of Bolivia and appear to 1991, personal communication). The sample is mature
continue northward to the Peruvian Andes where and at the peak of oil generation. This Kuibis seep appar-
basement ages of about 600 Ma are recorded (Dalmayrac ently influenced drilling of Tses-1 by Aquitaine in 1971
et al., 1980). and Vreda-281 by Artnell Exploration in 1963–64. Both
dry wells drilled Nama sections (Lawrence, 1989).
Interpretation In the Etosha basin of northern Namibia, oils have
been sourced from Vendian limestones and are slightly
Pan-African–Brasiliano events established the broad- biodegraded (McKirdy et al., 1983). In terms of biomarker
scale tectonic framework that guided the evolution of distributions, however, the Etosha and Nama oils are
Phanerozoic successor basins until the Cretaceous. (We distinctly different from each other. Gas seeps in the São
use the term successor basin to emphasize the structural Francisco basin and oil and gas seeps in the Corumbá
inheritance of a basin.) There is an overall simplicity in limestones near the Bolivia-Brazil border are also known
the way these basins evolved, probably because they (Dardenne, 1978; Braun et al., 1991; França et al., 1995).
were geodynamically linked. In their early history, these
basins were mostly extensional (Figure 4) and consider-
able thicknesses of sediments were deposited. From late PALEOZOIC SUCCESSOR BASINS
Riphean through Early Cambrian time, compressive
deformation altered the paleogeography (Figure 11). An The Phanerozoic basins evolved by exploiting much
important consideration is that the apparent linkage of the preexisting structural fabric. There are several clear
throughout Pan-African–Brasiliano evolution precluded examples (Figure 12). First, the old Puncoviscana belt of
continental-scale ocean opening and closing between northwestern Argentina hosted a variety of basin types,
South America and Africa. However, there is speculation for example, Permian–Carboniferous pull-apart basins
that rifting and seafloor spreading between Laurentia (Fernandez-Seveso and Tankard, 1995) and Cenozoic tilt
and Gondwana preceded the Famatinian orogeny of block basins (Fielding and Jordan, 1988). Second, the
northwestern Argentina (Dalla Salda et al., 1992; Dalziel Bolivian Chaco basin subsided along the Boomerang-
et al., 1994). Chiquitanas transcurrent shear system that is expressed
The tectonic backbone of this system comprised a in a distinct Bouguer gravity boundary (Figure 3). Third,
suite of northeast-oriented lineaments and shear zones. the northeast-oriented lineaments and shear zones in
Associated crustal attenuation locally resulted in creation Brazil largely controlled Cambrian–Ordovician depocen-
of small pull-apart oceanic basins, such as the Khomas ters (França et al., 1995).
trough, and possibly some of the Dom Feliciano and The principal Paleozoic depocenters were the Chaco-
Ribeira magmatic belts, which are believed to have had Tarija basin, the Paraná basin, and the Cape-Karoo basin
magmatic arc affinities (e.g., Brito Neves and Cordani, (Figure 1). Depth to basement below the sub-Andean
1991). The latter also contains a complex system of fold and thrust belt of the Chaco basin is at least 12 km.
nappes and thrust slices. Likewise a plutonic arc is asso- The Cape and Karoo basins accumulated up to 9 km of
ciated with the Trans-Brasiliano structure. Elsewhere, sediment in a fault-controlled trough. In contrast, the
dilation at releasing bends is associated with acid to Chaco-Paraná basin consists of a condensed cover
intermediate volcanics and granitoids. Transtension sequence above normal thicknesses of crust; positive
along the Paraná shear zones is reflected in a parallel set Bouguer trends show no significant variations in crustal
of rifts. thickness (Figure 3) (Wiens, 1995). During the Paleozoic,
The Chiquitanas-Boomerang suture appears to have the Chaco, Chaco-Paraná, and Paraná basins formed a
functioned as an accommodation zone that relayed strain continuous basin complex that was later dismembered
from the Puncoviscana belt near the active boundary of by Mesozoic extension.
Gondwana to the continental interior (Litherland et al., Figure 13 compares the stratigraphic columns for the
1989). Daly et al. (1989) have suggested that the principal depocenters. Overall, there was a common
northeast-striking branch of the Damara was linked to pattern of stratigraphic evolution, at least through the
the Mozambique belt in a similar way. Triassic. Some events are diachronous, reflecting the
A curiosity of the Nama basin is that oil occurs in nature of basin evolution as well as the vagaries of bio-
brecciated Kuibis sandstones along a fault zone on the stratigraphic analysis.
20 Tankard et al.

Figure 12—Cambrian–Ordovi-
cian paleogeography in which
CAMBRI AN - ORDOV IC IA N structural inheritance has
guided basin evolution. Uplift
N BF (inversion) along the Sierras
CH Pampeanas created a high
CH AC O that separated early Paleozoic
basins on either side of it.
Subsidence was most
KF
PU N A
pronounced east of this high
P UN C O

(Ventana-Cape basin complex)


and adjacent to the
V IS

PAR A NA
Boomerang-Chiquitanas shear
C AN A

PF zone (Chaco basin; see


Figure 3). Northeast-trending
LB
rift basins were important in
Brazil. BF, Boomerang fault;
CH, Chiquitanas front; KF,
Khenyani fault system; LB,
Las Breñas fault; LI, Lincoln
structure; PF, Puna fault; SG,
LI
Sierra Grande. (Data from
Visser, 1984; Pezzi and
Mozetic, 1989; França et al.,
1995; Petrobras, proprietary
reports; YPF, proprietary

N
VE N TA N A

reports.)

SG

Pla tfo rm C la st ics


De p oc en te r C la stic s
CA PE
Ca rb on a tes
Faja E ru pti va de la P un a
Gr an ito ids

Vo lca ni cs

Northeast-Oriented Structural Trends shales collected along the length of this structure were
mostly in excess of 4, implying high heat flows and
Northeast-trending structures were an important incipient mineral metamorphism (W. A. M. Jenkins, J. N.
control on basin formation. They are particularly well Theron, J. Utting, 1993, personal communications). TAI
documented in the Paraná basin from outcrop studies values as high as this commonly reflect heat associated
and petroleum exploration (Figure 12). with fault dilation or igneous intrusion (W. A. M.
In South Africa, a prominent northeast-southwest Jenkins, 1993, personal communication). (5) The shape of
(reoriented relative to South America) sidewall is the Permian–Triassic Cape fold belt is attributed to
inferred in the western part of the Cape basin (Figures basement topography (Söhnge, 1983, his figure 1; De
10, 12). (Sidewall refers to the lateral termination of a rift Beer, 1990). Cobbold et al. (1992) have attributed the
segment against a transfer fault; Gibbs, 1984.) The pattern of folding in the western branch of the Cape fold
evidence for this structure includes the following: (1) The belt, stretching lineations, and striations on fault planes
distribution of Pan-African faults suggests a major offset to right-lateral shear on this northeast-trending structure.
(Figure 10). (2) The shape of the Ordovician–Devonian This northeast-oriented crustal structure is believed to
Cape basin, its isopachs, and provenance and paleocur- have accommodated differences in structural style
rent studies reflect this geometry (Rust, 1973, his figures between a broadly subsiding southern Cape trough and
4–8; Theron and Loock, 1988, their figures 8, 10, 13). (3) a tract of extensional basins to the west (Figure 10). Strati-
The distribution of uppermost Cambrian–Ordovician graphic data (Rust, 1973; Theron and Loock, 1988)
braidplain deposits constrains the eastern and western suggest that a similar sidewall must have bounded the
edges of the rift basins (Rust, 1967, p. 84). (4) Thermal western edge of the tract of extensional basins. The
alteration indices (TAI) from 24 samples of Devonian structure of the Ventana basin and Sierra de la Ventana
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 21

The intracontinental extensional basins were formed


as offset and isolated depocenters that were probably
linked tectonically by the small-displacement wrench
faults or accommodation zones striking northeast.
Nelson et al. (1992) describe basement faults of this type
as compartmentalization faults. The regional interrela-
tionships between major structural trends and basin
development are shown in Figure 15. Transects A and B
are approximately orthogonal to the general Paleozoic
plate edge. Figure 15C is a transverse cross section that
emphasizes basin geometry with respect to accommoda-
tion zones and intervening arches.
In summary, the northeast-trending accommodation
zones were an important component of basin evolution.
The various depocenters were able to develop as inde-
pendent rift segments while the accommodation zones
maintained tectonic linkage. Figure 16 attempts to recon-
struct this tectonic framework.

Early Paleozoic Basins


The earlier Pan-African–Brasiliano episode responded
to a fundamental reorganization of stress fields, as
evidenced by the change from widespread extension to
contraction in latest Riphean times. In contrast, the transi-
tion to the Paleozoic Gondwana basins appears to have
been more orderly and gradational (Figures 11, 12, 17).
The Cape basin is believed to have evolved by
exploiting preexisting zones of weakness (De Villiers,
1956; Le Roux, 1983; Söhnge, 1983; De Beer, 1990). In the
western part of the basin, transition from the Klipheuwel
stratigraphy was locally marked by an angular unconfor-
mity (Figure 13), but elsewhere was conformable (Vos
and Tankard, 1981). Whereas the Klipheuwel sediments
are immature and mudstone rich, the basal Cape
Figure 13—Phanerozoic stratigraphic column and interpre- sediments consist of mature conglomerates and sand-
tation. AB, Andean batholith; AV, Andean cordilleran
stones deposited in braidplain environments (Rust, 1967;
volcanics; BG, Botswana gabbro; CA, Chon Aike-Tobifera
suite; CG, Cape granite suite; CH, Choiyoi complex; EB, Vos and Tankard, 1981). The succeeding stratigraphy,
Etendeka basalts; KB, Kuboos-Bremen intrusive suite; including enormous thicknesses (up to 2000 m ) of quartz
Pak-Ced, Pakhuis and Cedarberg formations; PB, Central arenite, was deposited in a broad, slowly subsiding sag
Patagonia batholith; PF, Paraná flood basalts; SB, basin which nevertheless remained extensional (Rust,
Somuncura batholith; SE, Stormberg extrusives; TUC, 1967, 1973; Vos and Tankard, 1981; De Beer, 1990; De
Tucavaca Group. (Data from SACS, 1980; Dingle et al., Wit, 1992).
1983; Hiller and Theron, 1988; Rapela and Kay, 1988; The Cape and Karoo successions are thickest (9 km)
Theron and Loock, 1988; Hiller, 1992; N. Hiller, 1993, (Cole, 1992) adjacent to a significant crustal anomaly
personal communication; Milner et al., 1992; Gresse and recorded in magnetic, gravity, and magnetotelluric data
Scheepers, 1993; Turner et al., 1994; França et al., 1995;
acquired across the Cape fold belt (De Beer, 1983; Pitts et
YPFB, proprietary reports).
al., 1992). The 30-km-wide Beattie magnetic anomaly is
modeled as a southward-dipping body located at least 7
km beneath the surface. Hälbich (1992) has interpreted
of eastern Argentina also reflects a long history of struc- this anomaly as a southward-dipping intracrustal
tural inheritance (Martínez, 1989). Basement rocks show detachment. The southern Cape conductivity belt is
northeast-trending ductile shear zones with retrograde colinear with the Tandilia lineament of Argentina
metamorphism that guided Permian–Triassic deforma- (Figure 2), which Brito Neves and Cordani (1991)
tion (Cobbold et al., 1986, 1992; Rossello and Massabie, interpret as a rift.
1992). The Paraná basin initially evolved by reactivation of
The Lincoln structure (Figure 1) is a small pull-apart old northeast-oriented structures that were associated
basin that was probably colinear with the Blumenau- with immature continental clastics containing
Soledade fault zone of the eastern Paraná basin (Figure interbedded andesitic and rhyolitic volcaniclastic
12). Figure 14 shows the seismic expression of this material (460–510 Ma). According to Almeida et al. (1981)
northeast-oriented transtensional structure. these “Castro cycle” deposits are postorogenic molasse
22 Tankard et al.

Figure 14—Migrated line 300. Uninterpreted (top) and interpreted (bottom) sections showing basin-forming characteristics of
northeast-striking lineaments. The Lincoln structure is a narrow pull-apart basin showing an early rift–postrift couplet of
probable Cambrian–Ordovician age, a conspicuous erosional unconformity separating this lower Paleozoic section from a
thick Carboniferous–Permian succession, and Cretaceous extensional reactivation, including 130-Ma Serra Geral volcanics.
During the Late Jurassic–Cretaceous, the northeast-striking lineaments were reactivated extensionally. Extensional subsi-
dence of the Salado basin took advantage of these trends; rift and sag stages are present. Stratigraphic interpretation of the
Lincoln basin is based on seismic correlation with the Camilo Aldao well 150 km to the northwest. See Figure 1 for location.
(Interpretation by L. Groenewoud; courtesy of Norcen International Ltd. and Pluspetrol S.A.)

that mark the transition from the Brasiliano paleogeog- An enormous episutural trough subsided along the
raphy to the cratonic cover sequence at the initiation of Arequipa-Pampeanas suture of western Bolivia, accumu-
the Paraná basin (Figure 13). The Paraná basin sensu lating over 10,000 m of terrigenous clastics from the
stricto is attributed to postrift thermal subsidence in the Cambrian to Devonian (Dalmayrac et al., 1980; Willner et
Late Ordovician (Zalán et al., 1991) which formed an al., 1987). In northwestern Argentina, igneous and sedi-
enormous gulf open to the west. This episode is equated mentary rock assemblages are attributed to a Neopro-
with the Tacsarian extension of Bolivia characterized by terozoic–early Paleozoic passive margin (Gonzalez
thick, shallow marine quartz arenites. Bonorino and Gonzalez Bonorino, 1991). The Faja
The early Paleozoic margin of southwestern Eruptiva de la Puna of Méndez et al. (1972) (Figure 12) is
Gondwana lay along the western edge of the Pampeanas a magmatic complex of Middle Ordovician age (Lork
terrane, west of which the pre-Cordilleran terrane and Bahlburg, 1993). This magmatic complex and associ-
supported a seaward-facing carbonate platform ated flysch are interpreted as fore-arc (Coira et al., 1982;
(Mpodozis and Ramos, 1989). On the basis of Knüver and Reissinger, 1982; Spalletti et al., 1989) to
reprocessed vibroseis data, Snyder et al. (1990) have back-arc–foreland (Bahlburg, 1990) in origin. There is
attributed early Paleozoic structural styles to extensional evidence that intermittent magmatism (Knüver and
and contractional processes. The Sierra de Valle Fértil Reissinger, 1982) and marginal orogens associated with
fault system is also a major crustal feature that was compressional and oblique-slip processes characterized
repeatedly reactivated throughout the Phanerozoic this Pacific margin until the Early Permian (Gonzalez
(Baldis et al., 1982; Fielding and Jordan, 1988). Bonorino, 1991; Sempere, 1995).

(text continues on p. 29)


Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 23

CHI LE PA GA N Z O SIER R A S CHACO - PARANA BASIN NO RT H PARANA BASI N


TRENCH BA S IN PA MPE A N AS BAS IN
C C-P PC Tr K
T K
SL T P
V
P
P
C
C
NA A T A T
ZC
A
PL
AT
E
A
OU T EN IQ UA KA R O O KAL AHARI GHANZ I -
SWART BERG BAS IN V C P
BA S IN O
BAS IN CHO BE
K PC
SL O P
R
C
O-D

B
B
Secti o n A SCC Sec ti o n B
CH A C O - PA R A N A CA M I LO A LD AO CAP E KA R O O
BA S IN AR C H BAS IN BAS IN
D

T SL O P
P K T K O
C K V
C T C
C S
C O

0 km 200

C VE X 20

Figure 15—Regional tectonostratigraphic cross sections. (A) and (B) Orthogonal to the edge of the continent. (C) Transect
cross cutting northeast-trending structures. See Figure 1 for locations. (Data from Le Roux, 1983; Visser, 1984; Cole, 1992;
França et al., 1995; GSD and YPF, proprietary reports.)
24 Tankard et al.

Figure 16—Schematic block


diagram of early Paleozoic
basin development in which
isolated rift segments are
linked by northeast-striking
accommodation zones.

Figure 17—Devonian paleo-


DEVONI A N geography. The widespread
distribution of the lower
N Paleozoic cover suggests that
basement geology was estab-
lished by the earliest Paleozoic.
CHACO
The Puna-Pampeanas arch has
the form of a giant pop-up
structure (see Figure 22). G,
Grande basin; V, Ventana basin.
(Data from Baldis et al., 1989;
PUNA

França et al., 1995; Petrobras


and YPF, proprietary reports.)
- PAM PEAN A
H RC

Pla tfo rm & B a sin


Ma rg in s
CAPE
De p oc en te r
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 25
26 Tankard et al.
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 27
28 Tankard et al.
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 29

Figure 22—Cross section


showing interpreted evolution
of Puna arch. Transpressional
uplift created a giant pop-up
structure, especially during the
Ocloyic and Chañic orogenies.
Early Carboniferous
Somuncura extension reacti-
vated the previous transpres-
sional structures. See Figure 1
for location. (Modified after Mon
and Salfity, 1995.)

The north-directed Puna structures interacted with Figures 12 and 17 show the distributions of the
the Boomerang-Chiquitanas transcurrent shear to create Cambrian–Ordovician and Devonian basins, respec-
a deep Phanerozoic basin (Figures 2, 3, 12, 17). Seismic tively. In South America, petroleum exploration
sections in Figures 18–21 show the stratigraphic relation- mapping has shown that the lower Paleozoic cover
ships and structural styles of this Chaco basin. Along the sequence is widespread and constrains the age of
Boomerang margin, extensional basement fault blocks accretion of the Patagonia terrane. Furthermore, the
and a basal Silurian detachment form a stepwise Sierra Grande fold belt (Figures 12, 17) consists of
geometry, above which Andean-age folds have Silurian quartz arenites, mudstones, and oolitic iron-
developed as a series of ramp-flat associated structures stones that were deposited in a platform environment
(Figure 18). Pre-Cretaceous erosion has stripped the (Cortés et al., 1984; Ramos and Cortés, 1984).
Carboniferous sequence south of the Enconada well and An important Phanerozoic structure is the south-
has cut deeply into the Devonian sequence in the area of southeast trending Puna-Pampeanas arch system (Figure
the hingeline, creating their wedge-shaped geometries. 17). Mon and Salfity (1995) describe the geology of this
In Figure 19 the entire structural envelop is the result of basement high in northwestern Argentina where it
basement structures and small rifts that are located formed on Brasiliano-age structures (Figure 22). In
beyond the influence of Andean compression. Argentina, this arch separates basins on either side of it,
The Chaco basin (Figure 17) is a westward-thickening much as the Bocono system separates the Maracaibo and
wedge of Phanerozoic sedimentary rocks, mostly of Barinas basins in Venezuela (Giraldo and Osuna, 1994).
terrigenous origin. The succession consists of several The Puna-Pampeanas arch has the characteristics of a
unconformity-bounded sequences, each with a distinct giant pop-up structure (Mon and Salfity, 1995, their
tectonic overprint. Depth to basement below the sub- figure 6) due to oblique-slip processes along the Pacific
Andean fold and thrust belt is at least 12 km. The Izozog margin (Gonzalez Bonorino, 1991; Sempere, 1995) and
arch (Figure 2) is a northwestward-plunging basement intraplate transpressional processes. This basement high
uplift. Isopach reconstructions show that the Devonian may also have had a counterpart south of the Cape basin
depocenter spanned the area of the arch. (Rust, 1967; Ryan, 1967; De Beer, 1983; Visser, 1987a,b).
During Tertiary Andean compression, uplift of the Visser describes middle Paleozoic Gondwana iceflow
Izozog arch resulted in erosion of much of the Paleozoic directions toward the north.
section. Figure 20 is a regional seismic line that shows the The Late Ordovician–earliest Silurian Ocloyic orogeny
most eastern Andean structure (Aimiri) adjacent to the was focused not only on the Puna-Pampeanas arch (Mon
foreland basin and Izozog arch. Sequence boundaries are and Salfity, 1995) but was more widespread as evidenced
essentially parallel in the Paleozoic section, indicating by a regional unconformity (Figure 13). Hälbich (1983a)
that the arch was largely inactive. Figure 21 shows the has described soft sediment disharmonic folding associ-
variety of structural styles in the Andean deformation ated with the Pakhuis diamictites and Cedarberg shales
front, where the principal levels of detachment were the in the Cape basin; he attributes this intraformational
Silurian Kirusillas and the Devonian Los Monos shales. folding to slumping into the tectonically deepened
The Phanerozoic stratigraphy and its dispersal Cedarberg trough. A Ashgillian–early Llandovery age
patterns reflect the evolving early Paleozoic tectonic (Cocks et al., 1970; Berry and Boucot, 1973) suggests a
environment of the Chaco-Tarija basin. During the correlation with the Ocloyic diastrophism. The
Cambrian–Ordovician, the westward-thickening wedge Pakhuis–Cedarberg interval is equivalent to the
was built by sediments derived from the craton. The Cancañiri–Kirusillas succession of Bolivia (see Sempere,
Silurian–Devonian was characterized by sedimentation 1985) (Figure 13 ).
from an orogen in the west that Ramos (1988) and The Devonian paleogeography (Figure 17) was char-
Forsythe et al. (1993) attribute to Silurian docking of the acterized by shallow epeiric seas and cosmopolitan
Arequipa terrane. Malvinokaffric invertebrate faunas (Boucot, 1971;
30 Tankard et al.

Copper, 1977). Intermittent subsidence in the Cape basin Carboniferous–Permian Basins


resulted in stacking of five shoal water depositional
sequences of the Bokkeveld Group (Tankard and Barwis, During the Devonian, underfilled shallow marine
1982; Theron and Loock, 1988). The thickness of the basins were ubiquitous. In contrast, the succeeding
Bokkeveld Group in the Cape trough varies up to 4000 m Carboniferous–Permian basins were generally overfilled
(Theron and Loock, 1988). Conversely, Ponta Grossa and characterized by mixed terrigenous continental envi-
deposits in the Paraná basin are generally less than 500 m ronments, periodic marine incursions, and development
thick. According to Zalán et al. (1991), northwest- of the Gondwana glaciation in several phases. (The
oriented normal faults compartmentalized the Paraná concept of underfilled and steady-state stages is
basin. This interpretation finds support in the pre- discussed by Covey, 1986.)
Cordillera, where northeast-trending shear faults had a The Karoo basin of South Africa is famous for its rich
right-lateral sense of displacement (Baldis et al., 1989). In mammal-like reptile biota (Tankard et al., 1982). The
the Chaco basin, the Devonian section was originally Karoo is generally separated from the Cape succession
more than 4000 m thick before it was eroded deeply by by a significant unconformity (Figure 13). Late Carbonif-
Tertiary uplift of the Izozog arch. erous Dwyka iceflow directions indicate topographic
Inversion of these early Paleozoic basins in the Late relief south of the flysch trough. There are also thick
Devonian and earliest Carboniferous is attributed to the accumulations (up to 3000 m) of Ecca sediments in this
Chañic orogeny. This event is marked by a widespread trough. Submarine fan deposits and deltaic sediments
unconformity. Renewed glaciation was associated with also support a southern provenance (Kingsley, 1977;
this Late Devonian orogenesis in the Titicaca region of Ryan and Whitfield, 1979; Wickens, 1992). This southern
Bolivia (Caputo, 1985). highland may have been a continuation of the Puna-

Figure 23—Carboniferous–
Permian paleogeography.
Right-lateral transpression
emphasized the Puna-
Pampeanas arch, along which
transtensional depocenters
formed. These stress fields
resulted in normal sense reac-
tivation of the northeast-
trending structures. Late
Carboniferous–Permian
postextensional subsidence
resulted in broad, shallow
epicontinental seaways in
which bituminous shales accu-
mulated. Note offsets of
highlands, especially between
Karoo and Kalahari basins. BF,
Boomerang fault; BL, Bahia La
Lancha suture; CH, Chiqui-
tanas front (San José fault);
GF, Guapiara fault; I, Irati
Formation; KF, Khenyani fault;
KL, Kalahari line; LB, Las
Breñas fault; MR, Mokoro rift;
P, Patquía Formation; WH,
White Hill Formation; ZF, Zoet-
fontein fault. (Data from Reyes,
1972; McLachlan and
Anderson, 1973; Anderson and
McLachlan, 1979; Van Vuuren
and Cole, 1979; Riccardi and
Rolleri, 1980; Visser, 1987a;
Pezzi and Mozetic, 1989;
Fernandez-Seveso and
Tankard, 1995; França et al.,
1995; Petrobras and YPF,
proprietary reports.)
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 31

Pampeanas arch of Argentina (Figure 23). Although bituminous shale-limestone drape of the Irati Formation,
tenuous, isopach and structural contour maps which has total organic carbon (TOC) values as high as
(McLachlan and Anderson, 1973; Van Vuuren and Cole, 20%. As fault activity decreased and the sag basin
1979) suggest that during Dwyka–Ecca time, the Karoo gradually widened, an epicontinental seaway linked the
basin may have been compartmented by northeast- Karoo basin in the southeast (Eyles, 1993, his figure 16.9).
oriented structures (Wickens et al., 1992) (Figure 23). The Irati Formation of Brazil and the White Hill
The Carboniferous–Permian Claromeco basin on the Formation of South Africa are contemporaneous. They
conjugate margin of Argentina has a similar southwest- are important because they contain the aquatic reptile
facing asymmetry (Figure 23). Paleocurrents in the Tunas Mesosaurus, once crucial supporting evidence for conti-
Formation are from the Puna-Pampeanas highland along nental drift (Du Toit, 1927).
the southwestern margin of the basin (Lopez-Gamundi The Carboniferous–Permian basins of northwestern
et al., in press a). These authors suggest a genetic linkage Argentina and Bolivia (Figure 23) resulted from
with the Karoo basin. The relationship of the Claromeco compressional, transtensional, and oblique-slip processes
basin to the fringing highland probably has significance related to the adjacent Pacific margin (Lopez-Gamundi et
for the southern Karoo basin. al., 1989; Rapalini, 1990; Gonzalez Bonorino, 1991; Mon
On its northern side, the Karoo basin thins deposition- and Salfity, 1995; Sempere, 1995; Fernandez-Seveso and
ally across the Kalahari craton and is separated from the Tankard, 1995). Northwest-trending en echelon folds in
Kalahari basin of Namibia and Botswana by a highland northern Chile have been attributed to dextral transcur-
(Figure 23) (Visser, 1987a). In the latter, a major hiatus rent movement (Breitkreuz, 1991). Like the Paraná basin,
separates the Cambrian Nama from the uppermost these basins have a history of Early Carboniferous exten-
Carboniferous stratigraphy. Extensional reactivation of sional subsidence that gradually diminished through the
Damaran structures created a suite of rift basins along Permian (Caminos, 1979). The Devonian succession is
the Damara-Ghanzi and Limpopo belts, local transpres- locally folded and metamorphosed and is separated
sional folding notwithstanding (Münch, 1974). Isopachs from the Carboniferous by an unconformity attributed to
of the Kalahari basin (Visser, 1987a) conform closely to the Chañic diastrophism of the Late Devonian–earliest
the outline of the Nosop basin complex and its bounding Carboniferous (Archangelsky et al., 1986; Mon and
Kalahari fault (Figures 1, 23). The Masetlheng Pan explo- Salfity, 1995).
ration well encountered Upper Carboniferous (Kasi- The Paganzo basin formed across the Pampeanas,
movian–Gzelian) glacial and varved restricted marine pre-Cordillera, and Chilenia terranes of northwestern
shales above Cambrian red beds (GSD, proprietary Argentina (Figure 23). It consists of the Guandacol, Tupe,
reports). Similar pollen and spore assemblages occur in and Patquía unconformity-bounded sequences
the coal measures of the middle Zambezi valley (Utting, (Fernandez-Seveso and Tankard, 1995). The Guandacol
1978). The orientation of the Mokoro rift of eastern accumulated in isolated pull-apart basins created by
Botswana to the Zoetfontein fault (Green et al., 1980) right-lateral wrench tectonics. It is characterized by rapid
suggests a right-lateral translation of the latter (Figure stacking of alluvial sediments adjacent to basin-
23). The actual onset of extension may be slightly older. bounding faults, massive synsedimentary deformation,
The Masetlheng Pan well intersected a thick gabbro sill at and glaciolacustrine sedimentation of Visean age. The
3600 m within the Witvlei section (Gariep equivalent). It overlying Tupe and Patquía sequences (Westphalian–
yielded a K-Ar age of 333 ± 17 Ma (middle Carbonif- Permian) reflect gradual cessation of fault-controlled
erous) (GSD, proprietary report). subsidence as the various depocenters were yoked
Tectonic uplift associated with Gondwana glaciation together in a regional downwarp (Fernandez-Seveso and
resulted in a 45-m.y. hiatus in the Paraná basin where Tankard, 1995). The Upper Permian Patquía Formation
Westphalian–Artinskian Itararé sedimentary rocks contains restricted marine and lacustrine bituminous
overlie the Devonian Ponta Grossa (Figure 13) (Gravenor shales of Irati–White Hill age.
and Rocha-Campos, 1983; Eyles and Eyles, 1993; Eyles et This Paganzo basin is linked to the Chaco-Tarija basin
al., 1995). Great thicknesses of Devonian strata were of northern Argentina and Bolivia via a system of
eroded by Gondwana glaciation. The Itararé Group diverging strike-slip faults. Mon and Salfity (1995)
comprises stacked glacigenic depositional sequences that describe transpressional uplift of the Puna arch during
record stepwise deepening of the basin across major the Chañic event and Early Carboniferous Somuncura
northeast-oriented faults. This structural relief separated extensional subsidence (Figure 22).
the Paraná and Kalahari basins (Figure 23). Glaciers The Carboniferous–Permian succession in Bolivia is
spreading westward into the Paraná basin were only half as thick as the Paganzo succession (1000 and
grounded on this highland (J. N. J. Visser, 1993, personal 2000 m, respectively) and may reflect a northward-
communciation). diverging transcurrent fault system (Fernandez-Seveso
The fault-controlled subsidence characteristic of the and Tankard, 1995). The Lower Carboniferous Machareti
lower Itararé Group progressively diminished and was Group is a counterpart of the Guandacol and stratigraph-
gradually replaced by regional subsidence that created a ically similar. The Mandiyuti Group is equivalent to the
broad epeiric basin. The style of Permian sedimentation Tupe. Deeply incised paleovalleys of the Tarija and
changed progressively through the sheetlike Chapéu do Escarpment formations have been mapped from the
Sol rain-out deposit and the coal-bearing Rio Bonito dense grid of reflection seismic data (Figures 19, 20, 24).
deltaic sequence. It culminated in the Upper Permian These paleovalleys are typically 500 m deep and several
32 Tankard et al.

Figure 24—Carboniferous paleo-


valleys of the Escarpment
Formation, Chaco basin, Bolivia,
based on a dense grid of seismic
data. These paleovalleys
possibly fed into transtensional
basins. The Boomerang-
Tucavaca erosional edge is
Mesozoic. SC, Santa Cruz de la
Sierra. (Data from Vargas and
Suárez, 1980; Suárez, 1986;
YPFB, proprietary reports.)

kilometers wide. Internally they consist of stacked dextral transcurrent movement in the pre-Cordillera
channel sandstones and mudstones. The network of (Baldis et al., 1989) and the northwest-southeast normal
paleovalleys is believed to have drained northward and fault subsidence in the Paraná basin (Zalán et al., 1991)
northwestward into Tupambi (Machareti Group) suggests that the direction of principal extensional stress
transtensional basins. Their orientation matches the (σ3) in the Devonian was toward the southwest.
extensional trend, suggesting a tectonic control of A similar history characterized Carboniferous–
glaciofluvial processes. Permian paleogeography. Extension was widespread in
the Carboniferous, as exhibited by the Paganzo, Chaco,
Interpretation Paraná, Claromeco, Karoo, and Kalahari basins.
Transtensional subsidence in the Paganzo and Chaco
The form of the Paleozoic basins was largely basins was consistent with reactivation in a normal sense
controlled by Brasiliano structural inheritance. In the of the northeast-oriented structures in the Paraná basin,
early stages of subsidence (Cambrian–Ordovician), the implying an extensional stress field (σ3) directed toward
northeast trends were repeated (e.g., Figure 14). the northwest. The Claromeco basin is believed to have
Gradually the fine detail of fault-controlled subsidence originated as a post-Chañic intracontinental rift (López-
was abandoned, and subsidence and sedimentation were Gamundí and Rossello, 1992) (Figure 23). A similar
more uniform through the Ordovician, culminating in response apparently occurred in southern Africa where
the Late Ordovician Ocloyic orogeny and regional the Kalahari fault (Visser, 1987a; Hall et al., 1990) and
unconformity. This pattern of postorogenic extensional trans-Karoo trends (McLachlan and Anderson, 1973; Van
subsidence and succeeding regional sag was repeated in Vuuren and Cole, 1979) also affected isopach thicknesses.
the Silurian–Devonian and was in turn terminated by the Middle Carboniferous extension also explains granitoid
Chañic diastrophism (Figure 13). Only the first-order emplacement in the western Pampean ranges (311 Ma)
tectonic trends persisted, such as the Puna-Pampeanas (Cingolani et al., 1993) and a thick gabbro sill in
arch and the Boomerang-Chiquitanas transcurrent shear. Botswana (333 Ma) (GSD, proprietary reports). Through
The gradual change to more uniform styles of subsidence the Late Carboniferous–Permian, the extensional basins
may reflect relaxation of extensional stresses (Bally and widened and coalesced as fault-controlled subsidence
Snelson, 1980). The northeast-southwest orientation of decreased (Polansky, 1970; Fernandez-Seveso and
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 33

and Cape-Karoo curves show similar patterns of subsi-


dence; both have two episodes of tectonic subsidence. In
marked contrast, the Paganzo basin is characterized by
rapid subsidence typical of a transtensional basin
(Christie-Blick and Biddle, 1985).
Compressional events appear to have been most
pronounced along the Puna-Pampeanas arch. The Late
Ordovician–earliest Silurian Ocloyic foldbelt (Mon and
Hongn, 1991) is associated with a widespread unconfor-
mity (Figure 13). Disharmonic Cedarberg folding in the
Cape basin (Hälbich, 1983a) may be contemporaneous.
The latest Devonian–earliest Carboniferous Chañic
orogeny (Bahlburg et al., 1987; Bell, 1987) was also
important. This geology may also explain the Claromeco
and southern Karoo troughs as foreland basins that
resulted from transpressional deformation along a
southern highland (Figure 23). This highland, the North
Patagonia-Chadileufu-Pampeanas basement complex of
López-Gamundí et al. (in press b), formed a drainage
divide between basins on either side. Caminos et al.
(1988) infer progressive uplift through the late Paleozoic.
The basement arch complex appears to divide, and
resulted in separation of, the Paraná, Kalahari, and Karoo
basins (Figure 23).
Uplift of the Puna-Pampeanas arch resembled a giant
pop-up structure, on either side of which basins formed,
including the Tepuel, San Rafael, Chaco-Paraná, and
Claromeco basins (Figure 23). In this respect, the arch
and its flanking basins may be similar to the modern
Bocono fault system of Venezuela, the outward
vergences of which face the Maracaibo and Barinas
basins on either side (Giraldo and Osuna, 1994).
Tectonic arching had important climatic effects which,
in turn, strongly influenced the stratigraphy and sedi-
mentology of flanking basins. Localized uplift has been
implicated as a causative mechanism for Gondwana
glaciation, whereby small ice covers may have triggered
regional cooling. It should be noted that the Late Ordovi-
cian–earliest Silurian Ocloyic orogeny and the Late
Devonian–Early Carboniferous Chañic orogeny
coincided with Gondwana glaciations. Amos (1972) and
Eyles (1993) attribute the expansion of these ice sheets to
topographic relief created by tectonism. This also
explains the earlier development of glacial conditions in
the Paganzo basin (Visean) than in the Paraná basin
(Westphalian–Sakmarian), because the pre-Cordillera
Figure 25—Backstripped tectonic and basement subsi- was the focus of the Chañic deformation and strike-slip
dence curves for Paganzo (Fernandez-Seveso and
Tankard, 1995), Paraná (Oliveira, 1987), and western Cape-
tectonics. The synchroneity of Paleozoic sea level
Karoo basin (Cloetingh et al., 1992). The steep Paganzo changes and history of glaciation appears to reflect
curves are typical of strike-slip processes (Christie-Blick tectonic behavior.
and Biddle, 1985). Finally, this interpretation has significance for the
petroleum geology of the area. The Devonian and
Permian paleogeographies are attributed to postexten-
Tankard, 1995). This regional subsidence reached its sion regional subsidence and development of broad,
acme during the Late Permian with accumulation of epicontinental seaways. Both periods were characterized
widespread drapes of bituminous calcareous shales: by organic-rich shales deposited in anoxic bottom water
White Hill, Irati, Patquía, and Vitiacua formations settings. Of the Devonian Bokkeveld, Ponta Grossa, and
(Figure 23). Los Monos formations, only the Los Monos of Bolivia is a
The subsidence curves shown in Figure 25 summarize prolific hydrocarbon source rock. The Upper Permian
the Carboniferous–Permian subsidence history for the bituminous shales in the White Hill, Irati, Patquía, and
Paganzo, Paraná, and Cape-Karoo basins. The Paraná Vitiacua formations also reflect restricted marine circula-
34 Tankard et al.

Figure 26—Triassic–Early
Jurassic extensional paleo-
TRI AS S IC - EAR LY J URAS S I C geography. Permian–Early
Triassic contraction is included
N to show the relationship of
Triassic extension to the
preceding orogenesis. BL,
Bariloche; CFB, Cape fold belt;
CU, Cuyo basin; GF, Gastre
fault system; IG, Ischigualasto
basin; MB, Molteno depocenter;
NQ, Neuquén basin; SG, Sierra
Grande; SP, Sierra Pintada; SS,
Sierra Septentrionales; SV,
Sierra de la Ventana. (Data from
IG
Coira et al., 1975; Bianchi, 1984;
Tr
Uliana et al., 1989; Zalán et al.,
1991; Cobbold et al., 1986,
1992; França et al., 1995; Wiens,
CU 1995; YPF, proprietary reports.)
SP

Tr

NQ
SV

SS

Tr

N
BL

SG

Tr

CF
B

Rif t Ba si n s ,
Ma i nl y Cl a st ic Fi ll

MB Th in Cr at o ni c Co v er
Str uc t ura l I nv e rs i on
La te Pe rm ia n - Ea rl y Tri as s i c
Ex te ns i on a l Fa ul ts
Tr Tr an s fe r Fau l ts

tion and high TOC values. None of these Permian source and became the locus of intense magmatic activity
rocks are known to have sourced significant accumula- (Caminos et al., 1988). Postorogenic extension followed
tions, although the Permian Ene Formation is one of the two paths. North-northwest trending rift basins formed
most important source rocks in Peru. In the Chaco basin along the fringes of southwestern Gondwana during the
of Bolivia, the Carboniferous paleovalley sandstones Triassic and Early Jurassic. Late Jurassic–Cretaceous
(Tarija and Escarpment formations) are a major explo- extension was linked to northwest-directed shear
ration play. Glacigenic lithologies of the Itararé couples that preceded Atlantic opening; overall these
Formation in the Paraná basin are a major gas play, basins developed inboard of the earlier Mesozoic
whereas the most encouraging oil shows occur in the extension (Uliana and Biddle, 1988; Uliana et al., 1989).
Permian Rio Bonito sandstone. The final important stage of basin evolution resulted
from Cenozoic Andean contraction and local inversion of
the Mesozoic basins.
OROGENY AND MESOZOIC BASINS
Permian–Triassic Orogenesis
An entirely new style of basin development began in
the late Paleozoic when many of the earlier Paleozoic The late Paleozoic orogens were intracratonic and
basins were inverted in a series of discontinuous intra- formed a set of isolated segments (Cobbold et al., 1986,
continental foldbelts (Figure 26) (Cobbold et al., 1986). 1992) that were associated with northeast-trending faults
The 2500-km-long San Rafael orogen flanked the Pacific and lineaments (Figure 26) in a way that suggests the
margin of Chile and Argentina (Azcuy and Caminos, latter may have functioned as accommodation zones.
1986; Llambias and Sato, 1990; Mpodozis and Kay, 1990) The distribution and probable tectonic linkage of the
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 35

preinversion extensional basins is reconstructed in been deformed with northeastward vergence (Cortés et
Figure 16. (By structural inversion, we mean the reactiva- al., 1984; Ramos and Cortés, 1984). The farfield effects of
tion in a reverse sense of basin-forming extensional this deformation are also recorded in the transpressional
faults; see Williams et al., 1989.) reactivation of structures in the Paraná basin (e.g.,
The Cape fold belt is 900 km long and up to 380 km Curitiba fault system) (Figure 27). Daly et al. (1991) and
wide. It consists of a southern branch and an orthogonal De Wit and Ransome (1992) discuss farfield deformation
western branch, both of which developed contemporane- in Gondwana.
ously (De Beer, 1990). On the basis of whole rock Ar-Ar The Late Permian San Rafael compression formed a
step-heating analyses and stratigraphic considerations, 500-km-wide orogenic belt adjacent to the plate margin
Hälbich (1983b) suggests an Early Permian–Middle of Chile and Argentina (Llambias and Sato, 1990;
Triassic age range for this orogeny, although these dates Mpodozis and Kay, 1990). This mountain belt was the
are poorly constrained. According to Hälbich, the onset locus of intense synorogenic and postorogenic magmatic
of deformation postdates Dwyka sedimentation. activity. The Choiyoi magmatic event comprises a volu-
A voluminous literature relates the style of Cape minous volcanic and plutonic suite dominated by a
deformation to inherited basement structures of Pan- rhyolite-ignimbrite association (Kay et al., 1989; Rapela
African age (e.g., De Villiers, 1956; Hälbich, 1983b; Le and Pankhurst, 1992).
Roux, 1983; Söhnge, 1983; De Beer, 1990; De Wit, 1992).
The southern branch is folded about axes subparallel to Triassic–Jurassic Extension and Magmatism
old basement structures. Folds are most intense close to
high-angle basement faults where upright to nearly Triassic–Early Jurassic lithospheric extension was
isoclinal recumbent folds occur. Folds verge northward located mainly along the Permian–Triassic orogenic belt
toward the craton. where it formed a broad band of half-grabens with an
The southern branch of the Cape fold belt meets the overall north-northwest trend (Figure 26) (Uliana and
western branch in a zone of syntaxis, which Söhnge (1983) Biddle, 1988; Uliana et al., 1989). These trends developed
attributed to interaction with a block-faulted basement; along pre-Triassic basement structures, including the
Ransome and De Wit (1992) have attributed these charac- pre-Cordillera–Pampeanas and Chilenia–pre-Cordillera
teristics to microplate tectonics. Elements of the western terrane boundaries. Basins such as the Cuyo, Bolsones,
branch include north-trending open folds near the and Ischigualasto of Argentina and the Molteno of South
syntaxis, a prominent north-trending monocline along the Africa formed inboard of the compressional belt, where
eastern margin of the belt, and folds that diverge toward they were landlocked and hosted continental deposi-
the northwest (Söhnge, 1983, his figure 1) where they tional systems (Turner, 1983; Uliana et al., 1989). Lacus-
coincide with older Klipheuwel rifts (De Beer, 1990). trine bituminous shales in the Cuyo basin (Cacheuta
The overall geometry of the Cape fold belt, its paral- Formation) and Molteno basin are of similar age.
lelism to early Paleozoic structural fabrics, and the partic- The extensional basins are offset and linked by
ipation of high-angle mylonitized basement faults in northeast-trending accommodation zones (Figure 26).
deformation all suggest that the fold belt has evolved by These accommodation zones even provide linkage across
inversion of previous rift basins. Furthermore, the uplifted areas, such as the rift jump between the
change of direction from a southern to western branch Neuquén and Cuyo basins (Welsink et al., 1995). Figure
mirrors the inferred sidewall configuration of the earlier 27 shows transpressional reactivation in the Paraná
rift segments (Figure 16). Cobbold et al. (1992) interpret a basin, where a Triassic unconformity has been eroded
north-striking right-lateral shear (sidewall) to explain the across a positive flower structure. Extension was also
pattern of folding, stretching lineations, and fault plane prevalent across the San Rafael orogenic belt and
striations. magmatic arc of the Cordillera Frontal and pre-
The Sierra de la Ventana (Australes) foldbelt of Cordillera (Rolleri and Criado Roqué, 1968; Rolleri and
Argentina trends northwestward for 180 km along the Garrasino, 1979). Throughout the region, the principal
southern margin of the Rio de la Plata craton (Figure 26). Triassic faults were guided by preexisting structures. For
Its Lower Devonian–Permian stratigraphy is similar to example, the Ischigualasto basin formed by reactivation
that of South Africa (Keidel, 1916; Du Toit, 1927; López- of Paganzo and older structures.
Gamundi et al., in press a). The Paleozoic cover was Legarreta et al. (1992) and Kokogian et al. (1993) have
deformed initially in the Permian by right-lateral described the structural framework of the Cuyo half-
wrenching (Cobbold et al., 1986). The structural style is grabens and accommodation zones. To the south, the
dominated by northeast-verging folds linked to Neuquén basin was subjected to Triassic–Early Jurassic
basement shear zones. Cobbold et al. (1986) interpret extension followed by a long period of Middle
these relationships as structural inversion of preexisting Jurassic–Paleogene subsidence (Uliana and Legarreta,
normal faults. Cratonward thrusting was influenced by 1993; Vergani et al., 1995). The early Neuquén basin was
northeast-southwest dextral wrenching (Cobbold et al., structurally compartmented, suggesting that Lower
1986, 1992; Rossello and Massabie, 1992). Jurassic source rocks probably accumulated in several
Similar late Paleozoic deformed belts occur at Sierras isolated depocenters.
Septentrionales, Sierra Grande, and Bariloche (Figure 26) This extensional period was accompanied by prodi-
(Cobbold et al., 1992). The Sierra Grande fold belt gious magmatism (e.g., Kay et al., 1989). The Gastre fault
contains Silurian–Devonian terrigenous clastics that have system of central Patagonia is a fundamental geologic
36 Tankard et al.

Figure 27—Seismic expres-


sion of Late Permian–Early
Triassic transpression along a
northeast-oriented fault in the
central Paraná basin, showing
the Lower Triassic unconfor-
mity. P, Permian, Tr, Triassic.

boundary and was subjected to right-lateral displace- the northeast-directed accommodation zones implies
ment in the Mesozoic (Coira et al., 1975; Marshall, 1994). that the direction of principal extensional stress (σ3) was
Development of mylonitic fabrics was also associated toward the southwest in Triassic–Early Jurassic time. We
with Triassic–Early Jurassic plutons of the central believe that the processes controlling extension of these
Patagonia batholith (Rapela and Pankhurst, 1992, 1993). basins was analogous to the Basin and Range province of
Transcurrent movement on this fault system is also western North America. The latter is a Cenozoic exten-
believed to explain the granites of the northern Pata- sional province that is probably driven by extensional
gonian massif which have now been dated as collapse of the Laramide orogen (Wernicke, 1985;
Permian–Triassic. (These northern Patagonian granites Dewey, 1988; Verrall, 1989; Jones et al., 1992). In
had previously been dated as Carboniferous and were an Argentina, an important component of the extensional
essential part of the justification for a middle Paleozoic process was subduction rollback (Uliana et al., 1989).
age of accretion of the Patagonia terrane. Also, the
geochemistry of the Somuncura batholith is not typical of Late Jurassic–Cretaceous Extension
subduction-related plutonism, according to Rapela and
Kay [1988].) In the Late Jurassic–Cretaceous, the Permian–Triassic
The Choiyoi is a Late Permian–Early Triassic suite of compressional belt was no longer the locus of extension,
silicic volcanic and volcaniclastic rocks. The Choiyoi has except in South Africa. Instead, extension had propa-
been attributed to orogenic crustal thickening and lower gated inboard and was more widely distributed
rates of subduction (Llambias and Sato, 1990; Mpodozis (Bianucci and Homovc, 1982; Uliana and Biddle, 1988;
and Kay, 1990). The Middle Jurassic (Bajocian, 180 Ma) Uliana et al., 1989). These Late Jurassic–Cretaceous rift
Marifil rhyolite-ignimbrite suite of Patagonia is a volumi- basins were associated with a northwest-trending dextral
nous bimodal volcanic province associated with the shear system (Figure 28). Transcurrent shear faults such
north-northwest trending system of extensional basins. It as the Agulhas-Malvinas transform, the Gastre fault
was followed by Middle–Late Jurassic magmatism of the system, and the Martin Garcia fault were conspicuous.
Tobifera and Chon-Aike formations (about 165 Ma), Clockwise rotation of the Antarctic Peninsula initiated
which has subduction affinities in the west (Bruhn et al., spreading in the Weddell Sea between 175 and 155 Ma,
1978; Rapela and Pankhurst, 1992). thus isolating east and west Gondwana (Grunow, 1993).
The close association of this Triassic–Early Jurassic rift The onset of opening in the South Atlantic was at about
system with the Permian–Triassic belt and inversion 130 Ma (Rabinowitz and LaBrecque, 1979). Widespread
segments (Figure 26) and with postorogenic bimodal extension was also associated with Late Jurassic–Creta-
volcanism and plutonism suggests an origin by exten- ceous subduction along the Pacific margin (Mpodozis,
sional orogenic collapse (Dewey, 1988). The overall 1984). Upper Jurassic–Neocomian deposition was
distribution of the basin-forming extensional faults and generally nonmarine, implying high continental
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 37

Figure 28—Late
Jurassic–Cretaceous paleo-
L ATE J URAS SI C - C RETA CEOUS geography. A, Alemania basin;
AG, Agulhas transform; AL,
N

.
F
AN
O Altiplano basin; AR, Asunción
I LI
AL S B RA
S
rift; BO, Bolsones basin; CO,
AN
TR Colorado basin; CU, Cuyo
IZ O Z O G
basin; DH, Dorsal de Huincul;
KF
AR CH
EL, Entre Lomas structure; ET,

G
MI CH I CO L A El Toro fault; GF, Gastre fault

U
A
AR CH

PI
system; LC, Bahia La

A
TC

R
A
.
ET SJ Lancha–Rio Chico lineament;

FA
LO

U
CU
F.

LT
GA
U TI N LO, Lomas de Olmedo basin;

R
IT
JA C

IB
A M N

A
EA M, Metán basin; MA, Malvinas

F.
MP
TR A N S PA M

PA A R C H AR
basin; MG, Magallanes basin;
CIN
HA NI, Nirihuau basin; NQ,
AR CH

L AN
BO Neuquén basin; OB, Orange
basin; OU, Outeniqua basin;
P E AN

RA, Rawson basin; SA, Salado


AR
TI

basin; SJ, Salta-Jujuy high;


N
G

CU SJB, San Jorge basin; SL,


AR
C

Santa Lucía basin; TC, Tres


IA

SL
Cruces basin. (Data from Coira
NQ EL
SA
et al., 1975; Bianchi, 1984;
Salfity, 1985; Chebli et al., 1989;

N
DH
CO
Santa Ana, 1989; Uliana et al.,
1989; Fouché et al., 1992;
Milani, 1992; Eyles and Eyles,
OB
NI GF 1993; Muntingh and Brown,
RA
1994; Welsink et al., 1995.)
S JB

Rif t B as in Fil l
L
AG

O Pla tfo rm C o ve r
U

MG Str uc tu ral Inv e rsio n


MA

freeboard. Exceptions were the Malvinas plateau and et al., 1992). Their synrift deposits vary up to 10 km in
Magallanes basin. thickness. A regional late Valanginian unconformity
The initiation of the extensional phase is poorly separates the synrift and postrift deposits. The formation
constrained. Eight sedimentary basins underlie the conti- of these basins has been attributed to relaxation or
nental shelf of Argentina. These are generally attributed collapse of the Cape orogenic belt (De Wit and Ransome,
to Late Jurassic–Neocomian extension (Urien et al., 1976; 1992; Fouché et al., 1992).
Stoakes et al., 1991; Keeley and Light, 1993). The base of Along the shelf and eastern seaboard of Argentina
the synrift succession is exposed in the Santa Lucía basin and Uruguay, the rifts are oblique or orthogonal to the
of Uruguay where Puerto Gomez tholeiitic basalts and continent–ocean boundary. In the Punta del Este, Salado,
intercalated lacustrine deposits overlie Neoproterozoic and Colorado basins, the northeast-trending accommo-
basement (ANCAP, proprietary report). The Gamtoos dation zones of earlier Mesozoic subsidence were reacti-
and Algoa basins offshore South Africa are believed to vated in a normal sense (e.g., Figure 14). Farther south,
contain Kimmeridgian–Tithonian synrift sediments the northeastern fault-bounded margin of the San Jorge
(Fouché et al., 1992). Structural inverson in the Neuquén basin is colinear with the Agulhas-Malvinas-Gastre fault
basin culminated in the latest Oxfordian–earliest system (Figure 28). This fault system coincides with a
Kimmeridgian Araucanian event (Vergani et al., 1995). small circle based on the pole of early South Atlantic
A suite of Late Jurassic–Cretaceous half-grabens opening (Rapela and Pankhurst, 1992). A prominent
underlies the continental shelf of South Africa. They are fault also defines the southern margin of the San Jorge
bounded by arcuate listric normal basin-forming faults basin (Figure 28).
that are genetically linked to the Agulhas-Malvinas A linear belt of transtensional subsidence links the
fracture zone (Figure 28) (Bate and Malan, 1992; Fouché Colorado basin with the rift system of northwestern
38 Tankard et al.

Figure 29—Uninterpreted (top) and interpreted (bottom) seismic line showing structural inversion of the Huincul arch in the
Neuquén basin. See Figure 1 for location.

Argentina, where en echelon rift segments have over- suggests that Paleozoic stratigraphy is preserved in
lapped and surrounded the Salta-Jujuy high (Figure 26). deeply down-faulted blocks (Cominguez and Ramos,
Synrift deposits in these Salta rift basins (Lomas de 1995). It would appear that a process of rifting and rift
Olmedo, Metán, Alemania, and Tres Cruces) include flank uplift would better explain the Michicola and
thick sequences of terrigenous sediments, evaporites, Pampean arches.
and intercalated volcanics. The Paleozoic section across The southern Altiplano is part of the Cretaceous
the Michicola and Pampean arches flanking the Lomas Andean rift. It was controlled by northwest-southeast
de Olmedo basin has been eroded, prompting interpreta- and northeast-southwest structures that reactivated a
tion of prerift thermal doming (e.g., Bianucci and Brasiliano-age structural grain. Variable extension was
Homovc, 1982). However, reprocessed vibroseis data accommodated by northwest-striking transfer faults
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 39

Continental flood basalts of the Paraná basin and the


smaller Etendeka basalt province on the conjugate
Namibian margin (Figure 30) have been dated at 130–135
Ma (Milner et al., 1992; Turner et al., 1994). The locus of
the Paraná basalts is the Alto Paraguay lithospheric
terrane and its associated structural anistropies (Figure
30). The isopach distribution of this lava pile mimics the
structural framework (Zalán et al., 1991; Milani, 1992).
Maximum thickness is 1800 m. The lava flows were fed
by dikes rooted in the extensional fabric; northwest,
northeast, and north-northeast trends predominate. Late
Jurassic–Neocomian magmatism in Paraguay, Uruguay,
northern Argentina, and Bolivia was generally related to
extensional faulting (Avila-Salinas, 1986; Uliana et al.,
1989). The Velasco alkaline magmatic province of eastern
Bolivia consists of several plutonic bodies that were
intruded along northeast-striking lineaments during
extensional episodes (Fletcher and Litherland, 1981).
It is generally assumed that the Paraná basalts formed
over a fixed Tristan hotspot. However, seismic tomo-
graphic studies do not support a plume head interpreta-
tion (Anderson et al., 1992). The Paraná flood basalts
appear to be related to the Alto Paraguay terrane, a long
history of Phanerozoic extension that preweakened the
lithosphere, and a reorganization of regional stress fields
associated with changing plate boundary forces
(compare Figures 26, 28) (Storey et al., 1992). The Paraná
and Etendeka basalts were extruded at the start of the
opening of the South Atlantic. Also by Late Jurassic time,
Figure 30—Distribution and isopachs of Paraná-Etendeka a continuous magmatic arc had evolved along the
magmatism in relation to the Alto Paraguay terrane and margin of the Pacific plate (Jensen et al., 1976; Cisternas,
prominent extensional structures. AP, Alto Paraguay 1979; Mpodozis, 1984).
terrane; BS, Blumenau-Soledade fault; CF, Curitiba fault; Late Jurassic–Cretaceous extension was controlled by
GF, Guapiara fault; LF, Lancinha-Cubatão fault; JF, a northwest-trending right-lateral shear system related to
Jacutinga fault; RP, Rio de la Plata terrane; VA, Velasco the opening of the Weddell Sea (Grunow, 1993). The
alkaline province. (Data from Fletcher and Litherland, 1981; principal extensional stress (σ3) was oriented toward the
Hawkesworth et al., 1992; Milani, 1992.) northwest. These characteristics resulted in a greater
dispersion of rifting and abandonment of the
during the Berriasian–Campanian interval (Welsink et Permian–Triassic compressive tract. This regional stress
al., 1995). field resulted in reactivation of the northeast-oriented
Structural inversion was common during the Late Triassic–Early Jurassic transfer faults in a normal sense
Jurassic and Cretaceous, depending on the relative orien- (Figure 14) and structural inversion of earlier rifts (Figure
tations of basin-bounding faults and prevailing stress 29). Comparison of the stratigraphies of the Lomas de
fields (e.g., Dellapé and Hegedus, 1995; Vergani et al., Olmedo and Cuyo basins (Figure 31) is a useful test of
1995). A spectacular example of structural inversion is the admissibility of this Mesozoic extensional dichotomy.
the Dorsal de Huincul in the Neuquén basin, which Their histories of subsidence are mutually exclusive. In
inverted the Triassic–Early Jurassic depositional trough the same vein, inversion of the Dorsal de Huincul in the
(Figure 29). Ploszkiewicz et al. (1984) attributed this Neuquén basin (Figure 29) dates the reorganization of
structure to wrench-induced inversion. The Dorsal de Mesozoic stress fields to about the latest Oxfordian–
Huincul, Loma La Lata, and Entre Lomas are a suite of earliest Cretaceous, which was about when Atlantic
structures that diverge toward the northwest (Vergani et extensional subsidence is first recorded along the eastern
al., 1995), suggesting that they form a horse-tail splay. In seaboard of Argentina and Uruguay, as well as in South
a right-lateral sense, the splay formed a restraining bend Africa. The occurrence of an extensive magmatic arc
along which the Dorsal was inverted; oblique slip and along the Pacific margin suggests that extension may
mild inversion characterized the Entre Lomas trend. have been facilitated by subduction pull and Atlantic
Northwest-striking transcurrent faults, albeit with mid-oceanic ridge push (Ziegler, 1993).
relatively small displacements, characterized the Paraná After the onset of spreading in the South Atlantic at
basin as well (Figure 28). A conspicuous example is the about 130 Ma (Rabinowitz and LaBrecque, 1979), intra-
Ponta Grossa arch (Renne et al., 1992). These northwest- continental rifting decreased. Extension persisted inter-
striking structures in the Paraná basin were most mittently in the San Jorge basin where it was associated
dynamic during the Cretaceous. with collapse of inversion rollovers (Cerdán et al., 1990)
40 Tankard et al.

and in the Salta rift complex where it was associated with


alkaline volcanism (Malumián et al., 1983), possibly
reflecting the influence of plate margin subduction.
Fault-driven subsidence was endemic to the Chile plate
margin after the Middle Jurassic (Charrier, 1984).

Tertiary Orogenesis
Compressive tectonics along the Andean belt started
in the early Cenozoic or even Late Cretaceous. However,
the climax of orogenesis was during the Miocene–
Pliocene, which is known colloquially as the Quechua
tectonic event. In Bolivia, Andean orogenesis encoun-
tered thick Phanerozoic accumulations in the Chaco
basin. This allowed the fold and thrust belt to surge
cratonward, forming the Chaco salient. These mountains
have much in common with the Mackenzie Mountains
of northern Canada (Eisbacher, 1985). Figure 18 shows
that along the Boomerang margin of the basin, Andean
folds developed as a series of ramp-flat associated struc-
tures above basement extensional fault blocks. Late
Tertiary contraction also resulted in marked uplift of the
Izozog arch and erosion of much of the Paleozoic cover
(Figure 20). A variety of structural styles resulted from
Andean deformation (Figure 21).
In the Sierras Pampeanas and pre-Cordillera,
northeast- and northwest-oriented basement faults were
reactivated to form a province of tilted block basins
much like those of Wyoming in the United States (Jordan
and Allmendinger, 1986; Fielding and Jordan, 1988).
Grier et al. (1991) and Salfity et al. (1993) describe the
Neogene inversion of Cretaceous rift basins in the Sierras
Pampeanas.
The importance of this Andean-age deformation to
petroleum geology is twofold. First, it was responsible
for very thick accumulations of terrigenous clastic
sediments that promoted maturation of various source
rocks. Depth to basement below the sub-Andean fold Figure 31—Comparative stratigraphic columns of Cuyo
and thrust belt of the Chaco basin is at least 12 km. In this and Lomas de Olmedo basins. Extensional subsidence of
belt, the Tertiary is as thick as 5 km. Second, many large the latter was related to a northwest-oriented shear couple.
oil and gas fields occur in folds above basement-involved These vectors resulted in inversion of the Cuyo basin
reverse faults. The scale of deformation spans a broad during this period (Figure 28). That these two stratigraphic
spectrum from subtle forced folds to the larger scale columns are mirror images of each other supports the
disruption of tilted fault blocks. tectonic reconstruction of southwest-directed extension in
the Triassic–Early Jurassic and northwest-directed
extension in the Late Jurassic–Cretaceous. (From
Legarreta et al., 1993; YPF, proprietary reports.)
DISCUSSION
This study has attempted to reconstruct the evolution stage Brasiliano–Pan-African cycle appears to
and tectonic linkage of the late Precambrian and have controlled subsequent basin development,
Phanerozoic basins of southwestern Gondwana. Figure pre–late Precambrian structural inheritance
32 summarizes the pattern of basin evolution. The most notwithstanding. Until the Permian–Triassic,
important conclusions to have emerged from this study basin development was largely intracratonic,
are as follows: while the edge of southwestern Gondwana was
little affected. The Permian–Triassic San Rafael
• The continental crust on which the Phanerozoic orogeny extended to the plate edge (e.g., inversion
basins developed was probably established by the of the Puncoviscana trend). The locus of Triassic
earliest Paleozoic at the latest. The structural extension by orogenic relaxation was generally
framework that controlled basin evolution largely closer to the margins. However, Late Jurassic–
cut across the Neoproterozoic terrane boundaries. Cretaceous extension was again located inboard
• The structural fabric formed during the multi- as it reverted to older structures.
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 41

• A set of northeast-oriented faults, transcurrent • Late Jurassic–Cretaceous extension related to


shear faults, or lineaments formed the backbone South Atlantic rifting was associated with
of this system. Many of these structures formed a northwest-trending right-lateral shear couples.
belt along which the Cretaceous South Atlantic The Triassic northeast transfers were reactivated
would eventually open. in a normal sense. The Late Jurassic onset of
• Late Precambrian–Early Cambrian basins were Atlantic rifting was locally translated into struc-
controlled mainly by these northeast-trending tural inversion of Triassic–Early Jurassic depo-
structures. Locally, transtensional pull-apart centers.
basins of various sizes progressed as far as oceanic • The rift basins beneath the continental shelf of
floor stage, as evidenced by ophiolites and eastern Argentina are also partially composite in
magmatic rocks in the Dom Feliciano, Ribeira, and form, preserving evidence of an earlier phase of
Matchless belts. However, the overall tectonic southwest-directed extension that was later
linkage of these and subsequent basins remained replaced by a northwest-oriented stress field in
until the Late Jurassic. the Late Jurassic–Cretaceous (e.g., San Julian
• Regional stress fields and the sense of motion on basin) (P. V. Zalán and A. M. de Figueiredo, 1994,
these inherited structures changed systematically. personal communication).
These periodic reorganizations probably
coincided with orogenic episodes, regional dias- The way in which the long-established structural
trophisms, and prominent unconformities. Mile- framework responded to changes in horizontal devia-
stones included the Ocloyic orogeny of Late toric stress in the lithosphere appears to be reasonably
Ordovician–Early Silurian age, the Chañic predictable. Less clear are the reasons for these periodic
orogeny of latest Devonian–Early Carboniferous changes in lithospheric stress and the sources of wide-
age, the San Rafael orogeny of middle Permian– spread tension. Orogenic events alternated with periods
Middle Triassic age, and the Araucanian orogeny of extension and wrench faulting throughout the
of latest Oxfordian–early Kimmeridgian age. Phanerozoic. During the Triassic, extension resulted in
• Each intervening stage of basin formation appears destruction of the irregular and discontinuous fold belts.
to have involved an early history of fault- In the Late Jurassic–Neocomian, a continuous magmatic
controlled subsidence and a later episode when arc along the Pacific margin co-existed with widespread
subsidence was more uniform, possibly due to extension and opening of the South Atlantic (Jensen et
relaxation of extensional stresses or thermal al., 1976; Charrier, 1984; Mpodozis, 1984). Furthermore,
contraction. the direction of Mesozoic extension was at a high angle
• Throughout the Paleozoic, the Puna-Pampeanas to the contemporaneous Pacific margins (Figure 32),
arch appears to have formed a persistent tectonic suggesting linkage between the plate boundary and
high and divide, resembling a giant pop-up intracontinental extension (Storey et al., 1992). Plate
structure on either side of which there were sedi- boundary forces such as subduction slab pull and ridge
mentary basins. Sedimentologic evidence push, as well as collision processes, are believed to
suggests that this structural morphology extended control intraplate tectonism (Forsyth and Uyeda, 1975;
south of South Africa. Dewey, 1988; Bott, 1992). De Wit and Ransome (1992)
• Following the Chañic orogeny, the Carboniferous discuss the influence of stress transmission related to the
was a time of widespread extension. Dextral fusion and fission of supercontinents.
strike-slip along the Puna-Pampeanas system was Following the Chañic diastrophism, the Early
associated with pull-apart basins (e.g., Paganzo- Carboniferous basins of western Argentina and Chile
Tarija-Chaco basin tract). Elsewhere, the developed as transtensional pull-aparts as a result of
northeast-trending accommodation zones of times right-lateral shearing due to oblique convergence
past were reactivated as normal faults. (Rapalini, 1990; Breitkreuz, 1991; Fernandez-Seveso and
• The Late Ordovician–earliest Silurian Ocloyic Tankard, 1995).
orogeny and the Late Devonian–Early Carbonif- Regardless of the precise causative mechanisms
erous Chañic orogeny coincided with Gondwana whereby the structural framework of southwestern
glaciations. This is attributed to the climatic effects Gondwana responded to changing plate boundary
of tectonic arching. stresses (e.g., Leighton and Kolata, 1990), a global
• Phanerozoic extensional depocenters have been tectonic control on the subsidence histories of widely
interpreted as isolated rift segments offset across separated but structurally interlinked basins in south-
accommodation zones. Permian–Triassic San western Gondwana is implied. This recognition paves
Rafael contraction is believed to have inverted the way for correlations of large-scale depositional
these basin segments to form discontinuous but sequences. Recent plate reconstructions by Dalziel et al.
tectonically linked fold belts. (1994), for example, confirm the striking coincidence in
• Triassic extension and magmatism straddled this time between the Sauk–Tippecanoe and Tippecanoe–
discontinuous mountain belt. Southwest-directed Kaskaskia sequence boundaries of the Laurentian craton
Basin and Range type extensional collapse of the with the Taconic and Acadian deformations along the
orogen was guided by northeast-striking eastern margin of North America. Correlative tectonism
transfers. in South America (the Ocloyic and Chañic orogenies)
42 Tankard et al.

VE N D IA N - C A M B R IA N CA M B R IA N - O R D O V IC IA N
N BF CH TB CH
CHACO
GF
PT
ET
OF
3 PA RANA
PF JF
3 PF

AC LB
LF LA

ML

KF
ZF
CA PE
N

DE V O N IA N CA R B O N IFE R O U S - P E R M IA N
BF
CH

CHACO 3 CH AC O
HI G H

3
PAR A NA
HI
G
H

LB CHACO -
PA RANA
HI

HI G
G
H

D
AN
H

HL
G
HI

HIGH
KA L A H AR I

HI
GH
HI

LA
GH

CAPE ND
LA

KA RO O
ND

Marine Incursion
Ice-Flow Direction

Figure 32—Reconstructions of the tectonic and paleogeographic framework for the principal stages of basin development.
Thrusts on Triassic–Jurassic map refer to Permian–Early Triassic contraction; they show the relationship to Triassic
extension by orogenic collapse. AC, Aconquija lineament; AL, Altiplano basin; AM, Agulhas-Malvinas fracture zone; BF,
Boomerang fault; BL, Bariloche; CFB, Cape fold belt; CH, Chiquitanas front; CO, Colorado basin; CU, Cuyo basin; ET, El
Toro lineament; GF, Goiânia belt; GFS, Gastre fault system; JF, Jacutinga fault; IA, Izozog arch; KL, Kalahari line; LB, Las
Breñas fault; LF, Lancinha-Cubatão fault; LO, Lomas de Olmedo basin; MA, Malvinas basin; MG, Magallanes basin; NQ,
Neuquén basin; OF, Ocloyic fault; OU, Outeniqua basin; PT, Puna trough; RA, Rawson basin; SA, Salado basin; SG, Sierra
Grande; SJ, San Jorge basin; SL, Santa Lucía basin; SP, Sierra Pintada; SS, Sierras Septentrionales; SV, Sierra de la
Ventana; TB, Trans-Brasiliano lineament; WS, Weddell Sea; ZF, Zoetfontein fault.
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 43

TR IA S SIC - E AR LY J U RA S SI C L AT E J UR A S SIC - C R ETA C EO U S


N
AL

IA

LO

CU
3 CU
SL
NQ
SP

SV

SA
SS

N
BL
SG CO

G
FS
RA
SJ

OU
CF
B

MG MA AM

WS

Figure 32 (continued)

suggests that broader interregional correlations can now Paraná basin, the Irati Formation varies from
be attempted. immature at shallow depths to mature or overma-
The repetitive history of basin development and struc- ture where it is associated with diabase dikes.
tural modification also influenced the petroleum geology Thick sequences of Patquía de la Cuesta source
and prospectivity of the region. This history involved rock shales occur in the Paganzo basin, but they
multiple phases of orogenesis, fault-controlled exten- have become involved in the Tertiary tilt block
sional subsidence, and decay or relaxation of extensional basin province and are deeply buried in places. The
stresses (Bally and Snelson, 1980), which resulted in Paganzo basin is largely unexplored.
uniform subsidence of broad epicontinental basins. 5. Several prolific oil-prone source rocks occur in the
There were several episodes of hydrocarbon source rock Triassic, Early Jurassic, and Early Cretaceous rifts.
deposition, each of which broadly coincided with the
culminating episodes of regional subsidence character- The importance of the sub-Andean basins is that they
ized by restricted marine circulation and comparatively are long-lived and composite in form and contain
low rates of influx of terrigenous sediments. There are multiple source rock accumulations. All of the commer-
several source rock intervals: cially producing basins in Argentina and Bolivia
developed above Paleozoic depocenters. The late
1. An uppermost Proterozoic interval in the Nama Tertiary Andean orogeny provided a thick cover
and São Francisco basins is little more than a fasci- sequence that enhanced burial maturation and
nating curiosity. migration, as well as creating traps by structural
2. Potential Silurian source rocks occur in the Tarija inversion. Most producing oil and gas fields in Argentina
and Chaco basins, and the Kirusillas shales sourced have an inversion component of trap formation.
Sara oils in the Boomerang area. In contrast, many of the rift basins of Argentina,
3. The Devonian Los Monos shale is the principal Uruguay, and offshore South Africa formed along the
source of hydrocarbons in the Tarija and Chaco tract of Permian–Triassic fold belts by orogenic collapse.
basins. Its counterparts in the Paraná basin are the They are attractive exploration areas because they are
Lima and Ponta Grossa formations. well structured and may contain prolific lacustrine
4. Probably the most enigmatic is the Upper Permian source rocks. An example is the Cuyo basin with its
(Kazanian) interval. It is characterized by high TOC Upper Triassic Cacheuta source rock. However, orogenic
values and kerogen types I and II. The White Hill collapse also presents the risk of an overabundance of
Formation in the Karoo basin is overmature. In the coarse material.
44 Tankard et al.

Acknowledgments We thank R. Allen, H. R. Balkwill, G. Bahlburg, H., C. Breitkreuz, and W. Zeil, 1987, Paleozoic basin
Bonorino, H. Di Benedetto, E. J. Ebner, F. Fernandez-Seveso, development in northern Chile (21°–27°S): Geologische
M. Fowler, P. G. Gresse, W. Hegenberger, G. Henry, N. Rundschau, v. 76, p. 633–646.
Hiller, M. P. A. Jackson, W. A. M. Jenkins, A. H. Knoll, P. Baldis, B. A., M. S. Beresi, O. Bordonaro, and A. Vaca, 1982,
Síntesis evolutiva de la pre-Cordillera Argentina: 5
Kress, O. R. López-Gamundí, J. V. Ploszkiewicz, J. A. Salfity,
Congreso Latinamericano de Geología, Argentina, Actas,
F. Schein, L. A. Spalletti, I. Stanistreet, J. N. Theron, R. v. 1, p. 99–445.
Unrug, C. M. Urien, J. Utting, B. van Hoorn, R. Viñes, J. N. Baldis, B. A., M. Beresi, O. Bordonaro, and A. Vaca, 1984, The
J. Visser, and P. A. Ziegler for helpful discussions and assis- Argentine Procordillera as a key to Andean structure:
tance at various stages of this project. Elf Aquitaine kindly Episodes, v. 7, p. 14–19.
provided lithologic logs of the Tses-1 well in Namibia. Typing Baldis, B., O. Bordonaro, C. Armella, M. Beresi, N. Cabaleri, S.
was done by K. Bojarski. Peralta, and H. Bastias, 1989, La cuenca Paleozoica inferior
de la pre-Cordillera Argentina, in G. Chebli and L. Spalleti,
eds., Cuencas sedimentarias Argentinas: Tucumán,
Universidad Nacional de Tucumán, Serie correlación
geológica 6, p. 101–121.
REFERENCES CITED Balkwill, H. R., G. Rodrigue, F. I. Paredes, and J. P. Almeida,
1995, Northern part of Oriente basin, Ecuador: reflection
Aceñolaza, F. G., and F. R. Durand, 1986, Upper Precam- seismic expression of structures, in A. J. Tankard, R.
brian–Lower Cambrian biota from the northwest of Suarez, and H. J. Welsink, Petroleum basins of South
Argentina: Geological Magazine, v. 123, p. 367–375. America: AAPG Memoir 62, this volume.
Aceñolaza, F. G., and A. J. Toselli, 1988, El sistema de Bally, A. W., and S. Snelson, 1980, Realms of subsidence, in A.
Famatina: su interpretación como orógeno de margen D. Miall, ed., Facts and principles of world petroleum
continental activo: 5 Congreso Geológico Chileno, occurrence: Canadian Society of Petroleum Geologists
Santiago, v. 1, p. A55–A67. Memoir 6, p. 9–94.
Almeida, F. F. M., and Y. Hasui, 1984, O PréCambriano do Bate, K. J., and J. A. Malan, 1992, Tectonostratigraphic
Brasil: São Paulo, Edgard Blücher, 378 p. evolution of the Algoa, Gamtoos, and Pletmos basins,
Almeida, F. F. M., Y. Hasui, and B. B. de Brito Neves, 1976, offshore South Africa, in M. J. De Wit and I. G. D.
The Upper Precambrian of South America: Boletim Ransome, eds., Inversion tectonics of the Cape fold belt,
Instituto Geosciencias, Univ. São Paolo, v. 7, p. 5–80. Karoo and Cretaceous basins of southern Africa:
Almeida, F. F. M., Y. Hasui, B. B. de Brito Neves, and R. A. Rotterdam, Balkema, p. 61–73.
Fuck, 1981, Brazilian structural provinces: an introduction: Bell, C. M., 1987, The late Paleozoic evolution of the Gond-
Earth Science Reviews, v. 17, p. 1–29. wanaland continental margin in northern Chile, in G. D.
Alvarenga, C. J. S., and R. Trompete, 1992, Glacially influ- McKenzie, ed., Gondwana six: Structure, tectonics, and
enced sedimentation in the later Proterozoic of the geophysics: American Geophysical Union Geophysical
Paraguay belt (Mato Grosso, Brazil): Palaeogeography, Monograph 40, p. 261–270.
Palaeoclimatology, Palaeoecology, v. 92, p. 85–105. Bernasconi, A., 1987, The major Precambrian terranes of
Amos, A. J., 1972, Las cuencas carbónicas y pérmicas de la eastern South America: a study of their regional and
Argentina: Anales Academia Brasileira de Ciencias, v. 44, chronological evolution: Precambrian Research, v. 37,
p. 21–36. p. 107–124.
Anderson, A. M., and I. R. McLachlan, 1979, The oil-shale Berry, W. B. N., and A. J. Boucot, 1973, Correlation of the
potential of the Early Permian White Band Formation in African Silurian rocks: GSA Special Paper 147, p. 1–83.
southern Africa, in A. M. Anderson and W. J. van Biljon, Bianchi, J. L., 1984, Interpretación tectogenética y paleoambi-
eds., Some sedimentary basins and associated ore deposits ental de la cuenca de Rawson, plataforma continental
of South Africa: Geological Society of South Africa Special Argentina: Noveno Congreso Geológico Argentino,
Publication 6, p. 83–89. Bariloche, Actas 3, p. 47–60.
Anderson, D. L., Y. S. Zhang, and T. Tanimoto, 1992, Plume Bianucci, H. A., and J. F. Homovc, 1982, Tectogénesis de un
heads, continental lithosphere, flood basalts and tomog- sector de la cuenca del Subgrupo Pirgua, Noroeste
raphy, in B. C. Storey, T. Alabaster, and R. J. Parkhurst, Argentino: Quinto Congreso Latino-Americano de
eds., Magmatism and the causes of continental break-up: Geología, Actas 1, p. 539–546.
Geological Society of London, Special Publication 68, Bott, M. H. P., 1992, The stress regime associated with conti-
p. 99–124. nental break-up, in B. C. Storey, T. Alabaster, and R. J.
Archangelsky, S., C. Azcuy, C. Gonzalez, and N. Sabattini, Pankhurst, eds., Magmatism and the causes of continental
1986, Correlación general de biozonas, in S. Archangelsky, break-up: Geological Society of London, Special Publica-
ed., El Sistema Carbonífero en la República Argentina: tion 68, p. 125–136.
Academia Nacional de Ciencias de Córdoba, p. 265–287. Boucot, A. J., 1971, Malvinokaffric Devonian marine
Avila-Salinas, W. A., 1986, Magmatismo cretácico en Bolivia: community distribution and implications for Gondwana:
PIGC-Proyecto 242 Cretácico de America Latina: Primer Anais da Academia Brasileira de Ciencias, v. 43 (suple-
Simposio, La Paz, p. 52–70. mento), p. 23–49.
Azcuy, C. L., and R. Caminos, 1986, Diastrofismo, in S. Braun, O. P. G., U. Mello, and H. Della Piazza, 1991, Bacias
Archangelsky, ed., El Sistema Carbonífero en la República Proterozoicas brasileiras com perspectivas exploratórias
Argentina: Academia Nacional de Ciencias de Córdoba, para hidrocarbonetos, in G. P. Raja Gabaglia and E. J.
p. 239–251. Milani, eds., Origem e Evolução de Bacias sedimentares:
Bahlburg, H. 1990, The Ordovician basin in the Puna of NW Rio de Janeiro, Petrobras, p. 115–132.
Argentina and N Chile: geodynamic evolution from back- Breitkopf, J. H., and K. J. Maiden, 1987, Geochemical patterns
arc to foreland basin: Geotektonische Forschungen, v. 75, of metabasites in the southern part of the Damara orogen,
p. 1–107. SWA/Namibia: applicability to the recognition of tectonic
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 45

environment, in T. C. Pharoah, R. D. Beckinsale, and D. G. D. Ransome, eds., Inversion tectonics of the Cape fold
Rickard, eds., Geochemistry and mineralization of Protero- belt, Karoo and Cretaceous basins of southern Africa:
zoic volcanic suites: Geological Society of South Africa Rotterdam, Balkema, p. 239–248.
Special Publication 33, p. 355–361. Cobbold, P., A. C. Massabie, and E. A. Rossello, 1986,
Breitkreuz, C.,1991, Permo-Carboniferous magmatism, basin Hercynian wrenching and thrusting in the Sierras
development, and tectonics in the north Chilean Andes Australes foldbelt, Argentina: Hercynica II, v. 2, p. 135–148.
(abs.): XII International Congress on Carboniferous and Cobbold, P. R., D. Gapais, E. R. Rossello, E. J. Milani, and P.
Permian Geology and Stratigraphy, Abstracts, p. 17–18. Szatmari, 1992, Permian–Triassic intracontinental deforma-
Brito Neves, B. B. de, 1991, Processos orogênicos no Pré- tion in SW Gondwana, in M. J. De Wit and I. G. D.
Cambriano do Brasil, in G. P. Raja Gabaglia and E. J. Ransome, eds., Inversion tectonics of the Cape fold belt,
Milani, eds., Origem e Evolução de Bacias Sedimentares: Karoo and Cretaceous basins of Southern Africa:
Rio de Janeiro, Petrobras, p. 99–114. Rotterdam, Balkema, p. 23–26.
Brito Neves, B. B. de, R. A. Fuck, U. G. Cordani, and T. Filho, Cocks, L. R. M., C. H. C. Brunton, A. J. Rowell, and I. C. Rust,
1984, Influence of basement structures on the evolution of 1970, The first lower Palaeozoic fauna from South Africa:
the major sedimentary basins of Brazil: a case of tectonic Geological Society of London Quarterly Journal, v. 125,
heritage: Journal of Geodynamics, v. 1, p. 495–510. p. 583–603.
Brito Neves, B. B. de, and U. G. Cordani, 1991, Tectonic Coira, B. L., F. Nullo, C. Prosperpio, and V. A. Ramos, 1975,
evolution of South America during the Late Proterozoic: Tectónica de basamento de la región occidental del Macizo
Precambrian Research, v. 53, p. 23–40. Norpatagónico: Revista Asociación Geológica Argentina,
Bruhn, R. L., C. R. Stern, and M. J. De Wit, 1978, The bearing v. 30, p. 361–383.
of new field and geochemical data on the origin and devel- Coira, B., J. Davidson, C. Mpodozis, and V. Ramos, 1982,
opment of a volcano-tectonic rift zone and back arc basin Tectonics and magmatic evolution of the Andes of
in southernmost South America: Earth and Planetary northern Argentina and Chile: Earth Science Reviews, v.
Science Letters, v. 41, p. 32–46. 18, p. 303–332.
Caminos, R., 1979, Sierras Pampeanas noroccidentales, Salta, Cole, D. I, 1992, Evolution and development of the Karoo
Tucumán, Catamarca, La Rioja y San Juan, in Simposio de basin, in M. J. De Wit and I. G. D. Ransome, eds., Inversion
Geología Regional Argentina: Academia Nacional de tectonics of the Cape fold belt, Karoo and Cretaceous
Ciencias, v. 1, p. 225–292. basins of southern Africa: Rotterdam, Balkema, p. 87–99.
Caminos, R., J. Llambias, C. W. Rapela, and C. A. Parica, 1988, Cominguez, A. H., and V. A. Ramos, 1991, La estructura
Late Paleozoic–Early Triassic magmatic activity of profunda entre pre-Cordillera y Sierras Pampeanas de la
Argentina and the significance of new Rb-Sr ages from Argentina: evidencias de la sísmica de reflexión profunda:
northern Patagonia: Journal of South American Earth Revista Geológica de Chile v. 18, p. 3–14.
Sciences, v. 1, p. 137–145. Cominguez, A. H., and V. A. Ramos, 1995, Geometry and
Caputo, M. V., 1985, Late Devonian glaciation in South seismic expression of the Cretaceous Salta rift system of
America: Palaeogeography, Palaeoclimatology, Palaeoe- northwestern Argentina, in A. J. Tankard, R. Suarez, and
cology, v. 51, p. 219–317. H. J. Welsink, Petroleum basins of South America: AAPG
Cerdán, J. J., J. S. Galeazzi, J. D. Enrique, A. I. Jouglard, C. J. Memoir 62, this volume.
Perrot, and M. A. Uliana, 1990, Sísmica de detalle mas Copper, P, 1977, Paleolatitudes in the Devonian of Brazil and
interpretación integrada: claves para el hallazgo de nuevas the Frasnian–Famennian mass extinction: Palaeogeog-
reservas en áreas maduras: Tercer Congreso Andino de la raphy, Palaeoclimatology, Palaeoecology, v. 21, p. 165–207.
Industria del Petróleo, Memorias, Tomo1, p. 27–67. Cordani, U. G., and B. B. de Brito Neves, 1982, The geologic
Charrier, R., 1984, Areas subsidentes en el borde occidental de evolution of South America during the Archaean and
la cuenca tras-arco Jurásico–Cretácica, Cordillera Principal Early Proterozoic: Revista Brasiliano Geosciencias, v. 12,
Chilena entre 34° y 34°30’S: Noveno Congreso Geológico p. 78–88.
Argentino, Actas 2, p. 107–124. Corner, B., 1991, Crustal architecture of the Precambrian of
Chebli, G. A., O. Tofalo, and G. E. Turzzini, 1989, the Kaapvaal province from geophysical data (extended
Mesopotamia, in G. A. Chebli and L. A. Spalletti, eds., abs.): Terra Nova, v. 3, p. 5.
Cuencas sedimentarias Argentinas: Universidad Nacional Cortés, J. M., R. Caminos, and H. A. Leanza, 1984, La
de Tucumán, Serie correlación geológica 6, p. 79–100. cobertura sedimentaria eopaleozoica: IX Congreso
Christie-Blick, N., and K. T. Biddle, 1985, Deformation and Geológico Argentino, Bariloche, Relatorio 1, p. 65–84.
basin formation along strike-slip faults, in K. T. Biddle and Covey, M., 1986, The evolution of foreland basins to steady
N. Christie-Blick, eds., Strike-slip deformation, basin state: evidence from the western Taiwan foreland basin, in
formation, and sedimentation: SEPM Special Publication P. A. Allen and P. Homewood, eds., Foreland basins: Inter-
37, p. 1–34. national Association of Sedimentologists Special Publica-
Cingolani,C., R. Varela, L. Dalla Salda, and K. Kawashita, tion 8, p. 77–90.
1993, Los granitoides del Cerro Veladero, Río de la Troya, Dalla Salda, L. H., I. W. D. Dalziel, C. A. Cingolani, and R.
Provincia de la Rioja: estudio geocronológico e implican- Varela, 1992, Did the Taconic Appalachians continue into
cias tectónicas: XII Congreso Geológico Argentino, II southern South America?: Geology, v. 20, p. 1059–1062.
Congreso de Exploración de Hidrocarburos, Actas, v. 4, Dalla Salda, L. H., R. Varela, and C. Cingolani, 1993, Sobre la
p. 68–74. colision de Laurentia–Sudamerica y el orogeno fama-
Cisternas, M. E., 1979, Litofacies de transición marino-conti- tiniano: XII Congreso Geológico Argentino y II Congreso
nental en el Jurásico del sector La Ola, al sur del Salar de de Exploración de Hidrocarburos Actas, v. 3, p. 358–366.
Pedernales: Actas Congreso Geológica Chileno 2, Arica, v. Dalmayrac, B., G. Lambacher, R. Marocco, C. Martinez, and B.
1, p. A65–A85. Tomasi, 1980, La chaine hercynienne d’Amérique du Sud-
Cloetingh, S., A. Lankreijer, M. J. De Wit, and I. Martinez, Structure et évolution d’orogéne intracratonique: Geolo-
1992, Subsidence history analysis and forward modelling gische Rundschau, v. 69, p. 1–21.
of the Cape and Karoo supergroups, in M. J. De Wit and I.
46 Tankard et al.

Daly, M. C., J. Chorowicz, and J. D. Fairhead, 1989, Rift basin Eyles, N., 1993, Earth’s glacial record and its tectonic setting:
evolution in Africa: the influence of reactivated steep Earth-Science Reviews, v. 35, p. 1–248.
basement shear zones, in M. A. Cooper and G. D. Williams, Eyles, N., and C. H. Eyles, 1993, Glacial geologic confirmation
eds., Inversion tectonics: Geological Society of London, of an intraplate boundary in the Paraná basin of Brazil:
Special Publication 44, p. 309–334. Geology, v. 21, p. 459–462.
Daly, M. C. S. R. Lawrence, D. Kimun’a, and M. Binga, 1991, Eyles, N., G. Gonzalez Bonorino, A. B. França, C. H. Eyles,
Late Paleozoic deformation in central Africa: a result of and O. Lopez, 1995, Hydrocarbon-bearing late Paleozoic
distant collision?: Nature, v. 350, p. 605–607. glaciated basins of southern and central South America, in
Dalziel, I. W. D., L. H. Dalla Salda, and L. M. Gahagan, 1994, A. J. Tankard, R. Suarez, and H. J. Welsink, Petroleum
Paleozoic Laurentia–Gondwana interaction and the origin basins of South America: AAPG Memoir 62, this volume.
of the Appalachian Andean mountain system: GSA Fernandez-Seveso, F., and A. J. Tankard, 1995, Tectonics and
Bulletin, v. 106, p. 243–252. stratigraphy of the late Paleozoic Paganzo basin, western
Dardenne, M. A., 1978, Síntese sobre a estratigrafia do Bambuí Argentina, and its regional implications, in A. J. Tankard,
no Brasil Central: Anais do 30 Congreso Brasileiro de R. Suarez, and H. J. Welsink, Petroleum basins of South
Geologia, v. 2, p. 597–610. America: AAPG Memoir 62, this volume.
Dardenne, M. A., 1981, Os grupos Paranoa e Bambuí na faixa Fielding, E. J., and T. E. Jordan, 1988, Active deformation at
Dobrada Brasília: Simpósio sobre o Craton do São the boundary between the pre-Cordillera and Sierras
Francisco e suas Faixas Marginais, Anais, SBG. Nucleo de Pampeanas, Argentina, and comparison with ancient
Bahia–CPM-SME, p. 140–157. Rocky Mountain deformation: GSA Memoir 171, p.
De Beer, J. H., 1983, Geophysical studies in the southern Cape 143–163.
Province and models of the lithosphere in the Cape Fold Finnemore, S. H., 1978, The geochemistry and origin of the
Belt, in A. P. G. Söhnge and I. W. Hälbich, eds., Geody- Matchless amphibolite belt, Windhoek District, South West
namics of the Cape fold belt: Geological Society of South Africa, in W. J. Verwoerd, ed., Mineralization in metamor-
Africa Special Publication 12, p. 57–64. phic terranes: Geological Society of South Africa Special
De Beer, C. H., 1990, Simultaneous folding in the western and Publication 4, p. 433–477.
southern branches of the Cape fold belt: South African Fletcher, C. J. N., and M. Litherland, 1981, The geology and
Journal of Geology, v. 93, p. 583–591. tectonic setting of the Velasco alkaline province, eastern
Dellapé, D. A., and A. G. Hegedus, 1995, Structural inversion Bolivia: Journal of the Geological Society of London, v. 138,
and oil occurrence in the Cuyo basin of Argentina, in A. J. p. 541–548.
Tankard, R. Suarez, and H. J. Welsink, Petroleum basins of Forsyth, R. D., and S. Uyeda, 1975, On the relative importance
South America: AAPG Memoir 62, this volume. of the driving forces of plate motion: Geophysical Journal
De Villiers, J., 1956, Die drie sintaksisse in die suidwestelike of the Royal Astronomical Society, v. 43, p. 163–200.
Kaapprovinsie: Tegnikon, v. 9., p. 75–87. Forsythe, R. D., J. Davidson, C. Mpodozis, and C. Jesinkey,
De Villiers, J., H. Jansen, and M. P. Mulder, 1964, Die geologie 1993, Lower Paleozoic relative motion of the Arequipa
van die gebied tussen Worcester en Hermanus: Pretoria, block and Gondwana; paleomagnetic evidence from Sierra
Republic of South Africa, Geological Survey, 69 p. de Almeida of northern Chile: Tectonics, v. 12, p. 219–255.
Dewey, J. F., 1988, Extensional collapse of orogens: Tectonics, Fouché, J., K. J. Bate, and R. van der Merwe, 1992, Plate
v. 7, p. 1123–1139. tectonic setting of the Mesozoic basins, southern offshore,
De Wit, M. J., 1992, The Cape fold belt: a challenge for an inte- South Africa: a review, in M. J. De Wit and I. G. D.
grated approach to inversion tectonics, in M. J. De Wit and Ransome, eds., Inversion tectonics of the Cape fold belt,
I. G. D. Ransome, eds., Inversion tectonics of the Cape fold Karoo and Cretaceous basins of southern Africa:
belt, Karoo and Cretaceous basins of southern Africa: Rotterdam, Balkema, p. 33–45.
Rotterdam, Balkema, p. 3–12. França, A. B., E. J. Milani, R. L. Schneider, O. López, M. López,
De Wit, M. J., and I. G. D. Ransome, 1992, Regional inversion R. Suarez, H. de Santa Ana, F. Wiens, O. Ferreiro, E. A.
tectonics along the southern margin of Gondwana, in M. J. Rossello, et al., 1995, Phanerozoic correlation in southern
De Wit and I. G. D. Ransome, eds., Inversion tectonics of South America, in A. J. Tankard, R. Suarez, and H. J.
the Cape fold belt, Karoo and Cretaceous basins of Welsink, Petroleum basins of South America: AAPG
southern Africa: Rotterdam, Balkema, p. 15–21. Memoir 62, this volume.
De Wit, M. J., M. Jeffery, H. Bergh, and L. Nicolaysen, 1988, Germs, G. J. B., 1974, The Nama Group in South West Africa
Geological map of sectors of Gondwana: American Associ- and its relationship to the Pan-African geosycline: Journal
ation of Petroleum Geologists map, 2 sheets, 1:10,000,000. of Geology, v. 82, p. 301–317.
De Wit, M. J., C. Roering, R. J. Hart, R. A. Armstrong, C. E. J. Germs, G. J. B., 1983, Implications of a sedimentary facies and
DeRonde, R. W. E. Green, M. Tredoux, E. Peberdy, and R. depositional environmental analysis of the Nama Group in
A. Hart, 1992, Formation of an Archaean continent: South West Africa/Namibia, in R. McG. Miller, ed.,
Nature, v. 357, p. 553–562. Evolution of the Damara orogen of South West
Dingle, R. V., W. G. Siesser, and A. R. Newton, 1983, Mesozoic Africa/Namibia: Geological Society of South Africa Special
and Tertiary geology of southern Africa: Rotterdam, Publication 11, p. 89–114.
Balkema, 375 p. Germs, G. J. B., and P. G. Gresse, 1991, The foreland basin of
Du Toit, A. L., 1927, A geologic comparison of South America the Damara and Gariep orogens in Namaqualand and
with South Africa., with a paleontological contribution by southern Namibia: stratigraphic correlations and basin
F. Cowper Reid: Washington, D.C., Carnegie Institute, dynamics: South African Journal of Geology, v. 94, p.
381 p. 159–169.
Eisbacher, G. H., 1985, Pericollisional strike-slip faults and Gibbs, A. D., 1984, Structural evolution of extensional basin
synorogenic basins, Canadian Cordillera, in K. T. Biddle margins: Journal of the Geological Society (London), v.
and N. Christie-Blick, eds., Strike-slip deformation, basin 141, p. 609–620.
formation, and sedimentation: SEPM Special Publication Giraldo, C., and S. Osuna, 1994, Excursión al frente surandino
37, p. 265–282. entre Barinas y Guanare: V Simposio Bolivariano Explo-
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 47

ración Petrolera en las Cuencas Subandinas, Excursión no. between Bushmanland and Gordonia subprovinces: Ph.D.
1, p. 1–31. dissertation, University of Cape Town, South Africa, 285 p.
Gonzalez Bonorino, G., 1991, Late Paleozoic orogeny in the Hartnady, C. J., 1978, The structural geology of the Naukluft
northwestern Gondwana continental margin, western nappe complex and its relationship to the Damara
Argentina and Chile: Journal of South American Earth orogenic belt, South West Africa: Ph.D. dissertation,
Science, v. 4, p. 131–144. University of Cape Town, South Africa, 256 p.
Gonzalez Bonorino, G., and F. Gonzalez Bonorino, 1991, Pre- Hartnady, C., P. Joubert, and C. Stowe, 1985, Proterozoic
Cordillera de Cuyo y Cordillera Frontal en el Paleozoico crustal evolution in southwestern Africa: Episodes, v. 8, p.
temprano: terrenos ‘Bajo sospecha’ de ser autoctonos: 236–244.
Revista Geológica de Chile, v. 18, p. 97–107. Hawkesworth, C. J., K. Gallagher, S. Kelley, M. Mantovani, D.
Gordillo, C. E., 1984, Migmatitas cordieríticas de la Sierra de W. Peate, M. Regelous, and N. W. Rogers, 1992, Paraná
Cordóba: condiciones físicas de migmatización: Academia magmatism and the opening of the South Atlantic, in B. C.
Nacional de Ciencias, Miscelánea, no. 68, p. 1–37. Storey, T. Alabaster, and R. J. Pankhurst, eds., Magmatism
Gravenor, G. P., and A. C. Rocha-Campos, 1983, Patterns of and the causes of continental break-up: Geological Society
late Paleozoic glacial sedimentation on the southeast side of London, Special Publication 68, p. 221–240.
of the Paraná basin, Brazil: Palaeogeography, Palaeoclima- Henry, G., C. W. Clendenin, I. G. Stanistreet, and K. J. Maiden,
tology, Palaeoecology, v. 43, p. 1–39. 1990, Multiple detachment model for the early rifting stage
Green, R. W. E., 1983, Seismic refraction observations in the of the late Proterozoic Damara orogen in Namibia:
Damara orogen and flanking craton and their bearing on Geology, v. 18, p. 67–71.
deep seated processes in the orogen, in R. McG. Miller, ed., Hiller, N., 1992, The Ordovician system in South Africa: a
Evolution of the Damara orogen of South West review, in B. D. Webby and J. R. Laurie, eds., Global
Africa/Namibia: Geological Society of South Africa Special perspectives on Ordovician geology: Rotterdam, Balkema,
Publication 11, p. 355–367. p. 473–485.
Green, D., R. N. Crockett, and M. T. Jones, 1980, Tectonic Hiller, N., and J. N. Theron, 1988, Benthic communities in the
control of Karoo sedimentation in mid-eastern Botswana: South African Devonian, in N. J. McMillan, A. F. Embry
Transactions of the Geological Society of South Africa, v. and D. J. Glass, eds., Devonian of the World: Canadian
83, p. 213–219. Society of Petroleum Geologists Memoir 14, p. 229–242.
Gresse, P. G., 1986, The tectono-sedimentary history of the Hoffmann, K. H., 1984, Direction of tectonic transport in the
Vanrhynsdorp Group: Ph.D. dissertation, University of southern margin thrust belt, Damara belt, SWA/Namibia
Stellenbosch, South Africa, 155 p. (abs.): Conference on middle to late Proterozoic Lithos-
Gresse, P. G., and G. J. B. Germs, 1993, The Nama foreland phere Evolution, Abstracts, p. 51.
basin: sedimentation, major unconformity bounded Hutchins, D. G., and C. V. Reeves, 1980, Regional geophysical
sequences, and multisided active margin advance: Precam- exploration of the Kalahari in Botswana: Tectonophysics,
brian Research, v. 63, p. 247–272. v. 69, p. 201–220.
Gresse, P. G., and R. Scheepers, 1993, Neoproterozoic to Jensen, O. L., J. C. Vicente, J. Davidson, and E. Goday, 1976,
Cambrian (Namibian) rocks of South Africa: a geochrono- Etapás de la evolución marina Jurásica de la cuenca
logical and geotectonic review: Journal of African Earth Andina externa (mioliminar) entre los paralelos 26° y
Sciences, v. 16, p. 375–393. 29°30’ sur: Primero Congreso Geológico Chileno, Actas 1,
Grier, M., J. A. Salfity, and R. W. Allmendinger, 1991, Andean p. A273–A295.
reactivation of the Cretaceous Salta rift, northwestern Jezek, P., and H. Miller, 1987, Petrology and facies analysis of
Argentina: Journal of South American Earth Sciences, v. 4, turbiditic sedimentary rocks of the Puncoviscana trough
p. 351–372. (upper Precambrian–Lower Cambrian) in the basement of
Grunow, A. M., 1993, Creation and destruction of Weddell the NW Argentine Andes, in G. D. McKenzie, ed.,
Sea floor in the Jurassic: Geology, v. 21, p. 647–650. Gondwana six: structure, tectonics, and geophysics:
Hälbich, I. W., 1983a, Disharmonic folding, detachment and American Geophysical Union, Geophysical Monograph 40,
thrusting in the Cape fold belt, in A. P. G. Söhnge and I. W. p. 287–293.
Hälbich, eds., Geodynamics of the Cape fold belt: Geolog- Jones, C. H., B. P. Wernicke, L. G. Farmer, D. S. Coleman, L.
ical Society of South Africa Special Publication 12, p.
W. McKenna, and F. V. Perry, 1992, Variations across and
115–123.
along a major continental rift: an interdisciplinary study of
Hälbich, I. W., 1983b, A tectogenesis of the Cape fold belt
the Basin and Range province, western U.S.A. : Tectono-
(CFB), in A. P. G. Söhnge and I. W. Hälbich, eds., Geody-
physics, v. 213, p. 57–96.
namics of the Cape fold belt: Geological Society of South
Jordan, T. E., and R. W. Allmendinger, 1986, The Sierras
Africa Special Publication 12, p. 165–175.
Pampeanas of Argentina: a modern analogue of Rocky
Hälbich, I. W., 1992, The Cape fold belt orogeny: state of the
Montain foreland deformation: American Journal of
art 1970s–1980s, in M. J. De Wit and I. G. D. Ransome, eds.,
Science, v. 286, p. 737–764.
Inversion tectonics of the Cape fold belt, Karoo and Creta-
Joubert, P., 1986, Namaqualand—a model of Proterozoic
ceous basins of southern Africa: Rotterdam, Balkema,
accretion?: Transactions of the Geological Society of South
p. 141–158.
Hall, J., J. Wright, and B. R. Hoffe, 1990, Deep seismic reflec- Africa, v. 89, p. 79–96.
tion profiling in frontier exploration: an example from Kaufman, A. J., S. B. Jacobsen, and A. H. Knoll, 1993, The
basins in Botswana with one billion years of subsidence, in Vendian record of Sr and C isotopic variations in seawater:
B. Pinet and C. Bois, eds, The potential of deep seismic implications for tectonics and paleoclimate: Earth and
profiling for hydrocarbon exploration: Paris, Éditions Planetary Science Letters, v. 120, p. 409–430.
Technip, p. 291–315. Kay, S. M., V. A. Ramos, C. Mpodozis, and P. Sruoga, 1989,
Harris, R. W., 1992, A structural analysis of the Hartbees River Late Paleozoic to Jurassic silicic magmatism at the
thrust belt, with special emphasis on the nature and origin Gondwana margin: analogy to the middle Proterozoic in
of the change in structural patterns across the boundary North America?: Geology, v. 17, p. 324–328.
48 Tankard et al.

Keeley, M. L., and M. P. R. Light, 1993, Basin evolution and Lombard, A. F., A. Gunzel, J. Inner, and L. Kruger, 1986, The
prospectivity of the Argentine continental margin: Journal Tsumeb lead-copper-zinc-silver deposits, South West
of Petroleum Geology, v. 16, p. 451–464. Africa/Namibia, in C. R. Anhaeusser and S. Maske, eds.,
Keidel, J., 1916, La geología de las sierras de la Provincia de Mineral deposits of southern Africa: Geological Society of
Buenos Aires y sus relaciones con las montanas de Sud South Africa, p. 1761–1787.
Africa y los Andes: Anales del Ministerio de Agricultura López, O., 1982, Reconocimiento y edad del Grupo Jacadigo
de la Nación, Sección Geología, Mineralogía y Minería, IX, (Paleozoico Inferior) en el Oriente Boliviano: V Congreso
v. 3, 1–78. Latinoamericano de Geología, Argentina, v. 1, p. 293–300.
Kingsley, C. S., 1977, Stratigraphy and sedimentology of the López-Gamundí, O. R., and E. A. Rossello, E. A., 1992, La
Ecca Group in the eastern Cape Province, South Africa: cuenca interserrana de Claromeco, Argentina: un ejemplo
Ph.D. dissertation, University of Port Elizabeth, South de cuenca de antepaís hercínica: VIII Congreso Geológico
Africa. Latinoamericano, Salamanca, Spain, Actas, p. 55–59.
Knüver, M., and M. Reissinger, 1982, The plutonic and meta- López-Gamundí, O. R., L. Alvarez, R. R. Andreis, I. Espejo, F.
morphic history of the Sierra de Ancasti (Catamarca Fernando-Seveso, D. A. Kokogian, L. Legarreta, C. O.
province, Argentina): Zentralblatt für Geologie und Limarino, and H. L. Sessarego, 1989, Cuencas intermon-
Palaeontologie, Teil I, 1981, p. 285–294. tanas neopaleozoicas: Décimo Congreso Geológico
Kokogian, D. A., F. Fernandez-Seveso, and A. Mosquera, Argentino, Simposio de Cuencas Sedimentarias,
1993, Las secuencias sedimentarias Triásicas: XII Congreso p. 123–167.
Geológico Argentino, II Congreso de Exploración de López-Gamundí, O. R, P. J. Conaghan, E. A. Rossello, and P.
Hidrocarburos, Relatorio, v. 1, p. 65–78. R. Cobbold, in press a, The Tunas Formation (Permian) in
Kröner, A., 1977, Precambrian mobile belts of southern and the Sierras Australes foldbelt, east central Argentina:
eastern Africa—ancient sutures or sites of ensialic evidence for syntectonic sedimentation in a Variscan
mobility?: Tectonophysics, v. 40, p. 101–135. foreland basin: Journal of South American Earth Sciences.
Kukla, P. A., and I. G. Stanistreet, 1991, Record of the López-Gamundí, O. R, I. S. Espejo, P. J. Conaghan, and C.
Damaran Khomas Hochland accretionary prism in central McA. Powell, in press b, Southern South America, in J. J.
Namibia: refutation of an “ensialic” origin of a late Protero- Veevers and C. McA. Powell, eds., Permian–Triassic basins
zoic orogenic belt: Geology, v. 19, p. 473–476. and foldbelts along the Panthalassan margin of Gond-
Kusznir, N. J., and S. S. Egan, 1989, Simple-shear and pure- wanaland: GSA Memoir 184.
shear models of extensional sedimentary basin formation: Lork, A., and H. Bahlburg, 1993, Precise U-Pb ages of
application to the Jeanne d’Arc basin, Grand Banks of monazites from the Faja Eruptiva de la Puna Oriental and
Newfoundland, in A. J. Tankard, and H. R. Balkwill, eds., the Cordillera Oriental, NW Argentina: XII Congreso
Extensional tectonics and stratigraphy of the North Geológico Argentino y II Congreso de Exploración de
Atlantic margins: AAPG Memoir 46, p. 305–322. Hidrocarburos Actas, v. 4, p. 1–6.
Lawrence, S. R., 1989, Prospects for petroleum in late Protero- Macedo, M. H., M. A. S. Basel, M. G. Bonhomme, and K.
zoic–early Paleozoic basins of southern-central Africa: Kawashita, 1984, Dados geocronológicos referente às
Journal of Petroleum Geology, v. 12, p. 231–242. rochas metassedimentares do Grupo Itajaí: Revista
Legarreta, L., C. A. Gulisano, and M. A. Uliana, 1993, Las Brasileira de Geociências, v. 14, p. 30–34.
secuencias sedimentarias jurasico-cretacicas, in V. A. Malumián, N., F. E. Nullo., and V. A. Ramos, 1983, The Creta-
Ramos, ed., Geología y recursos naturales de Mendoza: XII ceous of Argentina, Chile, Paraguay and Uruguay, in M.
Congreso Geológico Argentino y II Congreso de Explo- Moullade and A. E. M. Nairn, eds., The Phanerozoic
ración de Hidricarburos, Mendoza, Relatorio I, p. 87–114. geology of the world II, the Mesozoic B: Amsterdam,
Legarreta, L., D. A. Kokogian, and D. A. Dellapé, 1992, Estruc- Elsevier, p. 265–304.
turación terciaria de la cuenca cuyana: ¿Cuánto de Marshall, J. E. A., 1994, The Falkland Islands: a key element in
inversión tectónica?: Revista de la Asociación Geológica Gondwana paleogeography: Tectonics, v. 13, p. 499–514.
Argentina, v. 47, p. 83–86. Martinez, J. S., 1989, The structure of the Sierras Australes
Leighton, M. W., and D. R. Kolata, 1990, Selected interior (Buenos Aires Province, Argentina): an example of folding
cratonic basins and their place in the scheme of global in a transpressive environment: Journal of South American
tectonics: a synthesis, in M. Leighton, D. R. Kolata, D. F.
Earth Sciences, v. 2, p. 317–329.
Oltz, and J. J. Eidel, eds., Interior cratonic basins: AAPG
Mathalone, J. M. P., and M. Montoya, 1995, Petroleum
Memoir 51, p. 729–798.
geology of the sub-Andean basins of Peru, in A. J. Tankard,
Le Roux, J. P., 1983, Structural evolution of the Kango Group,
R. Suarez, and H. J. Welsink, Petroleum basins of South
in A. P. G. Söhnge and I. W, Hälbich, eds., Geodynamics of
America: AAPG Memoir 62, this volume.
the Cape fold belt: Geological Society of South Africa
McCourt, W. J., 1981, The geochemistry and petrography of
Special Publication 12, p. 47–56.
the coastal batholith of Peru, Lima segment: Journal of the
Litherland, M., and K. Bloomfield, 1981, The Proterozoic
Geological Society of London, v. 138, p. 407–420.
history of eastern Bolivia: Precambrian Research, v. 15, p.
McKirdy, D. M., A. K. Aldridge, and P. J. M. Ypma, 1983, A
157–179.
geochemical comparison of some crude oils from pre-
Litherland, M., R. N. Annells, D. P. F. Darbyshire, C. J. N.
Fletcher, M. P. Hawkins, B. A. Klinck, W. I. Mitchell, E. A. Ordovician carbonate rocks: Advances in Organic
O’Connor, P. E. J. Pitfield, G. Power, and B. C. Webb, 1989, Geochemistry 1981, p. 99–107.
The Proterozoic of eastern Bolivia and its relationship to McLachlan, I. R., and A. Anderson, 1973, A review of the
the Andean mobile belt: Precambrian Research, v. 43, p. evidence for marine conditions in southern Africa during
157–174. Dwyka times: Palaeontologica Africana, v. 15, p. 37–64.
Llambias, E. J., and A. M. Sato, 1990, El batolito de Colangüil McWilliams, M. O., and A. Kröner, 1981, Paleomagnetism and
(28–31°S), Cordillera frontal de Argentina; estructura y tectonic evolution of the Pan-African Damara belt,
marco tectónico: Revista geológica de Chile, v. 17, southern Africa: Journal of Geophysical Research, v. 86,
p. 89–108. p. 5147–5162.
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 49

Méndez, J., A. Navarini, D. Plaza, and V. Viera, 1972, Faja argentino: IX Congreso Geológico Argentino, Bariloche, v.
eruptiva de la Puna oriental: Actas Congreso Geológico 1, p. 384–398.
Argentina, Buenos Aires, v. 4, p. 83–100. Pezzi, E. E., and M. E. Mozetic, 1989, Cuencas sedimentarias
Milani, E. J., 1992, Intraplate tectonics and the evolution of the de la region Chacoparanense, in G. A. Chebli and L. A.
Paraná basin, SE Brazil, in M. J. De Wit and I. G. D. Spalleti, eds, Cuencas sedimentarias Argentinas: Tucumán,
Ransome, eds., Inversion tectonics of the Cape fold belt, Universidad Nacional de Tucumán, Serie correlación
Karoo and Cretaceous basins of southern Africa: geológica 6, p. 65–78.
Rotterdam, Balkema, p. 101–108. Pitts, B. E., M. J. Maher, J. H. de Beer, and D. I. Gough, 1992,
Miller, R. McG., 1980, Geology of a portion of central Damara- Interpretation of magnetic, gravity and magnetotelluric
land, South West Africa/Namibia: Geological Survey of data across the Cape fold belt and Karoo basin, in M. J. De
South Africa, South West Africa Series, Memoir 6, 78 p. Wit and I. G. D. Ransome, eds., Inversion tectonics of the
Miller, R. McG., 1983, The Pan-African Damara orogen of Cape fold belt, Karoo and Cretaceous basins of southern
South West Africa/Namibia, in R. McG. Miller, ed., Africa: Rotterdam, Balkema, 27–32.
Evolution of the Damara orogen of South West Ploszkiewicz, J. V., I. A. Orchuela, J. C. Vaillard, and R. F.
Africa/Namibia: Geological Society of South Africa Special Viñes, 1984, Compresión y desplazamiento lateral en la
Publication 11, 431–515. zona de falla Huincul, estructuras asociadas, provincia del
Miller, R. McG., and K. E. L. Schalk, 1980, South West Neuquen: IX Congreso Geológico Argentino, Actas, v. 2, p.
Africa/Namibia geologic map 1:1 000 000: Geological 163–169.
Survey of the Republic of South Africa and South West Polansky, J., 1970, Carbónico y Pérmico de la Argentina:
Africa/Namibia. Buenos Aires, Eudeba, 216 p.
Milner, S. C., A. R. Duncan, and A. Ewart, 1992, Quartz latite Porada, H., 1989, Pan-African rifting and orogenesis in
rheoignimbrite flows of the Etendeka Formation, north- southern and equatorial Africa and eastern Brazil: Precam-
western Namibia: Bulletin of Volcanology, v. 54, brian Research, v. 44, p. 103–136.
p. 200–219. Rabinowitz, P. D., and J. LaBrecque, 1979, The Mesozoic
Mon, R., and F. D. Hongn, 1991, The structure of the Precam- South Atlantic Ocean and evolution of its continental
brian and lower Paleozoic basement of the central Andes margins: Journal of Geophysical Research, v. 84,
between 22° and 32°S latitude: Geologische Rundschau, v. p. 5973–6002.
80, p. 745–758. Ramos, V. A., 1988, Late Proterozoic–early Paleozoic of South
Mon, R., and J. A. Salfity, 1995, Tectonic evolution of the America: a collisional history: Episodes, v. 11, p. 168–174.
Andes of northern Argentina, in A. J. Tankard, R. Suarez, Ramos, V. A., and J. M. Cortés, 1984, Estructura e inter-
and H. J. Welsink, Petroleum basins of South America: pretación tectónica: IX Congreso Geológico Argentino,
AAPG Memoir 62, this volume. Bariloche, Relatorio 1, p. 317–340.
Mpodozis, C., 1984, Dinámica de los margenes continentales Ransome, I. G. D., and M. J. De Wit, 1992, Preliminary investi-
activos: Seminario Actualización de la Geología de Chile, gations into a microplate model for the southwestern
Servicio Nacional de Geología y Minería, Santiago, Misce- Cape, in M. J. De Wit and I. G. D. Ransome, eds., Inversion
lanea no. 4, p. A-1/A-22. tectonics of the Cape fold belt, Karoo and Cretaceous
Mpodozis, C., and S. M. Kay, 1990, Provincias magmáticas basins of southern Africa: Rotterdam, Balkema, p. 257–266.
ácidas y evolución tectónica de Gondwana: Andes Rapela, C. W., and S. M. Kay, 1988, Late Paleozoic to Recent
Chilenos (28–31°S): Revista Geológica de Chile, v. 17, p. magmatic evolution of northern Patagonia: Episodes, v. 11,
153–180. p. 175–182.
Mpodozis, C., and V. Ramos, 1989, The Andes of Chile and Rapela, C. W., and R. J. Pankhurst, 1992, The granites of
Argentina, in G. E. Ericksen, M. T. Cañas Pinochet, and J. northern Patagonia and the Gastre fault system in relation
A. Reinemund, eds., Geology of the Andes and its relation to the break-up of Gondwana, in B. C. Story, T. Alabaster,
to hydrocarbon and mineral resources: Circum-Pacific and R. J. Pankhurst, eds., Magmatism and the causes of
Council for Energy and Mineral Resources, Earth Science continental break-up: Geological Society of London,
Series, v. 11, p. 59–90. Special Publication 68, p. 209–220.
Münch, H. G., 1974, The tectonics of the northern part of the Rapela, C. W., and R. J. Pankhurst, 1993, El volcanismo
Klein Karas and Groot Karas mountains, South West
riolítico del noreste de la Patagonia: un evento meso-
Africa: Geological Survey of South Africa Annals 9,
jurásico de corta duración y origen profundo: XII Congreso
p. 107–109.
Geológico Argentino, II Congreso de Exploración de
Muntingh, A., and L. F. Brown, 1994, Sequence stratigraphy of
Hidrocarburos, Actas, v. 4, p. 179–188.
petroleum plays, postrift Cretaceous rocks (lower Aptian
Rapalini, A. E., 1990, Variaciones paleolatitudinales de
to upper Maastrichtian), Orange basin, western offshore,
América del Sur en el Carbonífero-Pérmico: Décimo
South Africa, in P. Weimer and H. Posamentier, eds., Silici-
Primer Congreso Geológico Argentina, Actas, v. 2,
clastic sequence stratigraphy, recent development and
p. 259–262.
applications: AAPG Memoir 58, p. 71–98.
Renne, P. R., M. Ernesto, I. G. Pacca, R. S. Coe, J. M. Glen, M.
Nelson, R. A., T. L. Patton, and C. K. Morley, 1992,
Prévot, and M. Perrin, 1992, The age of Paraná flood
Rift–segment interaction and its relation to hydrocarbon
exploration in continental rift systems: AAPG Bulletin, volcanism, rifting of Gondwanaland, and the
v. 76, p. 1153–1169. Jurassic–Cretaceous boundary: Science, v. 258, p. 975–979.
Oliveira, L. O. A., 1987, Aspectos da evolução termomecânica Reyes, F. C., 1972, On the Carboniferous and Permian of
da Bacia do Paraná no Brasil: Tese de mestrado (inédito), Bolivia and northwestern Argentina: Anais da Academia
Universidade Federal de Ouro Preto, Escola de Minas, Brasileira de Ciências, v. 44 (suplemento), p. 261–277.
Departamento de Geologia, 179 p. Riccardi, A. C., and E. O. Rolleri, 1980, Cordillera patagonica
Omarini, R, and B. A. J. Baldis, 1984, Sedimentologia y mecan- austral: Segundo Simpsio de Geologia Regional Argentina,
ismos deposicionales de la Formación Puncoviscana Academia Nacional de Ciencias, Cordoba, p. 1173–1306.
(Grupo Lerma, Precámbrico–Cámbrico) del noroeste Rolleri, E. O., and P. Criado Roqué, 1968, La cuenca Triásica
50 Tankard et al.

del norte de Mendoza: Terceras Jornadas Geológicas Cape fold belt: Geological Society of South Africa Special
Argentinas, v. 1, p. 1–60. Publication 12, p. 1–6.
Rolleri, E. O., and C. A. F. Garrasino, 1979, Comarca Septentri- Spalletti, L. A., C. A. Cingolani, R. Varela, and A. J. Cuerda,
onal de Mendoza, in C. M. Turner, coord., Geología 1989, Sediment gravity flow deposits of an Ordovician
regional Argentina: Academia Nacional de Ciencias de deep-sea fan system (western Precordillera, Argentina):
Córdoba, v. 1, p. 771–809. Sedimentary Geology, v. 61, p. 287–301.
Rossello, E. A., and A. C. Massabie, 1992, Caracterización Stanistreet, I. G., P. A. Kukla, and G. Henry, 1991, Sedimen-
tectónica del kinking mesoscópico de las Sierras Australes tary basinal responses to a late Precambrian Wilson cycle:
de Buenos Aires: Revista de la Asociación Geológica the Damara orogen and Nama foreland, Namibia: Journal
Argentina v. 47, p. 179–187. of African Earth Sciences, v. 13, p. 141–156.
Rostirolla, S. P., F. F. de Alkmim, and P. C. Soares, 1992, O Stoakes, F. A., C. V. Campbell, R. Cass, and N. Ucha, 1991,
Grupo Itajai, Estado de Santa Catarina, Brasil: exemplo de Seismic stratigraphic analysis of the Punta del Este basin,
sedimentação em uma bacia flexural de Antepais: Boletim offshore Uruguay, South America: AAPG Bulletin, v. 75,
de Geociências da Petrobrás, v. 6, p. 109–244. p. 219–240.
Rust, I. C., 1967, On the sedimentation of the Table Mountain Storey, B. C., T. Alabaster, M. J. Hole, R. J. Pankhurst, and H.
Group in the western Cape Province: D.Sc. dissertation, E. Wever, 1992, Role of subduction–plate boundary forces
University of Stellenbosch, South Africa, 110 p. during the initial stages of Gondwana break-up: evidence
Rust, I. C., 1973, The evolution of the Paleozoic Cape basin, from the proto-Pacific margin of Antarctica, in B. C. Storey,
southern margin of Africa, in A. E. M Nairn and F. G. T. Alabaster, and R. J. Pankhurst, eds., Magmatism and the
Stehli, eds., The ocean basins and margins I, the South causes of continental break-up: Geological Society of
Atlantic: New York, Plenum, p. 247–276. London, Special Publication 68, p. 149–163.
Ryan, P. J., 1967, Stratigraphic and paleocurrent analysis of Suárez, R., 1986, El sistema Carbonífero de Bolivia. Un breve
the Ecca Series and lowermost Beaufort beds in the Karroo sumario: Trabajo presentado en el reunión final del
basin of South Africa: Ph.D. dissertation, University of Proyecto IGCP-211, Córdoba, Argentina, p. 1–5.
Witwatersrand, Johannesburg, South Africa. Tankard, A. J., and J. H. Barwis, 1982, Wave-dominated
Ryan, P. J., and G. G. Whitfield, 1979, Basinal analysis of the deltaic sedimentation in the Devonian Bokkeveld basin of
Ecca and lowermost Beaufort beds and associated coal, South Africa: Journal of Sedimentary Geology, v. 52,
uranium and heavy mineral beach sand occurrences, in A. p. 959–974.
M. Anderson and W. J. van Biljon, eds., Some sedimentary Tankard, A. J., M. P. A. Jackson, K. A. Eriksson, D. K. Hobday,
basins and associated ore deposits of South Africa: Geolog- D. R. Hunter, and W. E. L. Minter, 1982, Crustal evolution
ical Society of South Africa Special Publication 6, p. 91–101. of southern Africa; 3. 8 billion years of earth history: New
SACS (South African Committee for Stratigraphy), 1980, York, Springer-Verlag, 523 p.
Stratigraphy of South Africa, part I (Kent, L. E., compiler): Theron, J. N., and J. C. Loock, 1988, Devonian deltas of the
lithostratigraphy of the Republic of South Africa, South Cape Supergroup, South Africa, in N. J. McMillan, A. F.
West Africa/Namibia and the Republics of Embry, and D. J. Glass, eds., Devonian of the world, vol. I:
Bophuthatswana, Transkei, and Venda: Handbook of the regional syntheses: Canadian Society of Petroleum Geolo-
Geological Survey of South Africa 8, 690 p. gists Memoir 14, p. 729–740.
Salfity, J. A., 1985, Lineamientos transversales al rumbo Thomas, R. J., M. W. von Veh, and S. McCourt, 1993, The
andino en el noroeste Argentino: IV Congreso Geológico tectonic evolution of southern Africa: an overview: Journal
Chileno, v. 2, p. 119–137. of African Earth Sciences, v. 16, p. 5–24.
Salfity, J. A., C. R. Monaldi, R. A. Marquillas, and R. E. Turner, B. R., 1983, Braidplain deposition of the Upper
González, 1993, La inversión tectónica del umbral de Los Triassic Molteno Formation in the main Karoo
Gallos en la cuenca del Grupo Salta durante la fase Incaica: (Gondwana) basin, South Africa: Sedimentology, v. 30,
XII Congreso Geológico Argentino y II Congreso de Explo- p. 77–89.
ración de Hidrocarburos, Actas 3, p. 200–210. Turner, S., M. Regelous, S. Kelley, C. J. Hawkesworth, and M.
Santa Ana, H. de, 1989, Tectonic and depositional considera- Mantovani, 1994, Magmatism and continental break-up in
tions of the northern basin of Uruguay: exploración, estado the South Atlantic: high precision 40Ar-39Ar
actuel y posibles tendencias de su tecnología: Reunión de geochronology: Earth and Planetary Science Letters, v. 121,
Expertos de ARPEL LXXI, Calgary, Canada, 14 p. p. 333–348.
Sawyer, E. W., 1981, Damaran structural and metamorphic Uliana, M. A., and K. T. Biddle, 1988, Mesozoic–Cenozoic
geology of an area south-east of Walvis Bay, South West paleogeographic and geodynamic evolution of southern
Africa/Namibia: Memoirs of the Geological Survey of South America: Revista Brasileira de Geociências, v. 18,
South West Africa 7, p. 1–83. p. 172–190.
Schobbenhaus, C, D. de A. Campos, G. R. Derze, and H. E. Uliana, M. A., K. T. Biddle, and J. Cerdan, 1989, Mesozoic
Asmus, 1981, Mapa geológico do Brasil, e da área oceânica extension and the formation of Argentine sedimentary
adjacente incluindo depósitos minerais, 1:2 500 000. Brasil: basins, in A. J. Tankard and H. R. Balkwill, eds., Exten-
Ministério das Minas e Energia. sional tectonics and stratigraphy of North Atlantic
Sempere, T., 1995, Phanerozoic evolution of Bolivia and margins: AAPG Memoir 46, p. 599–614.
adjacent regions, in A. J. Tankard, R. Suarez, and H. J. Uliana, M. A., and L. Legarreta, 1993, Hydrocarbons habitat in
Welsink, Petroleum basins of South America: AAPG a Triassic to Cretaceous sub-Andean setting: Neuquén
Memoir 62, this volume. basin, Argentina: Journal of Petroleum Geology, v. 16,
Snyder, D. B., V. A. Ramos, and R. W. Allmendinger, 1990, p. 397–420.
Thick-skinned deformation observed on deep seismic Unrug, R., 1992, The supercontinent cycle and Gondwanaland
reflection profiles in western Argentina: Tectonics, v. 9, assembly: component cratons and the timing of suturing
p. 773–788. events: Journal of Geodynamics, v. 16, p. 215–240.
Söhnge, A. P. G., 1983, The Cape fold belt—perspective, in A. Urien, C. M., L. R. Martins, and J. J. Zambrano, 1976, The
P. G. Söhnge and I. W. Hälbich, eds., Geodynamics of the geology and tectonic framework of southern Brazil,
Structural and Tectonic Controls of Basin Evolution in Southwestern Gondwana 51

Uruguay and northern Argentina continental margin, their Wiens, F., 1995, Phanerozoic tectonics and sedimentation in
behavior during the Southern Atlantic opening: Academia the Chaco Basin of Paraguay, with comments on hydro-
Brasileira Ciencias, Anais, v. 48, p. 365–376. carbon potential, in A. J. Tankard, R. Suarez, and H. J.
Utting, J., 1978, Lower Karroo pollen and spore assemblages Welsink, Petroleum basins of South America: AAPG
from the coal measures and underlying sediments of the Memoir 62, this volume.
Siankondobo coalfield, mid-Zambezi valley, Zambia: Paly- Williams, G. D., C. M. Powell, and M. A. Cooper, 1989,
nology, v. 2, p. 53–68. Geometry and kinematics of inversion tectonics, in M. A.
Van Vuuren, C. J., and D. I. Cole, 1979, The stratigraphy and Cooper and G. D. Williams, eds., Inversion tectonics:
depositional environments of the Ecca Group in the Geological Society of London, Special Publication 44,
northern part of the Karoo basin, in A. M. Anderson and p. 3–15.
W. J. van Biljon, eds., Some sedimentary basins and associ- Willner, A. P., U. S. Lottner, and H. Miller, 1987, Early
ated ore deposits of South Africa: Geological Society of Paleozoic structural development in the NW Argentine
South Africa Special Publication 6, p. 103–111. basement of the Andes and its implication for geodynamic
Vargas, C., and J. M. Suárez, 1980, Anomalias tipo canal detec- reconstructions, in G. D. McKenzie, ed., Gondwana six:
tadas por métodos sísmicos en el Carbonífero de Bolivia: structure, tectonics, and geophysics: American Geophys-
XXXIX Reunión de Expertos de ARPEL, Bolivia, p. 1–28. ical Union, Geophysical Monograph 40, p. 229–239.
Vergani, G. D., A. J. Tankard, H. J. Belotti, and H. J. Welsink, Zalán, P. V., S. Wolff, J. C. J. Conceiçao, A. Marques, M. A. M.
1995, Tectonic evolution and paleogeography of the Astolfi, I. S. Vieira, V. T. Appi, and O. A. Zanotto, 1991,
Neuquén basin, Argentina, in A. J. Tankard, R. Suarez, and Bacia do Parana, in G. P. de Raja Gabaglia and E. J. Milani,
H. J. Welsink, Petroleum basins of South America: AAPG eds., Origem e evolução de Bacias Sedimentares: Rio de
Memoir 62, this volume. Janeiro, Petrobras, p. 135–168.
Verrall, P., 1989, Speculations on the Mesozoic–Cenozoic Ziegler, P. A. 1993, Plate-moving mechanisms: their relative
tectonic history of the western United States, in A. J. importance: Journal of the Geological Society of London, v.
Tankard and H. R. Balkwill, eds., Extensional tectonics and 150, p. 927–940.
stratigraphy of the North Atlantic margins: AAPG Memoir Zoback, M. L., and M. Zoback, 1980, State of stress in the
46, p. 615–631. conterminous United States: Journal of Geophysical
Visser, D. J. L., 1984, Geological map of the Republics of South Research, v. B85, p. 6113–6156.
Africa, Transkei, Bophuthatswana, Venda and Ciskei and
the kingdoms of Lesotho and Swaziland: Geological
Survey of the Republic of South Africa, scale 1:1,000,000.
Visser, J. N. J., 1987a, The palaeogeography of part of south-
western Gondwana during the Permo-Carboniferous Authors’ Mailing Addresses:
glaciation: Palaeogeography, Palaeoclimatology, Palaeoe-
cology v. 61, p. 205–219. A. J. Tankard
Visser, J. N. J., 1987b, The influence of topography on the Tankard Enterprises Ltd.
Permo-Carboniferous glaciation in the Karoo basin and P. O. Box 81002
adjoining areas, southern Africa, in G. D. McKenzie, ed., Calgary, Alberta T2J 7C9
Gondwana six: stratigraphy, sedimentology, and paleon- Canada
tology: American Geophysical Union, Geophysical
Monograph 41, p. 123–129.
Von Veh, M. W., 1990, Diamictites in the Gariep belt:
M. A. Uliana
Extended abstracts, Geocongress 90, Geological Society of ASTRA C.A.P.S.A.
South Africa, p. 586–589. Tucumán 744
Vos, R. G., and A. J. Tankard, 1981, Braided fluvial sedimenta- 1049 Buenos Aires
tion in the lower Paleozoic Cape basin, South Africa: Sedi- Argentina
mentary Geology, v. 29, p. 171–193.
Welsink, H. J., E. Martinez, O. Aranibar, and J. Jarandilla, H. J. Welsink
1995, Structural inversion of a Cretaceous rift basin, Perez Companc
southern Altiplano, Bolivia, in A. J. Tankard, R. Suarez, J. J. Lastra Sud 6000
and H. J. Welsink, Petroleum basins of South America: 8300 Neuquén
AAPG Memoir 62, this volume.
Wernicke, B., 1985, Uniform-sense normal simple shear of the
Argentina
continental lithosphere: Canadian Journal of Earth
Sciences, v. 22, p. 108–125. V. A. Ramos
Wickens, H. de V., 1992, Submarine fans of the Permian Ecca Dpto. Ciencias Geológicas
Group in the SW Karoo basin: their origin and reflection Universidad de Buenos Aires
on the tectonic evolution of the basin and its source areas, 1428 Nuñez
in M. J. De Wit and I. G. D. Ransome, eds., Inversion Argentina
tectonics of the Cape fold belt, Karoo and Cretaceous
basins of southern Africa: Rotterdam, Balkema, p. 117–125. M. Turic
Wickens, H. de V., G. J. Brink, W. Van Rooyen, A. H. Bouma, YPF
and L. F. Brown, 1992, The Tanqua turbidite and deltaic
complexes: depositional models, reservoir realities, and the
Avenida R. Saenz Peña 777
application of sequence stratigraphy: AAPG Field Seminar 1364 Buenos Aires
Guidebook, 180 p. Argentina
52 Tankard et al.

A. B. França M. Cirbián
E. J. Milani O. López Paulsen
Petrobras/Nexpar YPFB
Rua Padre Camargo 285 Casilla 1659
Curitiba, PR Santa Cruz
Brazil Bolivia

B. B. de Brito Neves G. J. B. Germs


Universidade de S. Paulo JCI Research Unit
01498 970 Sao Paulo–SP P.O. Box 976
Brazil Randfontein 1760
South Africa
N. Eyles
Department of Geology M. J. De Wit
University of Toronto Department of Geological Sciences
Scarborough, Ontario M1C 1A4 University of Cape Town
Canada Rondebosch 7700
J. Skarmeta South Africa
ENAP
Casilla 3556 T. Machacha
Santiago Geological Survey Department
Chile Private Bag 14
Lobatse
H. Santa Ana Botswana
ANCAP
Casilla 1090 R. McG. Miller
Montevideo NAMCOR
R. O. del Uruguay Private Bag 13196
Windhoek
F. Wiens Namibia
Geo Consultores
Casilla Postal 166
Asunción
Paraguay

You might also like