Sempere2002 PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Tectonophysics 345 (2002) 153 – 181

www.elsevier.com/locate/tecto

Late Permian–Middle Jurassic lithospheric thinning in Peru and


Bolivia, and its bearing on Andean-age tectonics
Thierry Sempere a,*, Gabriel Carlier b, Pierre Soler c, Michel Fornari d, Vı́ctor Carlotto e,
Javier Jacay f, Oscar Arispe g, Didier Néraudeau h, José Cárdenas e, Silvia Rosas i,
Néstor Jiménez j
a
IRD, Apartado Postal 18-1209, Lima 18, Peru
b
IRD and Laboratoire de Minéralogie, Muséum National d’Histoire Naturelle, 61 rue Buffon, 75005 Paris, France
c
IRD and LODYC-UMR 7617 CNRS-IRD-UPMC, Institut Pierre-Simon Laplace, Université Pierre-et-Marie-Curie, Boı̂te 100, 4 Place Jussieu,
75252 Paris cedex 05, France
d
IRD and Laboratoire de Géochronologie, UMR 6526 Géosciences Azur, Université de Nice-Sophia Antipolis, 06108 Nice cedex 02, France
e
Departamento de Geologı́a, Universidad Nacional de San Antonio Abad del Cusco, Cusco, Peru
f
Universidad Nacional Mayor de San Marcos, Apartado Postal 3973, Lima 100, Peru
g
Casilla 4836, La Paz, Bolivia
h
Laboratoire de Paléontologie, Géosciences, Université de Rennes I, Campus de Beaulieu, Avenue du Général Leclerc,
35042 Rennes cedex, France
i
Sociedad Geológica del Perú, Arnaldo Márquez 2277, Lima 11, Peru
j
Universidad Mayor de San Andrés, Casilla 6568, La Paz, Bolivia
Received 30 March 2000; received in revised form 10 December 2000; accepted 15 December 2000

Abstract

Integrated studies and revisions of sedimentary basins and associated magmatism in Peru and Bolivia (8 – 22S) show that
this part of western Gondwana underwent rifting during the Late Permian – Middle Jurassic interval. Rifting started in central
Peru in the Late Permian and propagated southwards into Bolivia until the Liassic/Dogger, along an axis that coincides with the
present Eastern Cordillera. Southwest of this region, lithospheric thinning developed in the Early Jurassic and culminated in the
Middle Jurassic, producing considerable subsidence in the Arequipa basin of southern Peru. This  110-Ma-long interval of
lithospheric thinning ended  160 Ma with the onset of Malm – earliest Cretaceous partial rift inversion in the Eastern
Cordillera area. The lithospheric heterogeneities inherited from these processes are likely to have largely influenced the
distribution and features of younger compressional and/or transpressional deformations. In particular, the Altiplano plateau
corresponds to a paleotectonic domain of ‘‘normal’’ lithospheric thickness that was bounded by two elongated areas underlain
by thinned lithosphere. The high Eastern Cordillera of Peru and Bolivia results from Late Oligocene – Neogene intense inversion
of the easternmost thinned area. D 2002 Elsevier Science B.V. All rights reserved.

Keywords: Lithospheric thinning; Andean orogeny; Mesozoic; Cenozoic; Peru; Bolivia

*
Corresponding author.
E-mail addresses: sempere@terra.com.pe (T. Sempere), gabi@cimrs1.mnhn.fr (G. Carlier), Pierre.Soler@lodyc.jussieu.fr (P. Soler),
fornari@taloa.unice.fr (M. Fornari), carlotto@chaski.unsaac.edu.pe (V. Carlotto), oarispe@hotmail.com (O. Arispe),
Didier.Neraudeau@univ-rennes1.fr (D. Néraudeau), roque@chaski.unsaac.edu.pe (J. Cárdenas).

0040-1951/02/$ - see front matter D 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 0 4 0 - 1 9 5 1 ( 0 1 ) 0 0 2 11 - 6
154 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

1. Introduction

The Andes Cordillera is classically considered as a


prime example of a mountain belt built along a
continental margin by tectonic processes related to
subduction of an oceanic plate, in an apparently non-
collisional setting. This outstanding orogenic belt
shows considerable longitudinal variations (Fig. 1).
The Central Andes form the largest and most
mountainous segment of the Andes Cordillera. This
 4000-km-long segment is itself segmented into the
northern Central Andes (5300 –  13S; entirely lo-
cated in Peru), Bolivian Orocline (  13– 28S; over
southern Peru, Bolivia, northern Chile, northwestern
Argentina), and southern Central Andes (28 – 37S;
over central Chile and west – central Argentina). The
mountainous area of the largest of these segments, the
Bolivian Orocline, covers  1,300,000 km2 (we use
here the term ‘orocline’ only in a descriptive sense,
which refers to the characteristic oceanward concavity
of this segment). This paper is concerned with the
northern Central Andes, and with the Peruvian and
Bolivian parts of the Bolivian Orocline.
The geological structure of the Central Andes is
relatively poorly known. Many tectonicists and geo-
Fig. 1. Segmentation of the Andes Cordillera as used in this paper.
physicists have implicitly tended to consider most of Topography from USGS data.
the orogen as somewhat homogeneous (e.g., Isacks,
1988; Lamb et al., 1997). Characteristically enough,
descriptions of the Central Andean tectonic structure structure and history of the Central Andes, a precise
are generally based on the present-day geomorphic knowledge of the pre-orogenic history and structure
features. A consensus exists, however, about the idea is obviously needed, because ancient structures are
that the Central Andean active continental margin has likely to have played significant tectonic roles during
undergone significant crustal thickening since the Late Andean shortening.
Cretaceous (e.g., Jaillard and Soler, 1996). This paper provides a detailed summary of new and
Little is known of the causes and processes that updated data, and of often entrenched information, and
result in the crustal thickening of a continental demonstrates that the study area was submitted to
margin submitted to an ocean – continent conver- lithospheric thinning during the Late Permian– Middle
gence. The Central Andes happen to provide geo- Jurassic. We observe that a Late Permian –Liassic rift
scientists with an extreme case of convergence- system developed along what is today the high Eastern
related continental thickening, and are thus worthy Cordillera of Peru and Bolivia. We infer that distribu-
of detailed studies. In order to constrain the actual tion of pre-orogenic thinned areas coincide with spe-

Fig. 2. Mesozoic – Cenozoic stratigraphic synopsis of the Central Andes from  10S to  20S. Note overall decrease in marine influence
from W to E, and N to S. Abbreviations: G = Gramadal Fm, Hé = Huancané Fm, Hh = Hualhuani Fm, Ipa = Ipaguazú Fm, L = Labra Fm,
M = Muni Fm, Mc = Murco Fm, P + C = Puente and Cachı́os Fms, S = Sipı́n Fm, Sc = Saracocha conglomerates. Framed numbers: 1 = Late
Permian – Triassic rifting (diachronous from N to S); 2 = downwarping of Arequipa basin; 3 = downwarping of Chicama basin; 4 = latest
Jurassic – Early Cretaceous uplifts (4 * = Lagunillas uplift). References for northern Chile are mainly from Ramı́rez and Huete (1981), Skarmeta
and Marinovic (1981), Bogdanic (1990), Ladino et al. (1999); for other areas, see text and Figs. 6 and 8.
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 155
156 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

cific Andean structural traits, and we explore the idea In Peru, strata of Permian through Jurassic age are
that pre-existent extensional and/or transtensional divided into the Mitu and Pucará groups, which were
structures determined many actual characteristics of respectively deposited in continental and marine set-
the northern Central Andes and Bolivian Orocline. tings (McLaughlin, 1924; Steinmann, 1929; Harrison,
The Mesozoic – Cenozoic stratigraphic evolution 1943, 1951; Jenks, 1951; Newell et al., 1953). The
concerning the study area is summarized in three clastic and volcanic Mitu Group is mostly known
stratigraphic transects (Fig. 2). For correspondence from the Eastern Cordillera of central and southern
between chronostratigraphic stages and absolute ages, Peru and accumulated in subsident grabens, reflecting
we use Hardenbol et al.’s (1998) chart. Our use of the Late Permian – Triassic rifting (e.g., Mégard, 1978;
term ‘Andean’ refers to the Andes (or proto-Andes) as Dalmayrac et al., 1980; Kontak et al., 1985). North-
a mountain belt, and to the related period of mountain west of Cusco, the carbonate-dominated Late Trias-
building. sic –Liassic Pucará Group has a wider distribution
(Fig. 3) and was deposited during the thermal sag that
followed rifting in this region; more to the southeast,
2. Late Permian – Middle Jurassic rifting in the thick fluvio-eolian sandstones were deposited during
Eastern Cordillera of Peru and Bolivia this thermal sag period (Sempere et al., 1998, 1999,
2000a).
2.1. Introduction In the grabens produced by rifting, conformable or
deformed Late Paleozoic strata were generally pre-
Late Permian– Triassic rifting diachroneously de- served below the Mitu Group, whereas they were
veloped in the Eastern Cordillera of Peru (Mégard, eroded out from the neighbouring rift shoulders.
1978; Laubacher, 1978; Noble et al., 1978; Dalmayrac Intense magmatism commonly occurred at depth
et al., 1980; Kontak et al., 1985; Rosas and Fontboté, beneath the floor of the grabens, and predominantly
1995; Rosas et al., 1997; Jacay et al., 1999) and alkaline volcanic rocks were erupted. Consistent iso-
extended into Bolivia in the Late Triassic –Middle topic ages obtained on volcanic and plutonic rocks
Jurassic (McBride et al., 1983; Sempere, 1995; Sem- point to a Late Permian –Middle Jurassic age for the
pere et al., 1998, 1999). The main axis of the rift rifting (see Kontak et al., 1985, 1990; Soler, 1991;
system appears to coincide with the axis of the Eastern Jacay et al., 1999, for summaries).
Cordillera in both countries (Fig. 3). In this paper we
dedicate special attention to data from the still little 2.2. Pre-rift strata and deformation
known Bolivian continuation of this rift system.
Current research in southern Peru and Bolivia shows 2.2.1. Pre-rift stratigraphic units and relationships
that Mesozoic Bolivian basins were mostly connected The commonly >500-m-thick Pennsylvanian –
to Peruvian basins, and not to southern, Argentine – Early Permian (Tarma-)Copacabana Group was
Chilean basins. deposited prior to rifting, and forms a guide unit as
Reconstruction of the rift system in map view shows it was frequently preserved in the Mitu grabens. It is
that it splits into two branches at about 19S (Figs. 3 of shallow-marine origin and consists of fossiliferous
and 4). The southeastern, ‘‘Entre Rı́os branch’’ extends limestones and subordinate sandstones, black shales,
into the Chaco Subandean belt and dies out in the and cherty limestones. To the southeast, in the Chaco
Bolivia– Argentina border area. The southern, ‘‘Tupiza Subandean belt and lowlands of Bolivia, the Copaca-
branch’’ strikes (presently) N10E and apparently bana carbonate platform grades into sandstones of
extends into the Argentine Puna. In map view, this eolian and fluvial origin (Cangapi Formation; Sem-
geometry is reminiscent of the present-day Red Sea rift pere, 1995). In some areas of the Eastern Cordillera
system, which to the north splits into the now inactive where the Copacabana Group was preserved in Mitu
Suez Gulf and the active Aqaba Gulf rift and Dead Sea grabens, a rapid transition with overlying black shales
wrench – fault system (Fig. 5). In the following, we is observed. In other localities of the Eastern Cordil-
consider the ‘‘Tupiza branch’’ as the southern contin- lera, strata of the Copacabana Group are folded
uation of the main rift axis. (locally intensely, and metamorphized) and either
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 157

Fig. 3. Synopsis of the main Mesozoic geologic elements of Peru and Bolivia. The axis of the Late Permian – Middle Jurassic rift system is
defined by occurrences of the Mitu Group, coeval granitoids, and basic dyke swarms, and approximately coincides with the axis of the Eastern
Cordillera of Peru and Bolivia. Localities: A: Arequipa, C: Cochabamba, Cu: Cusco, L: Lima, P: Potosı́, SC: Santa Cruz, Tu: Tupiza.
158 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

Fig. 4. Location map of relevant data concerning the northern part of the Bolivian orocline.

intruded by Mitu granitoids (e.g., Soler and Bon- This contrasted variety of stratigraphic relationships
homme, 1987) or unconformably overlain by the Mitu (Fig. 6) provides information on the tectonic frame-
Group (e.g., Laubacher, 1978; Mégard et al., 1983). work of early rifting (see below).
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 159

overall shallowing-upward succession. Marine black


shales are characteristic of the lower Vitiacua Forma-
tion, whereas restricted-marine chert-rich carbonates
are especially common in the upper part of the unit
(Sempere et al., 1992). The Vitiacua Formation overlies
the fluvio-eolian sandstones of the Cangapi Formation
with a rapid transition (marking a transgression), and is
sharply overlain by the mudstone-dominated, gypsum-
bearing, continental red strata of the Ipaguazú Forma-
tion.
Palynomorphs and a coelacanth jaw bone from the
organic-rich lower part of the Vitiacua Formation
indicate a Late Permian age (Sempere et al., 1992),
whereas fossil molluscs and fish from its uppermost
part suggest that it reaches the Triassic, and possibly the
Late Triassic (Beltan et al., 1987; Suárez-Riglos and
Dalenz, 1993), although an Early or Middle Triassic
age seems more likely (Sempere et al., 1992, 1998).
Conchostracans suggestive of a Late Permian age are
recorded from the Vitiacua Formation (Palaeolimnadia
sp., Palaeolimnadiopsis cf. eichwaldi) and from the
overlying Ipaguazú Formation (Cyzicus [Lioestheria]
sp.) (Tasch, 1987); however, ages indicated by South
American conchostracans are not securely calibrated.
The stratigraphy of Bolivia shares many aspects
with the stratigraphy of Brazil (Sempere, 1995). The
Late Permian Irati Formation of the Paraná basin, Bra-
zil, is quite similar to the organic-rich lower part of the
Vitiacua Formation and represents its eastern equiva-
lent (Sempere et al., 1992). The Irati Formation is
conformably overlain by units (Serra Alta, Teresina,
and Rio do Rasto formations in the south; Corumbataı́
Formation in the north) that correlate with overlying
members of the Vitiacua Formation and range into the
Triassic (Francßa et al., 1995); continental red strata
known as the Pirambóia Formation unconformably
Fig. 5. Comparison between the map view geometries of the Red overlie this Late Permian – Early Triassic stratigraphic
Sea rift system and of the Central Andean Late Permian – Middle
Jurassic rift system (now inverted and shortened). Branching of the
set, this contact lying somewhere in the Middle Triassic
rift axis is observed in both cases. (Francßa et al., 1995). As the continental red strata of the
Ipaguazú Formation sharply overlie the Vitiacua For-
mation, and as Bolivia and Brazil both belonged to the
2.2.2. Restricted-marine deposits coeval of early cratonic area of western Gondwana at that time, a
rifting: the Vitiacua and Chutani formations, Bolivia, Middle Triassic age for the Vitiacua/Ipaguazú contact
and Ene Formation, Peru is suggested. The sedimentary discontinuity at the base
The Vitiacua Formation of southern Bolivia consists of the Ipaguazú and Pirambóia formations is inter-
of black shales, siliceous carbonates (mainly lime- preted to mark a large-scale destabilization of conti-
stones and dolomites with common chert), dark red nental depositional systems and, given its age, to
mudstones, and subordinate sandstones, which form an distally reflect onset of rifting to the west.
160 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

Fig. 6. Synopsis of Permian through Jurassic main stratigraphic, deformational, and magmatic data discussed in text concerning the Eastern
Cordillera of Peru and Bolivia, the Chaco basin of Bolivia, and the Paraná basin of Brazil.

In the Eastern Cordillera of Bolivia, the Vitiacua numerous Botryococcus) indicate marine to restricted-
Formation is found only at three localities, which are marine environments, and Middle to Late Permian
located within the main rift axis: Pasto Waykho, 6 km ages (J. Doubinger, unpublished).
west of Vitichi; Quebrada Aymaraj Hueko, 6 km About 70 km southwest of the main rift axis, the
southwest of Torotoro; Iglesiani, 40 km NNW of Chutani Formation crops out in the Tiquina area of
Cochabamba (Fig. 4). It is likely that the Vitiacua southeastern Lake Titicaca, where it overlies the
Formation was preserved there due to tectonic down- Copacabana Group (Oviedo, 1962, 1964). The Chu-
warping during rifting. At all three localities the tani Formation displays facies that are similar to those
Vitiacua Formation is overlain by altered basalt flows of the Vitiacua Formation, including dark shales in its
and/or conglomerates with  15-cm-size basalt clasts lower part, and should thus be of similar Late Per-
that mark the base of the Serere Group. At the first mian – Early? Triassic age (Sempere et al., 1998).
two localities, the chert-rich facies and chert breccias Fossil plants of Late Permian age occur in the upper
that characterize the upper part of the Vitiacua For- Chutani Formation (Iannuzzi et al., 1997; Vieira et al.,
mation are observed at the top of the local succession, 1999a,b). Altered basalt flows are intercalated with
suggesting that the Vitiacua Formation is stratigraph- coastal plain deposits and stromatolitic beds at the top
ically complete; furthermore, the contact between the of this unit. These flows are conformably overlain by
Vitiacua Formation and overlying red strata is nearly the brown-red fluvial sandstones and mudstones of the
transitional there, and it seems obvious that no large Tiquina Formation, which correlates with the Ipa-
time gap separated deposition of the two units. At guazú Formation.
Iglesiani, two rich palynomorph assemblages (1: In Peru, the Ene Formation displays facies similar
Hamiapollenites karrooensis, Tornopollenites toreu- to those in the Vitiacua and Chutani formations. In
tis, Lueckisporites virkkiae, Corisaccites alutas, Pro- particular, the lower part of the Ene Formation pre-
tohaploxypinus enigmaticus, Taeniaesporites sp. [sp. dominantly consists of organic-rich black shales of
1 Jardiné, 1974], Paravittatina cincinnata, Punctatis- Late Permian age (e.g., Mathalone and Montoya,
porites gretensis, . . ., and numerous acritarchs includ- 1995; Carlotto et al., 2000). At two localities near
ing several species of Micrhystridium; and 2: the Mitu rift axis (Fig. 4), these black shales con-
Lueckisporites virkkiae, Corisaccites alutas, Guttula- formably overlie the Copacabana Group and grade
pollenites gondwanensis, Gondwanipollenites ovatus, upward into siliceous carbonates and/or shallow-
Schweitzerisporites maculatus, Weylandites sp., with marine to fluvial or eolian sandstones; this continuous
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 161

succession is in turn conformably overlain by altered Cordillera (Sempere, 1995); given the contemporane-
volcanic rocks and red strata (including mudstones, ity of Late Permian deformation and deposition, this
sandstones and gypsum) that represent the local facies belt was probably discontinuous, deformation occur-
of the Mitu Group (Carlotto et al., 2000). The Ene ring in specific areas while shale-dominated sedimen-
Formation is widespread in the Subandean belt and tation was quietly going on in other areas of the same
lowlands of Peru, i.e. east of the Mitu rift system Eastern Cordillera domain.
(Mathalone and Montoya, 1995). Such relationships are strongly suggestive of a
The black shales that characterize the lower parts transcurrent rift system in which transtensional seg-
of the Vitiacua, Chutani and Ene formations represent ments would have been separated by transpressional
a marine transgression of Late Permian age, which is ‘‘nodes’’. We favor the idea that local transpression
also represented by units in the Paraná basin, Brazil, caused deformation of pre-Mitu strata at the onset of
and the Karoo basin, South Africa (Sempere et al., continental dislocation, before general graben forma-
1992; Tankard et al., 1995). This widespread trans- tion and intense magmatism developed. Coeval trans-
gression is likely to have extended over a very broad tension produced slow downwarping of elongated
region of western Gondwana, but was also coeval areas, where the Copacabana Group was preserved
with the early stage of the rift-related Mitu magma- and deeper marine shales conformably deposited over
tism (Fig. 6). In southern Peru, limestones bearing it, before accelerating rifting processes enhanced mag-
Late Permian fusulinids are locally intercalated within matism and formed true grabens.
red strata of the Mitu Group (Laubacher, 1978), de- A similar scenario, albeit later in time, could also
monstrating that marine ingressions occurred within explain the occurrence of Late Triassic plutons show-
some Mitu grabens. ing deformation that was contemporaneous with their
The Vitiacua, Chutani and Ene formations con- emplacement. In the Cordillera Real of western Boli-
formably overlie the Copacabana Group, and, given via, the Zongo –Yani pluton yielded Late Triassic U –
their age, appear as restricted-marine time-equivalents Pb ages (Table 1); emplacement of this foliated,
of the older Mitu syn-rift deposits, whereas they pre- peraluminous, two-mica granite was contemporane-
date younger Mitu syn-rift deposits as they are over- ous with schistosity and low-pressure metamorphism,
lain by Mitu or Mitu-equivalent red strata (Fig. 6). reflecting a high heat flow (Bard et al., 1974). In
Given the approximate Middle Triassic age of the top nearby Peru, the similar, foliated and peraluminous,
of these restricted-marine units, the Mitu rifting ap- two-mica Limacpampa pluton is dated to near the
pears to have been twofold, with a first stage spanning Triassic –Jurassic boundary (Table 1). South of Aban-
the Late Permian – Early? Triassic interval, and a cay, a cataclastic ‘‘quartz-diorite’’ yielded a Late Tri-
second stage beginning in the Middle Triassic. This assic U –Pb age (Table 1). We suggest that emplacement
analysis agrees with the idea of a twofold Mitu Group and early deformation of these intrusions might have
set forth by Soler (1991) on independent bases (see occurred in local transpressional settings at a later, Late
below). Triassic, stage of rifting.
Triassic uplift of plutons is recorded by clasts of
2.2.3. Deformation and rifting Mitu-age granitoids that are commonly found in
As stressed above, the sedimentary continuity conglomerates and pyroclastites of the Mitu Group
commonly observed in Bolivia and southern Peru in central Peru (Mégard, 1978), suggesting a twofold
between the Pennsylvanian – Early Permian Copaca- development of rifting (Soler, 1991). Such uplifts are
bana Group and overlying Late Permian – Early? Tri- likely to have been caused by lithospheric deforma-
assic units markedly contrasts with the pre-Mitu, tion related to rifting.
locally intense deformation observed in the Copaca-
bana Group in some areas of the Eastern Cordillera 2.3. Rift-related magmatism
between  11S and  17300S. This deformation,
traditionally explained by Late Permian ‘‘tardi-hercy- 2.3.1. Southern Peru
nian tectonics’’ (e.g., Dalmayrac et al., 1980), is in Intense magmatism was associated with the Mitu
fact restricted to a narrow belt within the Eastern rifting in southern Peru (Egeler and De Booy, 1961;
162
Table 1
Isotopic ages mentioned in text. See original references for details; bi: biotite, mu: muscovite, pl: plagioclase, hb: hornblende, wr: whole rock, gl: glass

T. Sempere et al. / Tectonophysics 345 (2002) 153–181


Region Unit Method Age (Ma) Observations Source
Central Peru: Equiscocha granite K – Ar 253 ± 11 (mu), 204 ± 9 (pl) Soler, 1991
Eastern Cordillera San Ramón-type batholiths Rb – Sr, wr 246 ± 10 recalculated by Soler, 1991,
based on data by
Capdevila et al., 1977
Talhuis – Carrizal-type granitoids K – Ar, bi 245 ± 11 and 233 ± 10 intrusive into the San Soler, 1991
Ramón-type batholiths
Southern Peru: Quillabamba granite U – Pb 257 ± 3 Lancelot et al., 1978
Eastern Cordillera Abancay ‘‘quartz – diorite’’ U – Pb 222 ± 7 cataclastic Dalmayrac et al., 1980
monzodiorite dyke (Fig. 4) K – Ar, hb 185 ± 6 this paper (Table 2)
rhyolite dyke (Fig. 4) K – Ar, bi 244 ± 6 this paper (Table 2)
Allinccápac volcanics Rb – Sr 200 – 180 peralkaline Kontak et al., 1985, 1990
Macusani syenite K – Ar 184.2 Stewart et al., 1974
Macusani syenite K – Ar 173.5 ± 3.1 Kontak et al., 1985
Coasa granite U – Pb 238 ± 11 Lancelot et al., 1978
Aricoma adamellite U – Pb 234 ± 9 Dalmayrac et al., 1980
Limacpampa 2-mica pluton Rb – Sr 199 ± 10 foliated, peraluminous Kontak et al., 1990
Southern Peru: Antarane granodiorite Ar – Ar  277 Clark et al., 1990b
Altiplano Antarane granodiorite K – Ar, bi 275.2 ± 5.8 Clark et al., 1990b
Antarane granodiorite K – Ar, mu 263.6 ± 5.2 from an associated quartz vein Clark et al., 1990b
‘‘lava flows’’ K – Ar, wr 272 ± 10 type of lava was not indicated Klinck et al., 1986
Southern Peru: Punta Coles gabbro- U – Pb 188.4 and 184 Mukasa, 1986
Arequipa monzotonalite super-unit
Southern Peru: basaltic flow, lower Chala Fm Ar – Ar  177 Romeuf et al., 1993
coast
Bolivia: Nevado Tata Sabaya granite K – Ar 188.1 ± 4.0 and slightly cataclastic and altered Sempere et al., 1998
Western Cordillera 181.6 ± 3.9
Bolivia: Altiplano ‘‘basalt flow’’ Serranı́as de Chilla K – Ar 279.9 ± 3.3 (wr), S. McBride, in
244.9 ± 2.9 (gl) Kontak et al., 1985;
ages recalculated by
Kontak et al., 1990
‘‘altered gabbro south of K – Ar 258 ± 13 Bolivian National Oil Company Saavedra et al., 1986
Lake Titicaca’’ unpublished data
Bolivia: Zongo – Yani 2-mica pluton U – Pb 222.2 + 7.7/  9.1 pervasively foliated facies Farrar et al., 1990
axis of Eastern Zongo – Yani 2-mica pluton U – Pb 225.1 + 4.1/  4.4 weakly foliated facies Farrar et al., 1990
Cordillera Cordillera Real plutons K – Ar  212 McBride et al., 1983;
Kontak et al., 1985
Cerro Sapo alkaline complex K – Ar 97.7 ± 2.8 dated material is from a Kennan et al., 1995;
breccia-pipe bearing kimberlitic clasts Tawackoli et al., 1999
Cerro Grande high-K gabbroic K – Ar, bi 120.0 ± 0.5 obtained on 2 biotite fractions Tawackoli et al., 1999
to syenitic intrusion
Cornaca basanite dyke K – Ar, wr 184 ± 4.9 Tawackoli et al., 1996, 1999
Bolivia: Tarabuco – Uyuni sill K – Ar, wr 171.4 ± 4.2 continental tholeiite Sempere, 1995
Entre Rı́os branch Entre Rı́os sill (altered) K – Ar, wr 108, 104, continental tholeiite; Bolivian Saavedra et al., 1986;
of rift and younger ages Gulf Oil Company unpublished data Sempere et al.,

T. Sempere et al. / Tectonophysics 345 (2002) 153–181


1992; López-Murillo and
López-Pugliessi, 1995
Camiri basalt sill K – Ar, wr 233 Bolivian Gulf Oil Company Saavedra et al., 1986;
unpublished data Sempere et al.,
1992; López-Murillo and
López-Pugliessi, 1995
Rejará pluton K – Ar 141 ± 10 this Precambrian pluton is intruded Aranı́bar, 1979
by dolerite dyke swarms
NW Argentina Los Alisos alkaline ultrabasic dyke K – Ar 303 ± 10 high-K Méndez and Villar, 1979
Sierra de Rangel alkaline granites Rb – Sr, bi 146 ± 1.6 and 122 ± 1.5 Menegatti et al., 1997

163
164 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

Vivier et al., 1976; Noble et al., 1978; Dalmayrac et canic rocks are probably cogenetic (Kontak et al.,
al., 1980; Carlier et al., 1982; Kontak, 1984; Bon- 1985), suggesting that the related peralkaline magma-
homme et al., 1985; Kontak et al., 1985, 1990; Clark tism developed during the Liassic and, possibly, early
et al., 1990a; Cenki, 1998). In a major contribution, Dogger. A predominantly mantle-derived magmatism
Kontak et al. (1985) clearly identified that the entire thus developed in this segment of the Eastern Cordil-
Mitu-age magmatism in southern Peru was rift- lera during the first half of the Jurassic (Kontak et al.,
related, and recognized that ‘‘the predominantly basic 1985).
Mitu Group volcanics and the batholithic granodior- In contrast with the mantle-derived Mitu volcanic
ites and monzogranites (. . .) are most probably repre- rocks, the Carabaya batholith plutons derive from
sentative of a continuum with a cause and effect crustal melts, and are also similar to plutons known
relationship’’. In particular, these authors underlined in the Oslo rift (Kontak, 1984; Kontak et al., 1985,
that this magmatism was very similar to the one known 1990). Intrusion of the main plutons occurred in the
from the Early Permian Oslo aborted rift in Norway Late Permian and Triassic (Table 1). The Carabaya
(Kontak, 1984; Kontak et al., 1985). batholith is commonly cut by coeval and younger
Several types of apparently unrelated mantle- alkaline, Ti-rich, basic dykes that display character-
derived magmas, including alkaline basalts and istics similar to the basalts known in the Mitu Group
locally thick peralkaline facies, occur among the Mitu (Kontak et al., 1985, 1990).
volcanic rocks (Kontak et al., 1985). Basic volcanics In addition to the published ages on the Mitu
can form up to 20% of the total Mitu volcanism and magmatism in southern Peru, we have obtained new
consist of tholeiitic or alkaline spilitized basalt flows K –Ar ages of 244 ± 6 and 185 ± 6 Ma, from a rhyolite
that are generally intercalated with the Mitu Group dyke and a monzodiorite dyke, respectively (Tables 1
sedimentary rocks (Vivier et al., 1976; Kontak, 1984). and 2; Fig. 4).
Swarms of basic dykes and sills intruding Paleozoic All these data agree with the idea that Late Per-
and Mesozoic strata are known from the Altiplano, mian – Middle Jurassic lithospheric thinning in the
Eastern Cordillera, and Subandean belt of southern Eastern Cordillera and Altiplano of southern Peru
Peru (Newell, 1949; Laubacher, 1978), and are likely generated a variety of mantle-derived magmas, along
to have been produced by the Triassic – Jurassic litho- with a high heat flow that produced significant
spheric thinning as in nearby Bolivia. amounts of crustal melting in the Late Permian –
A Permian, early stage of magmatism is docu- Triassic (Kontak et al., 1985). In this area, lithospheric
mented in the Peruvian Altiplano by dated ‘‘lava thinning lasted over nearly 100 Ma.
flows’’ from 21 km northwest of Juliaca, and by the
Antarane granodiorite  30 km WSW of the same 2.3.2. Bolivia
city (Table 1; Fig. 4). Rift-related magmatism in Bolivia was dominated
In the Eastern Cordillera, the 2000-m-thick Early by basic magmas. Plutons derived from crustal melts
Jurassic, Allinccápac peralkaline volcanic rocks sur- are only known northwest of 17S, i.e. in the prolon-
round and overlie a large peralkaline syenite complex gation of the Eastern Cordillera of southern Peru, and
that yielded Early to Middle Jurassic apparent ages have yielded Late Triassic U – Pb and K – Ar ages
(Table 1). The syenite complex and surrounding vol- (Table 1). The basic magmatism related to the main

Table 2
Analytical data concerning the new ages presented in this paper
Sample # (lab. #) Nature Location Method and %K Rad. Ar Atm. Ar Age (Ma) ± 2 s
material dated (nl/g) (%)
98-08-18-01 (H475.98) rhyolite dyke  14420S – 70430W K – Ar, bi 6.815 69.218 6 244 ± 6
M-371 (H289.CP) monzodiorite dyke  13320S – 72330W K – Ar, hb 0.859 6.493 21 185 ± 6
Samples were dated at the Laboratorio de Geocronologı́a of SERNAGEOMIN, Santiago, Chile; see Fig. 4 for location; bi: biotite, hb:
hornblende.
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 165

rift axis displays alkaline compositions (Aldag, 1913; 2.3.2.1. Magmatism and heat flow in the main rift
Smulikowski, 1934; Soler and Sempere, 1993; Ta- axis. The Late Triassic granitoids in the Eastern
wackoli, 1999), whereas the giant sill known in the Cordillera northwest of 17S (Cordillera Real) are
‘‘Entre Rı́os branch’’ of the rift (Sempere et al., 1998) likely to have been emplaced at depth along the
has a tholeiitic composition (Fig. 7; Soler and Sempere, Eastern Cordillera rift system, because they are similar
1993). All these basic rocks indicate ‘‘intraplate’’ to those known in nearby southern Peru (McBride et
mantle-derived magmatism and lithospheric thinning. al., 1983; Kontak et al., 1985), and include peralumi-
Although several Permian to Jurassic isotopic ages nous plutons (Bard et al., 1974).
have been obtained (Table 1), samples from these Southeast of 17S, such granitoids are unknown
commonly altered basic rocks are difficult to date and the reconstructed rift axis is characterized by
securely, and many have yielded Cretaceous and Pale- elongated swarms of basic dykes and sills that intrude
ogene apparent ages (see Saavedra et al. (1986), Ru- mainly Paleozoic rocks (Figs. 3 and 4). When Mes-
biolo (1997) and Tawackoli et al. (1999) for reviews). ozoic strata occur in areas with basic dyke swarms,

Fig. 7. Plot of some Bolivian Mesozoic basaltic rocks in Cabanis and Thiéblemont’s (1988) Tb – Th – Ta diagram (modified after Soler and
Sempere, 1993).
166 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

basaltic flows, sills and/or dykes are generally also Ipaguazú Formation, ‘‘Entre Rı́os basalt’’, and Tacurú
observed in the Mesozoic, strongly suggesting that the sub-Group in the southern Chaco Subandean belt. The
dykes in the Paleozoic are of Mesozoic age (Stein- Entre Rı́os Basalt is a  130-m-thick sill (Pádula and
mann, 1906, 1923; Kozlowski, 1934; Sempere et al., Reyes, 1958; Sempere et al., 1998); it is altered in all
1998). This strip rich in basic dykes, sills and flows known outcrops and has yielded 108 Ma, 104 Ma, and
coincides with the axis of the present-day Eastern younger apparent ages (Table 1).
Cordillera, which was recognized by Kozlowski Samples from the Tarabuco –Uyuni and Entre Rı́os
(1934) in luminous words: ‘‘Tous les gisements des sills have a continental tholeiite composition so close
roches éruptives mésozoı̈ques que je connais se trou- that they cannot be geochemically distinguished (Soler
vent exclusivement dans la Cordillera Oriental et and Sempere, 1993). Both intrude Mesozoic strata and
surtout dans sa partie centrale, où ils sont disposés reach similar thicknesses (130 – 150 m). Their respec-
le long d’une zone parallèle à la direction de la tive outcrop areas are presently only  50 km apart,
chaı̂ne (. . .)’’ (‘‘All the occurrences of Mesozoic and are separated by the narrow Interandean structural
eruptive rocks that I know are exclusively found in belt where no rock unit younger than Permian is
the Eastern Cordillera and especially in its central known (Fig. 4). All this strongly suggests that, prior
part, where they are distributed along a strip parallel to to Andean deformation, the two basalt sills formed
the mountain belt’’). Kozlowski (1934), after Stein- one same giant sill (Sempere et al., 1998), which we
mann (1923), correctly conjectured that the alkaline propose to name ‘‘Tarabuco – Entre Rı́os sill’’. Given
basic rocks they collected and studied from the East- the alteration of the Entre Rı́os basalt, the early Dog-
ern Cordillera of Bolivia were probably of Triassic – ger age obtained on the Tarabuco basalt should be
Jurassic age. An unaltered basanite dyke that intrudes considered as a better estimate of the time of emplace-
Ordovician rocks northeast of Tupiza yielded a Toar- ment of this sill. Because Bolivia was part of the
cian age (Table 1). Gondwana craton prior to the Cretaceous (Sempere,
1995), it is interesting that latest Liassic – earliest
2.3.2.2. Magmatism in the ‘‘Entre Rı́os Branch’’ of the Dogger giant tholeiitic sills are known in South Africa
rift system. The ‘‘Tarabuco basalt’’ (from the area and Antarctica (Encarnación et al., 1996; Fleming et
located from  30 km north to  20 km southeast of al., 1997), as these regions were also located along the
Tarabuco) and the ‘‘Uyuni –Incapampa basalt’’ (from southwestern edge of Gondwana and underwent co-
the area located  30 to  50 km south of Tarabuco) eval lithospheric thinning.
intrude pre-Maastrichtian Mesozoic strata and belong A feeding dyke of the Entre Rı́os sill can be
to the same  100– 150-m-thick basaltic sill, here- observed at Abra Castellón, 8 km northwest of Entre
after called the Tarabuco – Uyuni sill. In the Uyuni– Rı́os, and it is possible that the dykes known west of the
Incapampa syncline, this noteworthy sill clearly Chaco Subandean belt (see below, and Fig. 4) served as
intrudes a red,  400-m-thick, fining-upward succes- feeding dykes of the giant sill. Present known differ-
sion that has yielded trackways of primitive tetrapods ences in rock composition does not preclude this
(currently under study, but suggesting a Triassic age) hypothesis, as alkaline and tholeiitic magmatisms can
and should be ascribed to the Ipaguazú ( = Sayari) coexist during rifting. Furthermore, the early Dogger
Formation (Sempere et al., 1998); the sill displays apparent age of the giant sill agrees with the Toarcian
gabbro facies in its middle part, and branches into apparent age obtained on a dyke from the Tupiza area.
three to four large sills in the west. Near Uyuni del This suggests that magmatic activity, including alkaline
Pilcomayo, 0.3 – 3-m-thick basic dykes are known and tholeiitic magmas, was significant in southern
from below and above the sill. The Tarabuco –Uyuni Bolivia in the late Liassic –early Dogger (190 –170
sill was sampled 5 km north of Tarabuco, where it Ma).
intrudes eolian sandstones, and yielded an early Mid- Our recent field study of the Camiri basalt (sensu
dle Jurassic apparent age (Table 1). Sempere et al., 1998) has showed that it is a  30-m-
The relationships and characteristics of the sedi- thick sill that intrudes the Pennsylvanian – Early Per-
mentary and igneous rocks in the Tarabuco – Incapa- mian Cangapi Formation in a small area southeast of
mpa area are nearly identical to the set formed by the Camiri (Fig. 4). This also altered rock yielded a
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 167

Middle Triassic apparent age (Table 1), suggesting At the northern foot of Nevado Tata Sabaya in the
that basic magmatism may have been active in the Western Cordillera of Bolivia, a slightly cataclastic
area as early as the Triassic; this age, however, needs and altered granite crops out below the Cenozoic
confirmation. No geochemical data are available on cover and possibly belongs to the Precambrian ‘‘Are-
the Camiri sill yet, but its overall aspect and relation- quipa massif’’ (Rivas, 1989). This granite yielded two
ships are similar to the Tarabuco – Entre Rı́os sill. Toarcian apparent ages (Table 1). Plausible hypoth-
eses are that these ages reflect a significant thermal
2.3.2.3. Other areas of magmatism and heat flow in reset in the Toarcian, or emplacement of the pluton at
Bolivia. Apart from the giant tholeiitic sill present in that time; both hypotheses agree with coeval exten-
the ‘‘Entre Rı́os branch’’, the Mesozoic magmatism sion-related magmatism and/or heat flow in the
known in Bolivia has proved so far to be of alkaline nearby southern extension of the Arequipa basin (see
and/or intraplate type, inclusively in some areas that below).
were apparently located outside the rift system (Fig. 4; Lead – zinc( – silver) ore deposits are commonly
Soler and Sempere, 1993; Tawackoli et al., 1999). associated with ancient rifts, and the stratabound
Close to the Bolivian border in northwestern Argen- deposits known in the Late Triassic – Liassic Pucará
tina and west of the Chaco Subandean belt, a potassic Group of central Peru are no exception (Rosas and
alkaline ultrabasic sheet intrusion (Meyer and Villar, Fontboté, 1995). Although they occur in the Paleo-
1984) is derived from a deep lithospheric mantle zoic, the lead –zinc( –silver) ore deposits known in the
source (Barbieri et al., 1997) and might represent Eastern Cordillera of Bolivia are distributed along
one of these manifestations (Rubiolo, 1997), in spite both sides of the main rift axis and thus possibly
of an apparent age of 303 ± 10 Ma (Table 1). North of formed at depth in relation with the rift system
the border,  55 km SSW of Tarija, the Rejará pluton, (Sempere et al., 1998).
a Neoproterozoic or Cambrian monzonite to grano-
diorite that underwent some cataclastic metamor- 2.3.3. Central Peru
phism, is intruded by dolerite dykes that also cut Intense magmatism also occurred in central Peru
overlying Ordovician strata; the K –Ar apparent age during the Mitu rifting (Mégard, 1978; Dalmayrac et
of 141 ± 10 Ma obtained on this pluton (Table 1) al., 1980; Carlier et al., 1982; Soler, 1991). Numerous
possibly reflects a Jurassic emplacement of the doler- plutons intrude the metamorphic basement exposed in
ite dykes and related heat flow (Rubiolo, 1997; the Eastern Cordillera and were probably emplaced in
Sempere et al., 1998). the rift ‘roots’. They consist of a variety of granites,
A basalt ‘‘flow’’ from the Serranı́as de Chilla south granodiorites and alkaline granitoids, which appa-
of Lake Titicaca yielded apparent Permian and Early rently display two clusters of ages: a set of mainly
Triassic ages (Table 1). In the same region, a Permian northern plutons was emplaced during the Mississip-
apparent age was obtained on an altered ‘‘gabbro’’ pian, and a larger one during the Late Permian and
from an unprecised area south of Lake Titicaca (Table Triassic (Jacay et al., 1999). This suggests that Late
1); given its apparent age, the sampled basic rock is Permian – Triassic rifting developed along an area
probably related to the Serranı́as de Chilla basalt. where magmatism had already been significantly
Permian magmatism is also known west of Juliaca active in the Mississippian.
in nearby Peru (Fig. 4). Soler (1991) recognized the following chronology
In the northwestern Beni Subandean belt, in the of pluton emplacement in the Eastern Cordillera of
Rı́o Yanamanu (Serranı́a de Uchupiamonas), a 8-m- central Peru (Table 1): (1) the Equiscocha granite; (2)
thick lava flow or sill, of ‘‘andesitic’’ aspect, occurs in the San Ramón-type batholiths, which are intruded by
the Jurassic Beu Formation, 400 m above its base (3) the Talhuis– Carrizal-type granitoids; and (4) basic
(Ponce de León et al., 1972). In the same area, dykes (including microgabbros and microdiorites) that
numerous basic dykes and sills intrude Early Devon- cut through the latter. This chronology is based in part
ian strata, whereas grains of basic volcanic rocks are on the observation that the Mitu Group syn-rift deposits
common in the Beu Formation sandstones (Ponce de contain clasts derived from the erosion of San Ramón-
León et al., 1972). type granitoids, but so far no clasts derived from
168 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

Talhuis– Carrizal-type granitoids. These relationships plutonic rocks that do not intrude overlying units (this
support the idea that rifting developed in two stages. is particularly clear in the Nuñoa area of southern
The early, small Equiscocha pluton is a fine- Peru; Fig. 4). Paleoenvironments identified in the
grained hololeucocratic granite that contains abundant Mitu Group include alluvial fans, fluvial depositional
muscovite and scarce small-size garnets, and is systems and (playa-)lakes.
derived from crustal melting. The large San Ramón- A thick coarsening-upward unit of reddish sand-
type granitoids are metaluminous to slightly peralu- stones and conglomerates known in the Ocoña –
minous, calc-alkaline granites that closely resemble Camaná area of coastal southern Peru is traditionally
the Mitu granitoids from southern Peru (Soler, 1991) assigned to the Mitu Group, but this attribution needs
and thus were probably mainly derived from crustal confirmation (see Soler, 1991 for a summary).
melts; Sr isotopic ratios (Capdevila et al., 1977) from
the San Ramón batholith suggest that these crustal 2.4.2. Syn-rift basins in Bolivia: the lower Serere
melts partly mixed with mantle-derived magmas Group
(Soler, 1991). In contrast, the Talhuis– Carrizal-type The lower part of the Serere Group of Bolivia
granitoids have a marked alkaline affinity and show (Sempere et al., 1998) comprises the red-coloured
many characteristics of granitoids emplaced in litho- Ipaguazú and Tiquina formations (because correla-
spheric thinning settings; the late basic dykes have tions seem now secured, the more recent denomina-
alkaline compositions similar to the alkaline basalts tion ‘‘Sayari Formation’’ should be abandoned in
from southern Peru, and were comagmatic with the order to simplify the stratigraphic nomenclature).
Talhuis– Carrizal-type granitoids (Soler, 1991). The Ipaguazú Formation consists of red mudstones
As in southern Peru, volcanic rocks associated with with subordinate sandstones and local evaporites
the Mitu rift system in central Peru are predominantly (gypsum, rarely halite) of dominantly lacustrine ori-
acid, and are mainly represented by dacitic to rhyolitic gin, and crops out in the Chaco Subandean belt
pyroclastic rocks. Most of the Mitu volcanics of (mostly in the ‘‘Entre Rı́os branch’’ of the rift axis)
central Peru have alkaline compositions and are prob- and in the main rift axis of the Eastern Cordillera. The
ably comagmatic of the Talhuis– Carrizal granitoids Tiquina Formation predominantly consists of interca-
(Soler, 1991). lations of red sandstones and mudstones, and of
Present knowledge of the rift-related magmatic locally thick conglomeratic sandstones or basalt-clasts
evolution in central Peru suggests that lithospheric conglomerates, and is of alluvial origin; it crops out in
thinning initially produced crustal melts (e.g., the some localities of the Eastern Cordillera (including the
Equiscocha granite) that were partly contaminated at main rift axis) and in the Tiquina area. The Ipaguazú
a later stage by mantle-derived melts (San Ramón- and Tiquina formations respectively represent fine-
type batholiths), and that these mantle-derived alka- grained and coarse-grained end-members of the Boli-
line magmas finally dominated (Talhuis – Carrizal- vian equivalent of the Mitu Group. Both units are
type granitoids and comagmatic basic dykes, and Mitu generally  500 m thick and conformably overlain by
volcanics). the thick fluvio-eolian sandstones of the Ravelo For-
mation (Andean domain) or Tacurú sub-Group (Chaco
2.4. Syn-rift deposits Subandean belt). The Ipaguazú and Tiquina forma-
tions generally occur at specific localities or in restricted
2.4.1. Syn-rift basins in Peru: the Mitu Group areas, suggesting they were deposited in paleograbens,
The Mitu Group consists of a red to purple, locally whereas the overlying fluvio-eolian sandstones are
>2000-m-thick succession of conglomerates, sand- present over much broader regions.
stones and mudstones, with local carbonates and In the rift axis, the Ipaguazú or Tiquina formations
evaporites, that accumulated in subsident grabens include basalt flows in their lower part; at four Andean
(Mégard, 1978; Dalmayrac et al., 1980; Carlotto, localities, these flows overlie the Late Permian –Early?
1998). These sedimentary rocks are commonly inter- Triassic Vitiacua or Chutani formations (see above).
bedded with locally dominant volcanic and volcani- The Ipaguazú Formation unconformably overlies the
clastic rocks, and/or intruded by subvolcanic to Vitiacua Formation in the ‘‘Entre Rı́os branch’’ of the
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 169

rift axis. The observed successions characteristically 2.5.2. Post-rift deposits in the Bolivian Orocline, Peru
include an unconformable basal unit, generally no and Bolivia: the upper Serere Group
more than several tens of meters thick, that consists In the Cusco – Sicuani area in southern Peru, the
of pale alluvial or eolian sandstones, reddish conglom- Mitu Group is post-dated by fluvio-eolian sandstones
erates and/or basalts; this basal unit rapidly grades into (Caycay Formation; Carlotto, 1998) that are locally
a thick, brown-red to brown-purple, mudstone-domi- intercalated with basalt flows. Northwest of Lake
nated unit of alluvial to lacustrine origin. Titicaca, the fluvio-eolian sandstones of the Quilca-
At Quebrada Aymaraj Hueko, 6 km southwest of punco Formation conformably overlie the Mitu Group
Torotoro (Fig. 4), a previously unknown >300-m- (where present) or the Paleozoic (outside the Mitu
thick clastic unit overlies the Vitiacua Formation grabens) (Sempere et al., 2000a). The Quilcapunco
with a very rapid transition, and represents a local, Formation is overlain by the Sipı́n Formation carbo-
coarse facies of the Tiquina Formation that strongly nates, which yielded Rhaetian –early Bajocian echinids
resembles some deposits in the Mitu Group of Peru. (see below). The red mudstones of the Muni Formation
The 2 –6-m-thick basal conglomerate of this unit unconformably overlie the latter, and are transitionally
contains abundant basaltic clasts up to 15 cm in overlain by the fluvio-eolian Huancané s.s. Formation
size. The overlying strata dominantly consists of (Fig. 2; Sempere et al., 2000a). The Muni Formation
basalt-clasts conglomerates and red conglomeratic was deposited in an alluvial to coastal plain environ-
sandstones, in which at least one basalt flow is ment and contains thin marine intercalations with
intercalated, and is intruded by a  100-m-thick fossils suggestive of a late Dogger – early Malm age
basalt sill. Furthermore, this Tiquina Formation and (Newell, 1949). The Sipı́n and Muni formations thin
underlying strata are folded and overlain by the out to the north and east, and the fluvio-eolian Quilca-
Cretaceous Torotoro Formation with a clear angular punco and Huancané s.s. formations become coales-
unconformity. cent in these directions (Sempere et al., 2000a).
These Jurassic fluvio-eolian units from southern
2.5. Post-rift deposits Peru correlate in Bolivia with the Beu Formation of
the Beni Subandean belt, the Ravelo Formation of
2.5.1. Post-rift deposits northwest of the Bolivian Andean Bolivia, and the Tacurú sub-Group of the
Orocline, Peru: the Pucará Group Chaco Subandean belt, which all are of similar facies
The Mitu rifting generated a thermal sag that and origin (Oller and Sempere, 1990; López-
progressively expanded the basin. The Pucará Group Pugliessi, 1995; López-Murillo and López-Pugliessi,
carbonates were deposited over the Mitu Group dur- 1995; Sempere, 1995; Sempere et al., 1998). These
ing the Norian –Liassic interval (Mégard, 1978; Stan- unfossiliferous sandstones can be over 1000 m thick
ley, 1994) and onlapped across the rift shoulders. and locally include basaltic flows and sills, and
They reflect a transgression that initiated in the Norian conglomerates with basalt clasts. Their distribution
and progressed from north to south following the Mitu shows they onlapped laterally from the initial rift axis
rift axis (Mégard, 1978; Loughman and Hallam, 1982; (Sempere, 1995). Their top is an erosional surface that
Rosas et al., 1997; Sempere et al., 1998). Maximum is onlapped by Cretaceous strata (Puca Group).
inundation is marked by the organic-rich shales and The Tacurú sub-Group includes, in stratigraphic
marls of the < 50-m-thick, late Rhaetian –early Sine- order, the (San Diego-)Tapecua (fluvio-eolian sand-
murian Aramachay Formation. To the east, in the stones), Castellón (fluvial sandstones and subordinate
Peruvian Oriente, red alluvial and eolian strata (lower red and green mudstones), Ichoa (eolian sandstones)
Sarayaquillo Formation) grade westwards into the and Yantata (fluvial sandstones partly equivalent to
Pucará carbonates (Mégard, 1978). Basalts with an the Ichoa) formations (see discussion in Sempere et
intraplate signature commonly occur in the Pucará al., 1998). Semionotiformes fish occur in the Castel-
Group, which hosts a number of lead – zinc( –silver) lón Formation (det. M. Gayet, Université de Lyon,
stratabound ore deposits (Rosas and Fontboté, 1995; France), and are also common in the Late Triassic/
Rosas et al., 1997). The Pucará Group is unknown in Early Jurassic lower Tacuarembó Formation of the
the Cordillera Oriental southeast of Cusco. Paraná basin in Uruguay (Sprechmann et al., 1981).
170 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

Four new species of ostracods (Bisulcocypris laci- In the coastal area, volcanic and volcaniclastic rock
niata boliviana, B. truncata, B. castellonensis, B. units traditionally correlated with the Chocolate For-
lubimovae) from the Castellón Formation are similar mation (Bellido and Guevara, 1963) were paleonto-
to Jurassic taxa of the Paraná basin in Brazil but were logically and isotopically dated (Roperch and Carlier,
tentatively assigned to the lowermost Cretaceous 1992; Romeuf et al., 1993, 1995). In the Chala area, a
(Wealdian) due to a traditional attribution of the unit basaltic flow from the lower part of the Chala For-
to the Cretaceous (Damiani-Pinto and Sanguinetti, mation yielded an early Dogger age (Table 1), dem-
1987). In spite of apparent discrepancies, all paleon- onstrating that this >3000-m-thick unit represents the
tologic ages published on the Vitiacua through Cas- southern extension of the Dogger Rı́o Grande For-
tellón stratigraphic set in the Chaco Subandean belt mation (Rüegg, 1956; Caldas, 1978); both formations
belong to the Late Permian –lowermost Cretaceous unconformably overlie Precambrian and Late Paleo-
interval (Sempere et al., 1998). zoic rocks. In the Tacna area, the ‘coastal’ Chocolate
Abundant eolian sandstones in Bolivia and south- Formation is overlain by the volcano-sedimentary,
east Peru complement to the west the large Jurassic >3000-m-thick Guaneros Formation, the base of
desertic domain defined by coeval eolian units in the which has yielded late Bajocian– Bathonian ammon-
Paraná and Karoo basin (e.g., Francßa et al., 1995). ites (Romeuf et al., 1993, 1995) and apparently
Global paleoclimatic models place Bolivia and south- records a sea level maximum. This ‘coastal’ Choco-
east Peru within a broad desertic domain in the late formation is constituted by a >3000-m-thick
Jurassic (Chandler et al., 1992). volcano-sedimentary succession intruded by Hettan-
gian to Toarcian granodiorites (Clark et al., 1990a;
Romeuf et al., 1993) and is thus likely to include
3. Early and Middle Jurassic lithospheric thinning Triassic deposits.
in southwestern Peru The Rı́o Grande, Chala, and Guaneros formations
show geochemical characteristics which suggest that
3.1. Introduction they accumulated in a subduction-related volcanic
arc setting (Romeuf et al., 1993, 1995). The consid-
The Arequipa basin (Fig. 4) originated by litho- erable thickness of these units points to high sub-
spheric thinning during the Liassic and Dogger. The sidence rates, and rather suggests that these volcanic
Jurassic infill of this basin is formed by a 4500– 6000- and volcano-sedimentary rocks accumulated in an
m-thick succession that provides a prime sedimentary extensional back-arc setting, close to the arc proper.
record of the regional geologic evolution (Jenks, As the coastal formations attributed to the Choc-
1948; Benavides, 1962; Vicente, 1981, 1989; Vicente olate Formation are neither chronologic nor genetic
et al., 1982). We refer to this succession as the ‘‘Yura equivalents, we suggest that these homonymous vol-
Group’’, in a modified sense. As used here, the Yura canic units should now be prudently distinguished.
Group comprises, from bottom to top, the Chocolate Therefore, the arc setting reconstructed for the Middle
s.s., Socosani, Puente, Cachı́os, and Labra formations, Jurassic coastal volcanic rocks should not be gener-
because we think that this name should reflect the alized to the Late Triassic – early Liassic Chocolate s.s.
entire activity of the basin in which it accumulated. Formation of the interior Arequipa basin.

3.2. Igneous rocks 3.3. Late Triassic –early Bajocian record of incipient
lithospheric thinning
The dominantly volcanic and volcaniclastic Choc-
olate s.s. Formation (i.e., sensu Jenks, 1948) is >900– Although the Chocolate s.s. volcanic rocks remain
1500 m thick (its base rarely crops out). Near its top, it virtually unstudied, the association of this Late Tri-
includes limestone beds that yielded Sinemurian assic –early Liassic thick volcanic unit with the over-
ammonites, and is disconformably overlain by the lying late Liassic– Bajocian Socosani shallow-marine
late Liassic carbonates of the Socosani Formation carbonates is strongly reminiscent of the genetic link
(Vicente, 1981). between the Late Permian – Triassic Mitu Group syn-
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 171

rift volcanics and the overlying Late Triassic – Liassic The accumulation of the thick Chocolate s.s. vol-
Pucará Group sag shallow-marine carbonates in cen- canic rocks reflects intense tectonic subsidence, as
tral Peru. Following Vicente et al.’s (1982) sugges- do the overlying, sedimentary, Toarcian– Bathonian
tion, we thus propose that the Arequipa basin Socosani and Puente formations, in which synsedi-
originated by rifting, and given the high subsidence mentary extensional features are abundant (Vicente et
evidenced in the coastal area, that this rifting probably al., 1982). Toarcian emplacement of the Punta Coles
developed in an extensional back-arc setting. gabbro-monzotonalite super-unit (Table 1) must have

Fig. 8. Stratigraphy of the Arequipa basin (modified after Vicente, 1989). Age indications are based on fossils (Vicente, 1981, 1989).
172 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

occurred in the same framework; the area of emplace- 3.5. Filling-up of the Arequipa basin in the Oxfor-
ment of these plutons probably underwent more in- dian –Kimmeridgian
tense lithospheric thinning and is located southwest of
the classical outcrop area of the Yura Group, where no The upper part of the Cachı́os Formation shows
synsedimentary magmatic manifestations are known. shallowing-upward facies and grades into the 300 –
In the northeastern extension of the Arequipa 1500-m-thick sandstone-dominated Labra Formation
basin, we found Diademopsis sp., an echinid genus (Vicente, 1981), which was deposited in a siliciclastic
indicative of the Rhaetian – early Bajocian interval shelf to shoreface setting (Vicente et al., 1982). The
(Thierry et al., 1997), in marine limestones of the overall thickening- and coarsening-upward trend indi-
Sipı́n Formation (Newell, 1949). This 0– 40-m-thick cates deltaic-like progradation. Sandstones are com-
unit is thus correlative of the Socosani Formation and monly interbedded with shallow-marine limestones
reflects the same regional subsidence-related trans- (Vicente, 1981). The age of the Labra Formation is
gression (Fig. 2). bracketed by late Callovian ammonites in the under-
lying Cachı́os Formation and early Tithonian ammon-
3.4. Downwarping of the Arequipa basin in the middle ites in the overlying Gramadal Formation (Vicente,
Dogger 1989); the Labra thus appears mostly of Oxfordian –
Kimmeridgian age (  159 –151 Ma).
Lithospheric thinning culminated in the late Bajo- Progradation of the thick Labra shallow-marine
cian – early Callovian (  162– 167 Ma) with consid- sandstones onto relatively deep-marine shales implies
erable deepening and tectonic downwarping of the that the basin shallowed markedly in the Oxfordian –
Arequipa basin. Early Bajocian shallow-marine lime- Kimmeridgian, while remaining very subsident. Paleo-
stones are abruptly overlain by late Bajocian ‘starved currents and cumulative sandstone thicknesses indi-
basin’ facies (Upper Socosani Formation) that are in cate that sands were being derived from the north and
turn overlain by a 700-m-thick Bathonian and early northeast. This invasion by sands perceptibly de-
Callovian turbidite succession (Puente Formation) creased during the early Tithonian (Gramadal Forma-
(Vicente et al., 1982; Vicente, 1989; Fig. 8). The tion), when transgressive shallow-marine carbonates
turbidites were deposited in an elongated trough were commonly deposited southwest of the present-
parallel to the present Andean trend, and show day Western Cordillera.
NW ! SE paleocurrents.
The overlying,  500-m-thick, Cachı́os Formation
predominantly consists of organic-rich shales. Subor- 4. Late Jurassic – earliest Cretaceous tectonism in
dinate sandstones are distributed in channels and in the Eastern Cordillera
coarsening-upward slumps and olistolites. Facies
overall indicate a submarine slope paleoenvironment The Eastern Cordillera of central Peru is tradition-
(Vicente, 1981; Vicente et al., 1982). The Cachı́os ally believed to have behaved as a structural high
Formation has yielded early and late Callovian (‘‘Marañón geanticline’’ or ‘‘Axial Swell’’) since the
ammonites (Vicente, 1989). Late Triassic (Mégard, 1978, 1987; Dalmayrac et al.,
Callovian ammonite-bearing shales and minor 1980; Jaillard, 1994), mainly because in this area Early
coarser sediments are widespread southwest of Lake Cretaceous strata onlap Precambrian and Paleozoic
Titicaca (Douglas, 1920; Jenks, 1948; Newell, 1949; rocks and are much thinner than to the west and east.
Benavides, 1962; Bellido and Guevara, 1963; Portu- Our reconstruction of a Late Permian – Liassic rift
gal, 1974; Vicente, 1981), which probably indicates system along the same area implies instead that syn-
that the Jurassic regional maximum inundation oc- rift, and probably thermal sag, deposition must have
curred during this stage. In the northeastern extension occurred in the Eastern Cordillera domain during this
of the Arequipa basin, fossiliferous marine beds time span—as is indeed observed at a few localities
intercalated in the Muni Formation (Newell, 1949) (Rosas and Fontboté, 1995). The absence of Late
are correlative of this late Dogger highstand (Fig. 2; Triassic – Jurassic deposits and the Early Cretaceous
Sempere et al., 2000a). onlap must thus reflect that this area underwent uplift
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 173

and erosion before the Early Cretaceous. Uplift of this tion (as correctly described by Newell (1949); these
previously rifted area suggests that some kind of gentle observations are at odds with Portugal (1974) and
rift inversion occurred in the Late Jurassic and/or ear- Jaillard and Santander (1992)). This demonstrates that
liest Cretaceous. the Jurassic marine strata near Lagunillas were
Latest Jurassic –earliest Cretaceous erosion in the deformed in the latest Jurassic and/or Early Creta-
Eastern Cordillera of central Peru is documented by ceous, and that reliefs were tectonically created in the
the Copuma (J. Jacay, unpublished) and Upper Sar- area at that time. Lower temperature portions of
39
ayaquillo conglomerates, which overlie Lower to Ar – 40Ar spectra ‘‘strongly suggest’’ that the region
Middle Jurassic strata respectively west and east of  35 km north of the Lagunillas area underwent
the ‘‘Marañón geanticline’’. The Copuma conglomer- thermal overprinting at  130– 120 Ma (Clark et al.,
ates underlie the Valanginian – Aptian Goyllarisquizga 1990b).
Group with an angular unconformity, and, to the west, In Bolivia, east-derived conglomerates are domi-
grade into red mudstones and sandstones. More to the nant in the uppermost unit of the Serere Group  25
west, their lateral equivalent is likely to be the con- km west of the Jurassic rift axis, suggesting that
glomerate-bearing Tinajones Formation, which is of erosion affected an uplifted structure derived from
probable Berriasian age (Jaillard, 1994) and forms the inversion of the rift system prior to the Early/Middle
upper unit of the Chicama Group of western central Cretaceous onlap. The angular unconformity observed
Peru (Fig. 3); older units of this group record a near Torotoro between Cretaceous strata and folded,
considerable and abrupt downwarping of the local Mitu-equivalent, volcaniclastic deposits (see above)
basin floor during the Tithonian, as well as neighbour- reflects this deformation.
ing uplifts (Jaillard and Jacay, 1989; Enay et al., Small igneous bodies emplaced after cessation of
1996). The Goyllarisquizga Group disconformably rifting show that magmatic activity did not disappear
overlies the Chicama Group in the west, and uncon- completely from the lithospheric heterogeneity cre-
formably onlaps older rocks to the east (Jaillard et al., ated by rifting (Fig. 4): the Cerro Sapo alkaline
1997), including the Precambrian in the Eastern complex includes a nepheline syenite locally enriched
Cordillera. The erosional surface at the base of the in sodalite, carbonatites, and a breccia-pipe (bearing
Goyllarisquizga Group has a regional importance kimberlitic clasts) of Middle Cretaceous age (Table 1);
(Fig. 2). the undated Carpacayma phonolite overlies the Copa-
In coastal southern Peru, the post-Dogger, pre- cabana Group  25 km southwest of Torotoro
Tithonian, unconformity identified by Rüegg (1961) (Kozlowski, 1934); farther south, a high-K, gabbroic
reflects at least local tectonic motions. Invasion of the to syenitic, Early Cretaceous intrusion occurs at Cerro
Arequipa basin by the northeast-derived, Oxfordian – Grande (Table 1); in the Argentine Puna, along the
Kimmeridgian, Labra sands suggests that they were same lineament, alkaline rocks are known (Rubiolo,
produced by coeval uplifts in the northeast, possibly by 1997) and include Early Cretaceous alkaline granites
gentle inversion of the Eastern Cordillera rift system. (Table 1).
The early Tithonian Gramadal Formation, up to 300 m Available data thus suggest that a number of uplifts
thick in the south, rapidly thins to the northeast and (and downwarps) occurred in Peru and Bolivia in the
grades into red silty mudstones of coastal plain origin, Late Jurassic – Early Cretaceous, and especially in the
as observed by us at Chivay; near this locality, this unit Tithonian –Berriasian interval (Jaillard, 1994; Sem-
abruptly overlies the Labra Formation and locally dis- pere et al., 1999); new data concerning uplifts of this
plays coarse conglomerates at its base. age in southern Peru and Bolivia will be presented in
In the Lagunillas area (Fig. 4), 150-m-thick, very detail elsewhere. Uplifts in Peru are post-dated by
coarse conglomerates overlie folded Sinemurian – easterly or northeasterly onlap of Valanginian to
Kimmeridgian strata with a marked angular uncon- Middle Cretaceous strata, and an erosional surface is
formity, and are transitionally overlain by an overall found at its base (Mégard, 1978; Laurent, 1985;
fining-upward succession consisting of conglomeratic Jaillard, 1994). In Bolivia, this erosional surface is
sandstones, sandstones and red mudstones, overlain in represented by the unconformity that separates the
turn by the late Middle Cretaceous Ayabacas Forma- Serere Group from the overlying Puca Group; depo-
174 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

sition on this surface started in the Middle Cretaceous Navarro, 1979; Jacay, 1996; Mégard et al., 1996;
(Sempere, 1995). Quispesivana, 1996; Jacay et al., 1999). Given these
geologic relationships, it is possible that these peri-
dotite bodies were tectonically emplaced in the Juras-
5. Consequences of Late Permian – Jurassic sic in relation with major stretching and/or wrenching
lithospheric thinning for Andean shortening of the crust, if, due to its earlier onset, rifting in this
segment reached a state significantly more advanced
5.1. Andean inversion of the Eastern Cordillera rift than in southern areas. It is also possible, however,
system that these peridotite bodies are fragments of Precam-
brian or Paleozoic mantle lithosphere that were tec-
As underlined above, the axis of the well docu- tonically emplaced into younger rock units during rift
mented Late Permian – Jurassic rift system closely inversion due to particularly intense upward expulsion
coincides with the axis of the present-day Eastern of deep material. In any case, the occurrence of
Cordillera of Peru and Bolivia. This coincidence peridotite bodies in this context indicates that in
strongly suggests that the Eastern Cordillera results central Peru rift inversion partly reworked structural
from tectonic inversion of this rift system. It is logical levels as deep as the mantle lithosphere.
that many thrusts in the Eastern Cordillera originated Because rift inversions can affect different struc-
by compressional to transpressional reactivation of tural depths, the amount of inversion-derived uplift
earlier extensional or transtensional faults (Sempere, can also be perceived from the distribution of Late
2000). The dip and overall geometry of the Late Paleozoic –Jurassic granitoids. Although the rift sys-
Permian– Jurassic rift faults determined the vergence tem continues into Ecuador (Rivadeneira and Baby,
of many Andean-age thrusts, as west (resp. east)- 1999), the abundance of exposed granitoids typically
vergent thrusts are predominant west (resp. east) of decreases north of 6S, where they nearly disappear.
the paleo-rift axis. In the N-trending segment of the This suggests that in Peru shortening in the Eastern
Eastern Cordillera, south of 19S, Andean-age tec- Cordillera considerably decreases north of 6S.
tonic displacements are likely to have been initially Shortening and/or depth of rift inversion in the
more transpressional (Hérail et al., 1996; Tawackoli, Eastern Cordillera apparently also decrease southeast
1999) due to the obliqueness of pre-existent structures of 17 – 18S, where exposed granitoids disappear and
relative to stress (although more recent, pure compres- only basic dyke swarms crop out southwards on (Fig.
sional slips have obliterated striations produced by 3). Lower shortening in this region is also reflected by
older motions on these faults). the fact that this segment of the Eastern Cordillera,
Tectonic inversion and wrenching were more although of tectonic origin, is not a true, high and
intense near the axis of the main rift system, as narrow, cordillera but a broad highland region affected
suggested by exposures of structurally deeper regions by large-scale erosional surfaces.
of the rift. For example, in the Eastern Cordillera
northwest of 16300S, the granitoids that had been 5.2. Paleotectonic status of the Altiplano
emplaced in the rift ‘roots’ are now exposed at the
highest altitudes. Outcrops of these ‘roots’ are gen- The Bolivian Orocline is the second most prom-
erally composed of Precambrian to Early Paleozoic inent mountain range on earth. Crustal thickness
metamorphic rocks (Sempere et al., 1999); along this locally reaches 75 km (Beck et al., 1996), a figure
axis, granitoids and metamorphic rocks northwest of comparable to the maximum crustal thickness in the
16300S are commonly intruded by basic dykes, as are Himalayas, which contrastingly result from conti-
non-metamorphic Paleozoic strata southeast of nental collision. In this region is found the Alti-
16300S. plano, which in size is the second high plateau on
In central Peru, poorly studied peridotite bodies earth after Tibet. Under a geomorphic point of view,
occur within Precambrian metamorphic rocks, Mis- the Altiplano is a large endorrheic basin largely
sissippian plutons and strata, and at the Mitu/Pucará covered (and filled) by Cenozoic sedimentary and
contact (Aumaı̂tre et al., 1977; Grandin and Zegarra- volcanic deposits. It is bounded by the Western
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 175

Cordillera, a high elongated area formed by clusters lithosphere (Myers et al., 1998; Schmitz et al., 1999).
of Neogene volcanoes, and by the Eastern Cordil- However, the current crustal thickness below the
lera. In contrast with Tibet, the origin of the Altiplano varies from  55 – 60 km at  16S to
Altiplano remains largely obscure and is currently 70 –74 km at 20S (Beck et al., 1996) and, because
a matter of vigorous debate. these values considerably exceed ‘‘normal’’ crustal
Although outcrops of pre-Cenozoic rocks are rare thickness, the crust below the Altiplano must have
on the Altiplano s.s. (i.e., excluding its northern been thickened during the Andean orogeny.
region), it appears that this domain includes Precam- However, the fact that the Altiplano underwent
brian rocks, a poorly known Paleozoic cover, and, in little surface shortening in the Cenozoic (  15 km,
its easternmost part, thin remnants of the Jurassic i.e. < 10%; Rochat et al., 1999) indicates that its upper
Ravelo Formation (which thickens to the east, i.e. crust behaved relatively rigidly during the orogeny.
toward the Triassic – Jurassic rift axis); the Altiplano Consistently, propagation of upper crustal shortening
domain was only later onlapped by Late Cretaceous or apparently ‘‘jumped’’ across the Altiplano in the late
younger strata (Sempere, 1995). Because there is no Oligocene, from an area west of the Altiplano to an
evidence for pre-Late Cretaceous erosion of would-be area that at least partly coincides with the Eastern
Mesozoic accumulations from the Altiplano, these Cordillera (Sempere et al., 1990). This implies that the
relationships indicate that the Altiplano domain was Eastern Cordillera resulted from the compressional
submitted to little or no subsidence during the Trias- failure of the homonymous rift system, and not from
sic –Jurassic interval. progressive eastward propagation of deformation in
The Eastern Cordillera bounds the present-day the upper crust (Sempere, 2000).
Altiplano to the east and, because it results from the Because thicknening of the Altiplano upper crust
inversion of the Triassic – Jurassic rift system, the was weak, crustal thickening mostly affected its lower
latter must have bounded the Altiplano domain prior crust. This decoupling between the evolutions of the
to the Andean orogeny. To the west, the Altiplano Altiplano upper and lower crusts (see also Yuan et al.,
domain was bounded by the Arequipa basin, which 2000) can have been caused by ductile flow into the
apparently continued southwards into northern Chile Altiplano lower crust from overthickened lower crust
(Muñoz et al., 1988; Muñoz and Charrier, 1993; beneath the Western and Eastern cordilleras (Sempere
Vicente, 1989). Although more studies remain neces- et al., 2000b). This important issue, however, remains
sary, Dogger rifting is recognized in the north-Chilean a matter of debate.
Precordillera (22S; Günther et al., 1997), whereas
Triassic – Liassic rifting is documented in the Domey- 5.3. Andean-age inversion of other basins
ko Cordillera (25 – 26S; Mpodozis and Cornejo,
1997); and Early to Middle Triassic rapid uplift and In central Peru, the main Andean thrusts appear to
exhumation of the Limón Verde block (Franz and coincide with paleogeographic boundaries (Janjou et
Lucassen, 1997) probably reflect coeval rifting west al., 1981; Mourier, 1988), suggesting that they derive
of this area. Triassic –Jurassic lithospheric thinning is from inversion of Mesozoic structures produced by
thus very likely to have developed west of the entire lithospheric thinning (Jaillard, 1990). In southern
present-day Altiplano (Figs. 3 and 4). Peru, the compressional tectonics that affects the
A major consequence of this analysis is that, before Arequipa basin is interpreted to be of Late Creta-
the Andean orogeny, the non-subsident lithospheric ceous –Early Paleogene age (Vicente, 1989) and pos-
domain that today corresponds to the Altiplano was sibly resulted from inversion of this thinned area.
bounded on both its east and west sides by areas Jaillard (1990) stated that ‘‘in northern Peru, the
where the lithosphere had been thinned. The litho- inversion of the crustal extensional structures and
sphere below the Altiplano must thus have had the related shortening of the superimposed sedimen-
‘‘normal’’ thickness characteristics before the Andean tary wedge can explain the whole observed crustal
orogeny. We find worthy of mention that, according to thickness, whereas in southern Peru, it acted as a
recent seismic tomography studies, the Altiplano crust minor but not negligible parameter.’’ In these areas,
is indeed still underlain by a 65– 80-km-thick mantle however, more detailed studies seem necessary to
176 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

assess the influence of pre-existent structures on lithospheric thinning occurred in the Middle Creta-
Andean shortening. ceous (Atherton and Webb, 1989; Atherton, 1990;
Farther east, in the central and northern Oriente Soler, 1991). These two episodes of lithospheric
region of Subandean Peru, Cenozoic structures appa- thinning are likely to have derived from evolving
rently derive from reactivation of pre-existent faults asthenospheric flow patterns related to large-scale
(Laurent, 1985). mantle convection. The major change recorded
at  89 Ma from stretching-dominated to shortening-
dominated conditions in the Andean lithosphere
6. Conclusions (Sempere, 1995) must therefore reflect a major change
in the regional mantle flow pattern around that time.
Recognition of pre-orogenic lithospheric thinning This conclusion is supported by increasing evidence
in the Andes of Peru and Bolivia sheds light on the that a major change in Pacific mantle dynamics
structure and building history of the Andes at these occurred at  84 Ma (Sager and Koppers, 2000).
latitudes. Although the information used to recon- In the areas that had undergone lithospheric thin-
struct subsident basins and regions of thinned litho- ning, it is obvious that the pre-orogenic crust re-
sphere does not necessarily come from the areas that mained thin until the Andean orogeny developed,
underwent the most intense lithospheric thinning, we which resulted in its thickening. On the other hand,
underline that it shows that such processes developed it remains unclear whether the mantle lithosphere was
in the Central Andes during the Late Permian –Middle partly reconstituted at the expense of cooling astheno-
Jurassic interval. sphere while heat flow waned. In any case, however,
The available evidence consistently demonstrates Andean-age shortening of the mantle lithosphere must
that the present-day Eastern Cordillera of Peru and have triggered delamination processes when its in-
Bolivia underwent significant lithospheric thinning creasing thickness reached a stability threshold (e.g.,
during the Late Permian – Middle Jurassic interval. Kay et al., 1994; Carlotto et al., 1999).
Development of this rift system is no extraordinary We suggest that regional-scale pre-existence of a
phenomenon, as coeval similar processes were com- thinned crust or lithosphere was a necessary condition
mon in western Gondwana (e.g., Tankard et al., 1995) for subsequent non-collisional development of con-
due to the contemporaneous dislocation of Pangea. siderable shortening and crustal thickening during the
Onset of rifting seems to have been diachronous, Central Andean orogeny. It is a trivial principle that
propagating from north to south. Isotopic ages for the thickening of a continental area is less difficult when
Mitu magmatism clearly tend to be older, Late Per- its crust has been previously thinned, as is illustrated
mian, in the north, although Permian (280 –260 Ma) by many ancient passive continental margins that
ages are known west and south of Lake Titicaca (Fig. have been considerably shortened (e.g., in the Teth-
6; Table 1). Mitu syn-rift strata were apparently yan belt). It is another trivial principle that the stress
deposited earlier in the north than in the south, where required for sliding on a pre-existent fault is less
they overlie a Late Permian – Early? Triassic partly than the stress required for a new fault to form. A
marine unit that was not deposited in a rift setting. consequence of this principle is that pre-existent
Carbonate deposition linked to thermal sag along the faults and other tectonic heterogeneities generally
Mitu rift axis (Pucará Group) progressed from north to influence, and may even control, the propagation of
south, but did not penetrate southeast of Cusco deformation in areas submitted to shortening. In
(Dalmayrac et al., 1980). Lithospheric thinning devel- particular, knowledge of the pre-Andean lithospheric
oped more to the southwest, in the Arequipa basin, heterogeneities is crucial to understand why and how
during the Liassic and Dogger. End of rifting in the the Bolivian Orocline formed. In this respect, the fact
Eastern Cordillera proceeded from onset of gentle that the Eastern Cordillera corresponds to a paleo-rift
inversion of the rift system in the Late Jurassic. system and the Altiplano to an unthinned paleotec-
The Late Permian – Middle Jurassic episode of tonic domain suggests that the regional pre-orogenic
lithospheric thinning developed to the east and south- structure was a key factor in the formation of the O-
east of coastal central Peru, where further considerable rocline.
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 177

Acknowledgements in der chilenischen Präkordillere (21 – 23S). Berliner Geowis-


senschaftliche Abhandlungen, Reihe A, Band 123, 117 pp.,
Berlin.
This study was funded by the Institut de Recherche Bonhomme, M.G., Audebaud, E., Vivier, G., 1985. K – Ar ages of
pour le Développement (IRD, previously Orstom) and Hercynian and Neogene rocks along an east – west cross sec-
conducted under an agreement with the Universidad tion in southern Peru. Comunicaciones, Santiago de Chile 18,
Nacional San Antonio Abad del Cusco (UNSAAC), 27 – 30.
Cabanis, B., Thiéblemont, D., 1988. La discrimination des tholéiites
Cusco, Peru. We thank J. Doubinger, R. Iannuzzi, E.
continentales et des basaltes d’arrière-arc: proposition d’un nou-
Robert, and J.-C. Vicente for providing us with veau diagramme: le triangle Th-3xTb-2xTa. Bulletin de la So-
important information. We thank G. Hérail for in- ciété Géologique de France (8) 4, 927 – 935.
formation regarding northern Chile, and E. Jaillard for Caldas, J., 1978. Geologı́a de los cuadrángulos de San Juan, Acari y
fruitful discussions. Yauca, Lima, Boletı́n del Instituto de Geologı́a Minerı́a y Meta-
lurgı́a (A).
Capdevila, R., Mégard, F., Paredes, J., Vidal, P., 1977. Le batholite
References de San Ramón, Cordillère Orientale du Pérou central. Geologi-
sche Rundschau 66, 434 – 446.
Aldag, A., 1913. Petrographische Untersuchung bolivianischer An- Carlier, G., Grandin, G., Laubacher, G., Marocco, R., Mégard, F.,
desit-und Diabasgesteine samt ihrer Einschlüssen. Inaug.-Dis- 1982. Present knowledge of the magmatic evolution of the East-
sertation, Bonn, Germany. ern Cordillera of Peru. Earth Science Reviews 18, 253 – 283.
Aranı́bar, O., 1979. Geologı́a regional de la parte sur de la hoja geo- Carlotto, V., 1998. Evolution andine et raccourcissement au niveau
lógica Padcaya (no. 6628), Dpto. de Tarija. Informe interno de Cusco (13 – 16S, Pérou). Thèse de doctorat, Université de
GEOBOL, La Paz, 27 pp. Grenoble, France, 159 pp.
Atherton, M.P., 1990. The coastal batholith of Peru: the product of Carlotto, V., Carlier, G., Jaillard, E., Sempere, T., Mascle, G., 1999.
rapid recycling of ‘‘new’’ crust formed within rifted continental Sedimentary and structural evolution of the Eocene – Oligocene
margin. Geological Journal 25, 337 – 349. Capas Rojas basin: evidence for a late Eocene lithospheric
Atherton, M.P., Webb, S., 1989. Volcanic facies, structure, and geo- delamination event in the southern Peruvian Altiplano. IV In-
chemistry of the marginal basin rocks of central Peru. Journal of ternational Symposium on Andean Geodynamics, Göttingen,
South American Earth Sciences 2, 241 – 261. pp. 141 – 146.
Aumaı̂tre, R., Grandin, G., Guillon, J.H., 1977. Données lithologi- Carlotto, V., Cárdenas, J., Dı́az-Martı́nez, E., Sempere, T., Hermoza,
ques et structurales relatives à un bloc précambrien surélevé de W., Cerpa, L., Acosta, H., 2000. La Formación Ene de la región
la Cordillère andine orientale (Pérou central). Les corps de de Cusco y su importancia en la exploración de yacimientos de
roches ultrabasiques qui y sont présents. Bulletin de la Société hidrocarburos. X Congreso Peruano de Geologı́a, Lima, CD-
Géologique de France 19, 983 – 989. ROM file GH1.
Barbieri, M., Ghiara, M.R., Stanzione, D., Villar, L.M., Pezzutti, Cenki, B., 1998. Le volcanisme permo-triasique et/ou mésozoı̈que
N.E., Segal, S.J., 1997. Trace-element and isotope constraints on de la région de Cusco-Sicuani: contexte géologique, caractéris-
the origin of ultramafic lamprophyres from Los Alisos (Sierras tiques pétrographiques, minéralogiques et géochimiques, inter-
Subandinas, northern Argentina). Journal of South American prétation géodynamique. Mémoire de maı̂trise, Université de
Earth Sciences 10, 39 – 47. Grenoble, Francia, 33 pp.
Bard, J.-P., Botello, R., Martinez, C., Subieta, T., 1974. Relations Chandler, M.A., Rind, D., Ruedy, R., 1992. Pangean climate during
entre tectonique, métamorphisme et mise en place d’un granite the Early Jurassic: GCM simulations and the sedimentary record
éohercynien à deux micas dans la Cordillère Real de Bolivie of paleoclimate. Geological Society of America Bulletin 104,
(massif de Zongo-Yani). Cahiers ORSTOM, Série Géologie 6, 543 – 569.
3 – 18. Clark, A.H., Farrar, E., Kontak, D.J., Langridge, R.J., Arenas, M.J.,
Beck, S., Zandt, G., Myers, S.C., Wallace, T.C., Silver, P.G., Drake, France, L.J., McBride, S.L., Woodman, P.L., Wasteneys, H.A.,
L., 1996. Crustal-thickness variations in the central Andes. Sandeman, H.A., Douglas, D.A., 1990a. Geologic and geochro-
Geology 24, 407 – 410. nologic constraints on the metallogenic evolution of the Andes
Bellido, E., Guevara, C., 1963. Geologı́a de los Cuadrángulos de of southeastern Peru. Economic Geology 85, 1520 – 1583.
Punta Bombón y Clemesi. Carta Geológica Nacional, Lima, 92 pp. Clark, A.H., Kontak, D.J., Farrar, E., 1990b. The San Judas Tadeo
Beltan, L., Freneix, S., Janvier, P., López-Paulsen, O., 1987. La W ( – Mo, Au) deposit: Permian lithophile mineralization in
faune triasique de la formation de Vitiacua dans la région de southeastern Peru. Economic Geology 85, 1651 – 1668.
Villamontes (Département de Chuquisaca, Bolivie). Neues Jahr- Dalmayrac, B., Laubacher, G., Marocco, R., 1980. Caractères gén-
buch fuer Geologie und Palaeontologie, Monatshefte, 99 – 115. éraux de l’évolution géologique des Andes péruviennes. Trav-
Benavides, V., 1962. Estratigrafı́a pre-terciaria de la región de Are- aux et Documents de l’ORSTOM, Paris 122, 501 pp.
quipa. Boletı́n de la Sociedad Geológica del Perú 38, 5 – 63. Damiani-Pinto, I., Sanguinetti, Y.T., 1987. Lower Cretaceous ostrac-
Bogdanic, T., 1990. Kontinentale Sedimentation der Kreide und des odes from Bolivia. Anais do X Congresso Brasileiro de Paleon-
Alttertiärs im Umfeld des subduktionsbedingten Magmatismus tologia 2, 761 – 781.
178 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

Douglas, J.A., 1920. Geological sections through the Andes of Peru livia. Revista de la Universidade de Guarulhos-Geociências II
and Bolivia: 2. From the port of Mollendo to the Inambari River. (no. especial), 225.
Geological Society of London Quarterly Journal 76, 58 pp. Isacks, B.L., 1988. Uplift of the central Andean plateau and bending
Egeler, C.G., De Booy, T., 1961. Preliminary note on the geology of of the Bolivian orocline. Journal of Geophysical Research, B4
the Cordillera Vilcabamba (SE Peru), with emphasis on the 93, 3211 – 3231.
essentially pre-Andean origin of the structure. Geologie en Jacay, J., 1996. Geologı́a del cuadrángulo de Singa. Boletı́n IN-
Mijnbouw 40, 319 – 325. GEMMET, Serie A 67, 214 pp.
Enay, R., Barale, G., Jacay, J., Jaillard, E., 1996. Upper Tithonian Jacay, J., Sempere, T., Carlier, G., Carlotto, V., 1999. Late Paleo-
ammonites and floras from the Chicama basin, northern Peru- zoic – Early Mesozoic plutonism and related rifting in the East-
vian Andes. GeoResearch Forum, Transtec, Switzerland 1 – 2, ern Cordillera of Peru. IV International Symposium on Andean
221 – 234. Geodynamics, Göttingen, pp. 358 – 363.
Encarnación, J., Fleming, T.H., Elliot, D.H., Eales, H.V., 1996. Jaillard, E., 1990. Mesozoic extension and crustal thickening in the
Synchronous emplacement of Ferrar and Karoo dolerites and Peruvian Andes. I International Symposium on Andean Geo-
the early breakup of Gondwana. Geology 24, 535 – 538. dynamics, Grenoble, pp. 269 – 272.
Farrar, E., Clark, A.H., Heinrich, S.M., 1990. The age of the Zongo Jaillard, E., 1994. Kimmeridgian to Paleocene tectonic and geody-
pluton and the tectonothermal evolution of the Zongo – San Ga- namic evolution of the Peruvian (and Ecuadorian) margin. In:
bán Zone in the Cordillera Real, Bolivia. I International Sym- Salfity, J.A. (Ed.), Cretaceous Tectonics of the Andes. Vieweg,
posium on Andean Geodynamics, Grenoble, 171 – 174. pp. 101 – 167.
Fleming, T.H., Heimann, A., Foland, K.A., Elliot, D.H., 1997. Ar – Jaillard, E., Jacay, J., 1989. Les ‘‘Couches Chicama‘‘ du nord du
Ar geochronology of Ferrar Dolerite sills from the Transantarc- Pérou: colmatage d’un bassin né d’une collision oblique au
tic Mountains, Antarctica: implications for the age and origin of Tithonique. Comptes Rendus de l’Académie des Sciences de
the Ferrar magmatic province. Geological Society of America Parı́s, Série II 308, 1459 – 1465.
Bulletin 5, 533 – 546. Jaillard, E., Santander, G., 1992. La tectónica polifásica en escamas
Francßa, A.B., et al., 1995. Phanerozoic correlation in southern South de la zona de Mañazo – Lagunillas (Puno, sur del Perú). Bulletin
America. In: Tankard, A.J., Suárez, R., Welsink, H.J. (Eds.), de l’Institut Francßais d’Etudes Andines, Lima 21 (1), 37 – 58.
Petroleum Basins of South America. AAPG Memoirs, vol. 62, Jaillard, E., Soler, P., 1996. Cretaceous to early Paleogene tectonic
pp. 129 – 161. evolution of the northern Central Andes (0 – 18S) and its rela-
Franz, G., Lucassen, F., 1997. Upper Paleozoic crustal thickening: tions to geodynamics. Tectonophysics 259, 41 – 53.
the basement of the Sierra de Limón Verde in N-Chile (Región Jaillard, E., Bulot, L.G., Robert, E., Dhondt, A., Villagómez, R.,
Antofagasta). Actas del VIII Congreso Geológico Chileno 2, Rivadeneira, M., Paz, M., 1997. La transgresión del Cretácico
1271 – 1274. inferior en el margen andino (Perú y Ecuador). IX Congreso
Grandin, G., Zegarra-Navarro, J., 1979. Las rocas ultrabásicas en el Peruano de Geologı́a, 331 – 335.
Perú: las intrusiones lenticulares y los silles de la región de Janjou, D., Bourgois, J., Mégard, F., Sornay, J., 1981. Rapports
Huanuco-Monzón. Boletı́n de la Sociedad Geológica del Perú paléogéographiques et structuraux entre les cordillères occiden-
63, 99 – 115. tales et orientales des Andes nord-péruviennes: les écailles du
Günther, A., Haschke, M., Reutter, K.-J., Scheuber, E., 1997. Re- Marañón (7S, départements de Cajamarca et Amazonas, Pér-
peated reactivation of an ancient fault zone under changing ou). Bulletin de la Société Géologique de France (7) 32, 697 –
kinematic conditions: the Sierra-de-Moreno fault system 705.
(SMFS) (N-Chilean Precordillera). Actas del VIII Congreso Jardiné, S., 1974. Microflores des formations du Gabon attribuées
Geológico Chileno 1, 85 – 89. au Karoo. Revue de Paléobotanique et Palynologie 17, 75 – 112.
Hardenbol, J., Thierry, J., Farley, M.B., Jacquin, T., de Graciansky, Jenks, W., 1948. Geologı́a de la hoja de Arequipa, al 1/200.000.
P.-C., Vail, P.R., 1998. Mesozoic and Cenozoic sequence chro- Boletı́n del Instituto Geológico del Perú 9, 104 pp.
nostratigraphic framework of European basins, chart 1. In: de Jenks, W.F., 1951. Triassic stratigraphy near Cerro de Pasco, Peru.
Graciansky, P.-C., Hardenbol, J., Jacquin, T., Vail, P.R. (Eds.), Geological Society of America Bulletin 62, 203 – 220.
Mesozoic and Cenozoic Sequence Stratigraphy of European Kay, S.M., Coira, B., Viramonte, J., 1994. Young mafic back arc
Basins SEPM Special Publication 60. volcanic rocks as indicators of continental lithospheric delami-
Harrison, J.V., 1943. The geology of the Central Andes in part of the nation beneath the Argentine Puna Plateau, central Andes. Jour-
province of Junı́n, Peru. Quarterly Journal of the Geological nal of Geophysical Research 99, 24323 – 24339.
Society of London 99, 1 – 36. Kennan, L., Lamb, S., Rundle, C., 1995. K – Ar dates from the
Harrison, J.V., 1951. Geologı́a de los Andes Orientales del Perú Altiplano and Cordillera Oriental of Bolivia: implications for
central. Boletı́n de la Sociedad Geológica del Perú, vol. 21, 97 pp. Cenozoic stratigraphy and tectonics. Journal of South American
Hérail, G., Oller, J., Baby, P., Bonhomme, M., Soler, P., 1996. Earth Sciences 8, 163 – 186.
Strike-slip faulting, thrusting and related basins in the Cenozoic Klinck, B.A., Allison, R.A., Hawkins, M.P., 1986. The geology of
evolution of the southern branch of the Bolivian Orocline. Tec- the Cordillera Occidental and Altiplano west of Lake Titicaca,
tonophysics 259, 201 – 212. southern Peru. British Geological Survey, Nottingham, and IN-
Iannuzzi, R., Guerra-Sommer, M., Dı́az-Martı́nez, E., Grader, G.W., GEMMET, Lima, 353 pp.
1997. Presence of the Late Permian genus Glossopteris in Bo- Kontak, D.J., 1984. The magmatic arc and metallogenetic evolution
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 179

of a craton – orogen interface: the Cordillera de Carabaya, Cen- Mégard, F., 1978. Etude géologique des Andes du Pérou central.
tral Andes, Southeast Peru. PhD Dissertation, Queen’s Univer- Travaux et Documents de l’ORSTOM, Parı́s 86, 310 pp.
sity, Kingston, Ontario, 631 pp. Mégard, F., 1987. Cordilleran and marginal Andes: a review of An-
Kontak, D.J., Clark, A.H., Farrar, E., Strong, D.F., 1985. The rift dean geology north of the Arica elbow (18S). In: Monger, J.W.H.,
associated Permo-Triassic magmatism of the Eastern Cordillera: Francheteau, J. (Eds.), Circum-Pacific belts and evolution of the
a precursor to the Andean orogeny. In: Pitcher, W.S., Atherton, Pacific ocean basin. American Geophysical Union, Geodynamic
M.P., Cobbing, J., and Beckinsale R.D. (Eds.), Magmatism at a series 18, 71.
plate edge: The Peruvian Andes. Blackie, Glasgow, and Halsted Mégard, F., Marocco, R., Vicente, J.-C., Mégard-Galli, J., 1983.
Press, New York, pp. 36 – 44. Découverte d’une discordance angulaire tardi-hercynienne (Per-
Kontak, D.J., Clark, A.H., Farrar, E., Archibald, D.A., Baadsgaard, mien moyen) dans les Andes du Pérou central. Comptes Rendus
H., 1990. Late Paleozoic – Early Mesozoic magmatism in the de l’Académie des Sciences de Paris 296, 1267 – 1270.
Cordillera de Carabaya, Puno, southeastern Peru: geochronol- Mégard, F., Caldas, J., Paredes, J., De La Cruz, N., 1996. Geologı́a
ogy and petrochemistry. Journal of South American Earth Sci- de los cuadrángulos de Tarma, La Oroya y Yauyos. Boletı́n
ences 3, 213 – 230. INGEMMET, Serie A 69, 279 pp.
Kozlowski, R., 1934. Esquisse de la répartition des roches éruptives Méndez, V., Villar, L.M., 1979. Los filones ultrabásicos del Rı́o
dans les Andes de Bolivie. Archiwom Mineralogiczne Towarzyst- Piedras, Sierras Subandinas de Salta y Jujuy. Actas VII Con-
wa Naukowego Warszawskiego (Archives de Minéralogie de la greso Geológico Argentino 6, 119 – 129.
Société des Sciences et des Lettres de Varsovie) 10, 123 – 162. Menegatti, N., Omarini, R., Del Moro, A., Mazzuoli, R., 1997. El
Ladino, M., Tomlinson, A., Blanco, N., 1999. New constraints for the granito alcalino de la Sierra de Rangel (Crétacico inferior), Pro-
age of Cretaceous compressional deformation in the Andes of vincia de Salta, Argentina. Actas VIII Congreso Geológico Chi-
northern Chile (Sierra de Moreno, 21 – 22100S). IV Internation- leno, Antofagasta 2, 1379 – 1384.
al Symposium on Andean Geodynamics, Göttingen, 407 – 410. Meyer, H.A.O., Villar, L.M., 1984. An alnoite in the Sierras Sub-
Lamb, S., Hoke, L., Kennan, L., Dewey, J., 1997. Cenozoic evolu- andinas, northern Argentina. Journal of Geology 92, 741 – 751.
tion of the Central Andes in Bolivia and northern Chile. In: Mourier, T., 1988. La transition entre Andes marginales et Andes
Burg, J.P., Ford, M. (Eds.), Orogeny Through Time. Geological cordilléraines à ophiolites: évolutions sédimentaire, magmatique
Society Special Publication, vol. 121, pp. 237 – 264. et structurale du relais de Huancabamba (3 – 8S, nord-Pérou,
Lancelot, J.R., Laubacher, G., Marocco, R., Renaud, U., 1978. U/Pb sud-Equateur). Doctoral thesis, Université Paris XI, 302 pp.
radiochronology of two granitic plutons from the Eastern Cor- Mpodozis, C., Cornejo, P., 1997. El rift triásico-sinemuriano de
dillera (Peru): extent of Permian magmatic activity and conse- Sierra Exploradora, Cordillera de Domeyko (25 – 26S): aso-
quences. Geologische Rundschau 67, 236 – 243. ciaciones de facies y reconstrucción tectónica. Actas del VIII
Laubacher, G., 1978. Géologie de la Cordillère Orientale et de Congreso Geológico Chileno 1, 550 – 554.
l’Altiplano au nord et nord-ouest du lac Titicaca (Pérou). Trav- Mukasa, S.B., 1986. Zircon U – Pb ages of superunits in the Coastal
aux et Documents de l’ORSTOM 95, 217 pp. Batholith of Peru: implications for magmatic and tectonic pro-
Laurent, H., 1985. El pre-Cretáceo en el Oriente peruano: su dis- cesses. Geological Society of America Bulletin 97, 241 – 254.
tribución y sus rasgos estructurales. Boletı́n de la Sociedad Geo- Muñoz, N., Charrier, R., 1993. Jurassic – Early Cretaceous facies
lógica del Perú 74, 33 – 59. distribution in the western Altiplano (18 – 21300S): implica-
López-Murillo, R.D., López-Pugliessi, J.M., 1995. Estratigrafı́a del tions for hydrocarbon exploration. II International Symposium
Grupo Tacurú de las Sierras Subandinas. Revista Técnica de on Andean Geodynamics, Oxford, 307 – 310.
YPFB 16, 27 – 36. Muñoz, N., Elgueta, S., Harambour, S., 1988. El sistema jurásico
López-Pugliessi, J.M., 1995. Grupo Tacurú de las Sierras Subandi- (Fm. Livilcar) en el curso superior de la Quebrada de Azapa, I-
nas de Bolivia: nominación de las unidades formacionales que Región: implicancias paleogeográficas. Actas V Congreso Geo-
lo integran. Revista Técnica de YPFB 16, 55 – 68. lógico Chileno, Santiago, vol. 1, pp. A403 – A415.
Loughman, D.L., Hallam, A., 1982. A facies analysis of the Pucará Myers, S.C., Beck, S., Zandt, G., Wallace, T., 1998. Lithospheric-
group (Norian to Toarcian carbonates, organic-rich shales and scale structure across the Bolivian Andes from tomographic
phosphates) of Central and Northern Peru. Sedimentary Geol- images of velocity and attenuation por P and S waves. Journal
ogy 32, 161 – 194. of Geophysical Research 103 (B9), 21233 – 21252.
Mathalone, J.M.P., Montoya, M., 1995. Petroleum geology of the Newell, N.D., 1949. Geology of the Lake Titicaca region, Peru and
sub-Andean basins of Peru. In: Tankard, A.J., Suárez, R., Wel- Bolivia. Geological Society of America Memoir 36, 111 pp.
sink, H.J. (Eds.), Petroleum Basins of South America. American Newell, N.D., Chronic, J., Roberts, T., 1953. Upper Paleozoic of
Association of Petroleum Geologists Memoirs, vol. 62, pp. Peru. Geological Society of America Memoir 58, 276 pp.
423 – 444. Noble, D.C., Silberman, M.L., Mégard, F., Bowman, H.R., 1978.
McBride, S.L., Robertson, R.C.R., Clark, A.M., Farrar, E., 1983. Comendite (peralkaline rhyolites) in the Mitu Group, central
Magmatic and metallogenetic episodes in the northern tin belt, Peru: evidence of Permian – Triassic crustal extension in the
Cordillera Real, Bolivia. Geologische Rundschau 72, 685 – 713. Central Andes. U.S. Geological Survey Journal of Research 6,
McLaughlin, D.H., 1924. Geology and physiography of the Peru- 453 – 457.
vian Cordillera, departments of Junı́n and Lima. Geological Oller, J., Sempere, T., 1990. A fluvio-eolian sequence of probable
Society of America Bulletin 35, 591 – 632. middle Triassic – Jurassic age in both Andean and Subandean
180 T. Sempere et al. / Tectonophysics 345 (2002) 153–181

Bolivia. I International Symposium on Andean Geodynamics, and Bolivia: tectonic implications. Actas VIII Congreso Geo-
237 – 240. lógico Chileno, Antofagasta 3, 1719 – 1723.
Oviedo, C., 1962. Contribución al conocimiento estratigráfico de la Rüegg, W., 1956. Geologie zwischen Cañete und San Juan, 13000S –
penı́nsula de Copacabana. Tesis de Grado de la Universidad 15240S, Süd-Peru. Geologische Rundschau 45, 775 – 858.
Mayor de San Andrés, La Paz, 49 pp. Rüegg, W., 1961. Hallazgo y posición estratigráfico-tectónica del
Oviedo, C., 1964. Estratigrafı́a de la penı́nsula de Copacabana. In- Titoniano en la costa del sur del Perú. Boletı́n de la Sociedad
forme inédito GXG-YPFB no. 895, Santa Cruz. Geológica del Perú 36, 203 – 208.
Pádula, L.E., Reyes, F.C., 1958. Contribución al léxico estratigráfi- Saavedra, A., Santivañez, R., Shimada, N., 1986. Edades radiométr-
co de las Sierras Subandinas, República de Bolivia. Revista icas de Bolivia. Publicación especial del IGE (UMSA)-JICA, La
Técnica de YPFB 1 (1), 9 – 70 (and reimpresión 1960, v. 2 Paz.
(9), pp. 9 – 70). Sager, W.W., Koppers, A.A.P., 2000. Late Cretaceous polar wander
Ponce de León, V., Mariaca, J., Hochstatter, H., Llanos, R., Vargas, of the Pacific plate: evidence of a rapid true polar wander event.
C., 1972. Informe geológico regional de la Faja Subandina del Science 287, 455 – 459.
Norte (sector noroccidental). Informe interno GXG-YPFB no. Schmitz, M., et al., 1999. The crustal structure beneath the Central
1725, Santa Cruz. Andean forearc and magmatic arc as derived from seismic stud-
Portugal, J., 1974. Mesozoic and Cenozoic stratigraphy and tectonic ies — the PISCO 94 experiment in northern Chile. Journal of
events of Puno – Santa Lucı́a area, Department of Puno, Peru. South American Earth Sciences 12, 237 – 260.
American Association of Petroleum Geologists Bulletin 58, Sempere, T., 1995. Phanerozoic evolution of Bolivia and adjacent
982 – 999. regions. In: Tankard, A.J., Suárez-Soruco, R., Welsink, H.J.
Quispesivana, L., 1996. Geologı́a del cuadrángulo de Huanuco. (Eds.), Petroleum Basins of South America. AAPG Memoir,
Boletı́n de INGEMMET, Serie A 75, 138 pp. vol. 62, pp. 207 – 230.
Ramı́rez, C.F., Huete, C., 1981. Geologı́a de la hoja Ollagüe, Re- Sempere, T., 2000. Discussion of ‘‘Sediment accumulation on top of
gión de Antofagasta. Carta Geológica de Chile, no. 40, Instituto the Andean orogenic wedge: Oligocene to late Miocene basins
de Investigaciones Geológicas, Santiago. of the eastern Cordillera, southern Bolivia’’ (Horton, 1998).
Rivadeneira, M.V., Baby, P., 1999. La cuenca Oriente: estilo tectó- Geological Society of America Bulletin, in press.
nico, etapas de deformación y caracterı́sticas geológicas de los Sempere, T., Hérail, G., Oller, J., Bonhomme, M.G., 1990. Late
principales campos de petroproducción Publicación Petropro- Oligocene – early Miocene major tectonic crisis and related ba-
ducción-IRD, 88 pp. sins in Bolivia. Geology 18, 946 – 949.
Rivas, S., 1989. La Cordillera Occidental, ventana para el conoci- Sempere, T., Aguilera, E., Doubinger, J., Janvier, P., Lobo, J., Oller,
miento geológico continental. VIII Congreso Geológico Boli- S., Wenz, S., 1992. La Formation de Vitiacua (Permien moyen à
viano, La Paz, 205 – 225. supérieur-Trias? inférieur, Bolivie du Sud): stratigraphie, paly-
Rochat, P., Hérail, G., Baby, P., Mascle, G., 1999. Bilan crustal et nologie et paléontologie. Neues Jahrbuch fuer Geologie und
contrôle de la dynamique érosive et sédimentaire sur les méca- Palaeontologie, Abhandlungen 185, 239 – 253.
nismes de formation de l’Altiplano. Comptes Rendus de l’Aca- Sempere, T., Carlier, G., Carlotto, V., Jacay, J., 1998. Rifting Pér-
démie des Sciences de Paris, Serie IIa: Sciences de la Terre et mico superior-Jurásico medio en la Cordillera Oriental de Perú y
des Planètes 328, 189 – 195. Bolivia. Memorias XIII Congreso Geológico Boliviano, Potosı́
Romeuf, N., Aguirre, L., Carlier, G., Soler, P., Bonhomme, M., 1, 31 – 38.
Elmi, S., Salas, G., 1993. Present knowledge of the Jurassic Sempere, T., Carlier, G., Carlotto, V., Jacay, J., Jiménez, N., Rosas, S.,
volcanogenic formations of southern coastal Peru. II Interna- Soler, P., Cárdenas, J., Boudesseul, N., 1999. Late Permian –
tional Symposium on Andean Geodynamics, Oxford, 437 – 440. Early Mesozoic rifts in Peru and Bolivia, and their bearing on
Romeuf, N., Aguirre, L., Soler, P., Féraud, G., Jaillard, E., Ruffet, Andean-age tectonics. IV International Symposium on Andean
G., 1995. Middle Jurassic volcanism in the Northern and Central Geodynamics, Göttingen, pp. 680 – 685.
Andes. Revista Geológica de Chile 22, 245 – 259. Sempere, T., Acosta, H., Carlotto, V., 2000a. Estratigrafia del Mes-
Roperch, P., Carlier, G., 1992. Paleomagnetism of Mesozoic rocks ozoico y Paleogeno en la región del Lago Titicaca: hacia una
from the Central Andes of southern Peru: importance of rota- solución? X Congreso Peruano de Geologı́a, Lima, CD-ROM
tions in the development of the Bolivian Orocline. Journal of file GR50, 41 pp.
Geophysical Research 97 (B12), 17233 – 17249. Sempere, T., Carlier, G., Fornari, M., 2000b. The Altiplano plateau,
Rosas, S., Fontboté, L., 1995. Evolución sedimentológica del Grupo Bolivia and southern Peru: constraints and insights from the
Pucará (Triásico superior – Jurásico inferior) en un perfil SW – geology of the Bolivian Orocline. AGU Fall Meeting, San Fram-
NE en el centro del Perú. Sociedad Geológica del Perú, vol. cisco, supplement to Eos. Trans. Am. Geo. Un. 81 (48), F1118.
jubilar A. Benavides, pp. 279 – 309. Skarmeta, J., Marinovic, N., 1981. Geologı́a de la hoja Quillagua,
Rosas, S., Fontboté, L., Morche, W., 1997. Vulcanismo de tipo Región de Antofagasta. Carta Geológica de Chile, no. 51, In-
intraplaca en los carbonatos del Grupo Pucará (Triásico superi- stituto de Investigaciones Geológicas, Santiago de Chile.
or – Jurásico inferior, Perú central) y su relación con el vulcan- Smulikowski, K., 1934. Les roches éruptives des Andes de Bolivie.
ismo del Grupo Mitu (Pérmico superior – Triásico). IX Congreso Archiwom Mineralogiczne Towarzystwa Naukowego Warszaw-
Peruano de Geologı́a, 393 – 396. skiego (Archives de Minéralogie de la Société des Sciences et
Rubiolo, D., 1997. Alkaline rocks in Central Andes from Argentina des Lettres de Varsovie) 10, 163 – 234.
T. Sempere et al. / Tectonophysics 345 (2002) 153–181 181

Soler, P., 1991. Contribution à l’étude du magmatisme associé aux Tawackoli, S., Jacobshagen, V., Wemmer, K., Andriessen, P.M.,
zones de subduction. Pétrographie, géochimie et géochimie iso- 1996. The Eastern Cordillera of southern Bolivia: a key region
topique des roches intrusives sur un transect des Andes du Pérou to the Andean back-arc uplift and deformation history. Extended
central. Implications géodynamiques et métallogéniques. Thèse Abstracts, III International Symposium on Andean Geodynam-
de doctorat d’Etat, Université Pierre-et-Marie-Curie (Paris VI), ics, Saint-Malo, France, pp. 505 – 508.
950 pp. Tawackoli, S., Rössling, R., Lehmann, B., Schultz, F., Claure-Za-
Soler, P., Bonhomme, M., 1987. Données radiochronologiques K/ pata, M., Balderrama, B., 1999. Mesozoic magmatism in Boli-
Ar sur les granitoı̈des de la Cordillère Orientale des Andes du via and its significance for the evolution of the Bolivian
Pérou Central. Implications tectoniques. Comptes Rendus de Orocline. Extended Abstracts, IV International Symposium on
l’Académie des Sciences de Paris, Série II 304, 841 – 845. Andean Geodynamics, Göttingen, Germany, pp. 733 – 740.
Soler, P., Sempere, T., 1993. Stratigraphie, géochimie et significa- Thierry, J., Clavel, B., Hantzpergue, P., Néraudeau, D., Rigollet, L.,
tion paléotectonique des roches volcaniques basiques mésozoı̈- Vadet, A., 1997. Distribution chronologique et géographique
ques des Andes boliviennes. Comptes Rendus de l’Académie des échinides jurassiques en France: essai d’utilisation biostrati-
des Sciences de Parı́s, Série II 316, 777 – 784. graphique. In: Cariou, E., Hantzpergue, P. (Eds.), Biostratigra-
Sprechmann, P., Bossi, J., da Silva, J., 1981. Cuencas del Jurásico y phie du Jurassique Ouest-Européen et Méditerranéen. Mémoires
Cretácico del Uruguay. In: Volkheimer, W., Musacchio, E.A. Elf-Aquitaine, vol. 17, pp. 253 – 271.
(Eds.), Cuencas Sedimentarias del Jurásico y Cretácico de Vicente, J.-C., 1981. Elementos de la estratigrafı́a mesozoica sur-
América del Sur. Comité Sudamericano del Jurásico y Cretáci- peruana. In: Volkheimer, W., Musacchio, E.A. (Eds.), Cuencas
co, Buenos Aires, vol. 1, pp. 239 – 270. Sedimentarias del Jurásico y Cretácico de América del Sur, vol.
Stanley, G.D., 1994. Paleontology and stratigraphy of Triassic to 1. Comité Sudamericano del Jurásico y Cretácico, Buenos Aires,
Jurassic rocks in the Peruvian Andes. Paleontolographica, Abteil pp. 319 – 351.
A 233, 208 pp. Vicente, J.-C., 1989. Early Late Cretaceous overthrusting in the
Steinmann, G., 1906. Die Entstehung der Kupferlagerstätten von Western Cordillera of southern Peru. In: Ericksen, G.E., Cañas,
Corocoro und verwandten Vorkomnisse in Bolivia. Rose- M.T., Reinemund, J.A. (Eds.), Geology of the Andes and its
nbusch-Festschrift, Stuttgart. Relation to Hydrocarbon and Mineral Resources, Houston,
Steinmann, G., 1923. Umfang, Beziehungen und Besonderheiten vol. 11, pp. 91 – 117.
der Andinen Geosynklinale. Geologische Rundschau 14. Vicente, J.-C., Beaudoin, B., Chávez, A., León, I., 1982. La cuenca
Steinmann, G., 1929. Geologie von Peru. Karl Winter, Heidelberg, de Arequipa (Sur Perú) durante el Jurásico – Cretácico inferior.
448 pp. V Congreso Latinoamericano de Geologı́a 1, 121 – 153.
Stewart, J.W., Evernden, J.F., Snelling, N.J., 1974. Age determina- Vieira, C.L., Iannuzzi, R., Dı́az-Martı́nez, E., Grader, G.W., 1999a.
tions from Andean Peru: a reconnaissance survey. Geological Contribuicßão ao registro de pecopterı́deas da Formacßão Chutani
Society of America Bulletin 85, 1107 – 1116. (Grupo Titicaca, Bolivia) e algumas consideracß ões sobre suas
Suárez-Riglos, M., Dalenz, A., 1993. Pteriomorphia (Bivalvia) nor- idades e paleoambiente. Abstracts, XVI Congresso Brasileiro de
iano de la Formación Vitiacua, del área de Villamontes (Tarija). Paleontologia, Crato, 124.
In: Suárez-Soruco, R. (Ed.), Fósiles y Facies de Bolivia, vol. II. Vieira, C.L., Iannuzzi, R., Dı́az-Martı́nez, E., Grader, G.W., 1999b.
Revista Técnica de YPFB, vol. 13 – 14, pp. 155 – 160. Presence of the morphogenus Pecopteris in Late Permian de-
Tankard, A.J., et al., 1995. Structural and tectonic controls of basin posits from western Bolivia (Chutani Formation, Titicaca
evolution in southwestern Gondwana during the Phanerozoic. Group). Anais da Academia Brasileira de Ciências 71 (4-I),
In: Tankard, A.J., Suárez, R., Welsink, H.J. (Eds.), Petroleum 809 – 810.
Basins of South America. AAPG Memoirs, vol. 62, pp. 5 – 52. Vivier, G., Audebaud, E., Vatin-Pérignon, N., 1976. Le magmatisme
Tasch, P., 1987. Fossil Conchostraca of the Southern Hemisphere tardi-hercynien et andin le long d’une transversale sud-péruvi-
and continental drift. Geological Society of America Memoir enne: bilan géochimique des éléments incompatibles. Réunion
165, 290 pp. Annuelle des Siences de la Terre, Paris, 396.
Tawackoli, S., 1999. Andine Entwicklung der Ostkordillere in der Yuan, X., et al., 2000. Subduction and collision processes in the
region Tupiza (Südbolivien). Berliner Geowissenschaftliche Ab- Central Andes constrained by converted seismic phases. Nature
handlungen, Reihe A 203, 116 pp. 408, 958 – 961.

You might also like