You are on page 1of 24

Phanerozoic Evolution of Bolivia and

Adjacent Regions

Thierry Sempere
Convenio YPFBOrstom
Santa Cruz, Bolivia

Abstract

T he geologic evolution of Bolivia and the central Andean system during the past 500 m.y. was largely
controlled by the geodynamics of the South American margin of western Gondwana. The Phanerozoic
strata were deposited in mainly marine environments until the Early Triassic, after which continental environ-
ments predominated. However, there were six restricted marine transgressions in the Late CretaceousDanian
and one in the late Miocene.
The Late CambrianEarly Ordovician margin was initially a passive margin. It became an active one during
a Middle Ordovician compressional episode and was controlled by large-scale transtensional or transpres-
sional conditions from the Late Ordovician to the Triassic. The Late OrdovicianMississippian evolution was
characterized by vigorous subsidence of the marine foreland, which was filled with thick, shallowing-upward
sequences showing northeastward onlaps. Ashgill and latest DevonianMississippian glaciomarine and fluctu-
ating sea level processes are recorded in the succession. Shallow marine carbonates, marls, and sandstones, as
well as some evaporites and eolianites, were deposited during PennsylvanianEarly Triassic time.
After Middle Triassic rifting was aborted, the Bolivian basin behaved in a cratonic way until it was caught
up in the Andean system due to the onset of transtension along the margin in the Late Jurassic. It became part
of the Andean foreland domain in early Senonian time. Andean thrust deformation propagated into Bolivia
from the west in the late Oligocene and progressed eastward through Neogene time.
Organic-rich units correlate with Paleozoic highstand deposits and younger transgressions. Generation,
migration, and trapping of hydrocarbons depended mainly on Cenozoic sedimentary burial and tectonic
loading and hence on propagation of Andean deformation.

Resumen

L a evolucin geolgica de Bolivia y de los Andes centrales durante los ltimos 500 Ma fue controlada en
gran parte por la geodinmica del margen sudamericano del Gondwana occidental. Los estratos fanero-
zoicos se depositaron en ambientes principalmente marinos hasta el Trisico inferior, y predominantemente
continentales posteriormente. Sin embargo, seis transgresiones marinas restringidas ocurrieron en el Cretcico
superior-Daniano, and una en el Mioceno superior.
El margen Cmbrico superior-Ordovcico inferior era inicialmente un margen pasivo. Se volvi activo
durante un episodio compresivo en el Ordovcico medio, y estuvo controlado por condiciones transtensionales
o transpresionales de gran escala del Ordovcico superior al Trisico. Una fuerte subsidencia del antepas
marino, rellenado por espesas secuencias de somerizacin que presentan un traslape hacia el noreste, carac-
teriz el perodo Ordovcico superior-Mississippiano. Procesos glaciomarinos (Ashgilliano y Devnico
terminal-Mississippiano) y fluctuaciones eustticas se registran en la sucesin. Carbonatos, margas, y areniscas,
de agua somera, y algunas evaporitas y eolianitas, se depositaron durante el intervalo Pennsylvaniano-Trisico
inferior.
Posteriormente al aborto del rifting del Trisico medio, la cuenca boliviana se comport de manera
cratnica, hasta su captura en el sistema andino por la iniciacin en el Jursico superior de condiciones
transtensivas a lo largo del margen. Se volvi parte del antepas andino en el Senoniano inferior. En el
Oligoceno superior, la deformacin andina empez a propagarse en Bolivia desde el oeste por medio de cabal-
gamientos, y progres hacia el este durante el Neogeno.
Las unidades ricas en materia orgnica corresponden a depsitos paleozoicos de alto nivel marino y a trans-
gresiones ms jvenes. La generacin, la migracin y el entrampamiento de los hidrocarburos han dependido
principalmente del soterramiento sedimentario y de la carga tectnica cenozoicos, es decir de la propagacin
de la deformacin andina.

Sempere, T., 1995, Phanerozoic evolution of Bolivia and adjacent regions, in A. J. Tankard, 207
R. Surez S., and H. J. Welsink, Petroleum basins of South America: AAPG Memoir 62,
p. 207230.
208 Sempere

Figure 1General structure of the


area of interest. Study area and
Bolivia-Peru basin are shown in
insert. Symbols: dotted lines,
drainage divides; finely dashed
lines, national boundaries;
hatchured, western Andes
(Sempere et al., 1989, 1990b, 1991);
crystalline pattern, Khenayani
Turuchipa paleostructural corridor
(CPKT) in the northwest, and
northeastern boundary of Chaco
Boreal high in the southeast.
Structural elements and domains
(most are Spanish): CALP, main
Altiplanic thrust; CANP, main
Andean thrust; CFP, main frontal
thrust; FA, Apurimac fault; FLIA,
intra-Andean boundary fault
(conjectural); SFK, Khenayani fault
system; CB, Chapare buttress; CI,
Cuzco indenter; CK, Calama knot;
MI, Mizque indenter; SB, Susques
buttress; VH, Vilcamba hinge;
FPCH, Huarina fold and thrust belt;
SAB, sub-Andean belt; SBS, Santa
Brbara system; TTB, Tarija-
Teoponte belt; UCU, Ulloma-
Coipasa-Uyuni domain. Cities: An,
Antofagasta; Ar, Arequipa; As,
Asuncin; Cb, Cochabamba; Ch,
Charagua; Cr, Corumb; P, Potos;
S, Salta; SC, Santa Cruz; T, Tarija.

INTRODUCTION young and only started to form in the late Oligocene


(Sempere et al., 1990a). In the Bolivian orocline, late
Bolivia is of particular interest because it possesses an OligoceneRecent deformation propagated toward the
almost complete Phanerozoic stratigraphy that has east and northeast through what had been the eastern
recorded the evolution of the nearby Pacific margin of part of the Andean depositional area. Consequently, the
western Gondwana. New data collected in the past 10 Bolivian Andes provide many excellent exposures of
years, combined with the use of modern methods in Phanerozoic strata, of better quality and easier access
stratigraphy and basin analysis, permit a better under- than their counterparts in the sub-Andean rain forests of
standing of its regional Phanerozoic geologic history. central and northern Peru. Andean Bolivia is an ideal
Bolivia was located on the edge of cratonic Gondwana place to study the stratigraphy and paleogeography of
during most of the Phanerozoic. This geotectonic location the margin of western Gondwana.
explains the high subsidence rates reflected in the great During most of Late OrdovicianPaleogene time,
thicknesses of many of the stratigraphic units and the Bolivia was located close to the southern tip of a trough
comparatively small amount of magmatism and tecton- that ran parallel to the Pacific margin of South America
ism prior to Neogene time. and that generally deepened toward the north or
The purpose of this paper is to summarize new and northwest along its axis. This trough, called the Bolivia-
mostly unpublished stratigraphic and paleogeographic Peru basin in this paper, was confined to the south by the
data concerning Bolivia, to outline the Phanerozoic Sierras Pampeanas area in northwestern Argentina
geologic evolution of the central Andean area, and to during this time interval. An important consequence of
suggest how these data could be useful to hydrocarbon this geometry is that, from Late Ordovician to early
exploration. Paleocene time, marine transgressions flooded into
Bolivia from the northwest.
Between about 450 and 145 Ma, most of Bolivia
STRUCTURAL SETTING consisted of a large sedimentary basin that was subjected
to only local compressional deformation. The Silurian
Three broad geomorphic provinces are recognized in Devonian to Middle Jurassic stratigraphy of Bolivia
Bolivia: the Andes mountains in the southwest, the closely resembles that of the widespread Chaco-Paran
Guapor shield in the northeast, and between them, the basin of southern Brazil, eastern Paraguay, northwestern
Beni-Chaco plain, which is the present-day foreland Uruguay, and northeastern Argentina (Sempere, 1990),
basin (Figure 1). The Bolivian Andes are comparatively suggesting that the Bolivian strata were deposited in a
Phanerozoic Evolution of Bolivia and Adjacent Regions 209

region that was relatively stable and cratonic in nature al., 1980) and the basal part of the Marcona Formation of
(Oller and Sempere, 1990). The present-day Beni-Chaco southern coastal Peru (Faro and San Juan members)
basin, into which Andean deformation is still propa- (Shackleton et al., 1979) apparently have similar litholo-
gating, is the undeformed remnant of the Bolivia-Peru gies and deformation styles.
basin. Deformed rocks of NeoproterozoicEocambrian age
The extensional processes that culminated in the in eastern Bolivia define a west-northwest striking
opening of the South Atlantic from the latest Jurassic wrench fault zone along the southern edge of the
onward dissected and diminished the old depocenter, Guapor shield. This fault zone splays in central Bolivia
the western part of which became involved in the Pacific and becomes an east-northeast trending fold and thrust
margin geotectonic system (Sempere, 1994). Little sedi- belt (Zubieta-Rossetti et al., 1993). This large-scale
mentation but considerable pedogenesis occurred in the structure is EarlyMiddle Cambrian in age; the Chapare
sub-Andean-Beni-Chaco basin between Late Jurassic and buttress (Hrail et al., 1990; Sempere et al., 1990a) is
early Oligocene time. The productive sub-Andean-Beni- related to it.
Chaco basin thus records two superimposed histories of The Phanerozoic chronostratigraphy of Bolivia is not
evolution spanning the early PaleozoicJurassic (pericra- well constrained and requires more biostratigraphic,
tonic Gondwana) and late OligoceneRecent (Andean) geochronologic, and sequence stratigraphic work. For
intervals. instance, the Devonian and Carboniferous biostratig-
Andean thrust deformation propagated into Bolivia raphy suffers from many discrepancies among authors
after 27 Ma by taking advantage of several dcollement and among fossil groups (e.g., see Racheboeuf et al.,
levels in Paleozoic shales of the Bolivia-Peru basin 1993, for the Devonian). The origin of some of these
(Sempere et al., 1989, 1990a). A close control on the discrepancies may partially lie in continental-scale
geometry of Andean deformation was imposed by the diachroneity of fossil groups.
geometry of the Paleozoic basin and its sedimentary fill Most Late CambrianEarly Triassic strata are of
(Sempere et al., 1989, 1991; Hrail et al., 1990). Propaga- marine origin, with many dark shale units recording sea
tion of thrust deformation and sedimentary and tectonic level highs. Middle TriassicRecent strata were mainly
burying of Paleozoic organic-rich units have also deposited in continental environments.
controlled hydrocarbon migration and distribution. The Bolivian Phanerozoic stratigraphy can be divided
Present-day geometries of Phanerozoic paleogeogra- into eight main supersequences (with approximate
phies suggest that Andean deformation was consider- durations given in parentheses) (modified from Sempere,
able in the Bolivian orocline, involving a higher short- 1990):
ening in its northwest-trending segment and trans-
current motions (Figure 2). The southern half of Andean 1. Tacsara supersequence: Late Cambrianmiddle
Bolivia is crossed by a major northeast-striking structural Caradoc (80 m.y. or more)
element known as the Khenayani-Turuchipa paleostruc- 2. Chuquisaca supersequence: late Caradocmiddle
tural corridor (CPKT) (Figure 1), which controlled Famennian (85 m.y.)
Phanerozoic sedimentation and deformation in south- 3. Villamontes supersequence: late Famennian
western Bolivia (Sempere et al., 1991). Paleogeographic Mississippian (40 m.y.)
data suggest that the northeastern edge of the Chaco 4. Cuevo supersequence: PennsylvanianEarly
Boreal high was the continuation of the CPKT before Triassic (85 m.y.)
Andean deformation (Figure 1). It is believed that offset 5. Serere supersequence: Middle TriassicMiddle
of this paleogeographic marker by Andean deformation Jurassic (95 m.y.)
offers an eloquent illustration of the dextral tectonic 6. Puca A+B supersequence: latest JurassicTuronian
displacement that occurred in the southern part of the (56 m.y.)
orocline (~500 km cumulative). Large amounts of trans- 7. Puca C supersequence: Senonianmiddle
current displacements in this region are attributed to Paleocene (31 m.y.)
rotation of the Vilcabamba hinge zone in southern Peru 8. Corocoro supersequence: latest PaleoceneRecent
at the northern end of the orocline (VH in Figure 1) (57 m.y. or more)
(Sempere et al., 1989).
Definition of these eight stratigraphic intervals is
based solely on the sedimentary record and not on
tectonic or magmatic information. The boundaries
STRATIGRAPHY between these supersequences are generally sharp. In the
Chaco basin, the boundary between the Villamontes and
In Andean Bolivia, the basement consists of Upper Cuevo supersequences is not as sharp as it is in north-
Proterozoic or Lower Cambrian rocks that crop out in a western Bolivia. It is difficult to separate the Cuevo and
few places. Deformed epimetamorphic rocks are known Serere intervals in western Bolivia and southern Peru
in the Tarija area, along the Chapare road northeast of because of magmatism and deformation during the
Cochabamba, and at Cerro Chilla south of Lake Titicaca. middle PermianEarly Jurassic. Grouping of the Villam-
These rocks show similarities to the Puncoviscana ontes, Cuevo, and Serere supersequences, which coincide
Formation of northwestern Argentina. The Yanahuanca with the formation and subsequent dispersal of Pangea,
Formation of the central Peruvian Andes (Dalmayrac et would thus be possible.
210 Sempere
Phanerozoic Evolution of Bolivia and Adjacent Regions 211

(facing page) In this paper, the terms supersequence, megasequence,


Figure 2Paleozoic paleogeography for various time and sequence, in order of decreasing thickness, are used
intervals. Geographic reference marks and Neogene for stratigraphic units bounded by sharp contacts.
tectonostratigraphic boundaries as in Figure 1. Hachures
show approximate extent of basin. Arrows are mean paleo-
currents; arrow tips are location of data. (A) Late
CambrianEarly Ordovician paleogeography. Depositional
facies are mainly turbidites around the Puna aulacogen LATE CAMBRIANMIDDLE CARADOC
and distal to proximal shelf sediments in other outcrop
areas. 1, Mainly mafic eruptive rocks; 2, high gravity Strata of Late Cambrianmiddle Caradoc age are
anomaly of Gtze et al. (1988). Paleocurrents outside exposed in the Cordillera Oriental, southern Altiplano,
Bolivia after Niemeyer et al. (1985), Moya (1988), and and some parts of the sub-Andean belt of Bolivia
Bahlburg (1990). (B) Llandeilomiddle Caradoc paleogeo- (Surez-Soruco, 1976). They are unknown in the Ulloma-
graphy. All known depositional facies are shallow marine. Coipasa-Uyuni domain (Figure 1). The oldest outcrop-
1, Caradoc mafic volcanics (Avila, 1992); 2, acid plutonism, ping rocks in the north are of late Proterozoic or Early
possibly of magmatic arc; 3, known area affected by Middle Cambrian age and are overlain by Upper Cretaceous
Ordovician deformation; 4, structural strikes. (C) Ashgill strata, whereas in central and southern Bolivia, the
Llandovery paleogeography. Facies are mainly resedi-
ments. Dotted line, eastern limit of Tokochi Formation; Ashgill age Cancairi Formation crops out as the basal
tight hachures, Cancairi Formation is >250 m thick; loose dcollement of the local thrusts. Avila (1992) has
hachures, Cancairi is <100 m thick; 1, Iron-rich rocks; 2, reviewed the coeval igneous rocks of Bolivia.
probable ongoing magmatism. Note influence of CPKT on The base of this supersequence is exposed in the Tarija
paleogeography. The Caradoc mafic volcanics of shown in area and along the Cochabamba-Chapare road. In most
(B) may have been erupted at beginning of this interval. of the outcropping area, the basal dcollement of the
Data from Paraguay after Dyck (1992). (D) Wenlockmiddle Andean thrusts is located in LowerMiddle Ordovician
Famennian paleogeography. Loose hachures, mainly shales and no older strata are exposed. These successions
shallow marine shale and sandstone (Bolivia-Peru basin);
tight hachures, DevonianEarly Mississippian turbidites, are generally thick and monotonous.
highly deformed in northwest-trending recumbent folds In the Tarija area, the sequence begins with shallow
during Mississippian (El Toco Formation, northern Chile); marine clastics, which grade upward to thick, open
1, Wenlock limestone; 2, PridoliLochkovian paleocur- marine graptolitic shales with subordinate turbidites and
rents; 3, latest Eifelianearly Givetian paleocurrents; slumps. Paleocurrent data and facies distributions from
4, Frasnianmiddle Famennian paleocurrents; 5, little southern Bolivia, northwestern Argentina, and northern
deformed shallow marine Middle Devonian sandstones Chile define a north to north-northeast trending trough
associated with Arequipa massif; 6, interstratified tuffs; (Figure 2A) (Sempere, 1989, 1991a; Bahlburg, 1990). In
7, San Nicols batholith; 8, erosional boundary of Devonian
in Bolivia. Successive edges of basin: A-L, AshgillLlan-
the southern part of this trough, ophiolitic rocks
dovery; W, Wenlock; L, earliest Lochkovian; E, Emsian. (Allmendinger et al., 1983) and a large positive gravity
Mafic sills, dikes, and flows occur in Devonian strata of the anomaly suggesting the presence of a dense substrate
Peruvian and northwestern Bolivian sub-Andean belt. Data (Gtze et al., 1988) are documented. This early Paleozoic
partially from Mgard (1978), Dalmayrac et al. (1980), trough was apparently underlain by a substantially
Breitkreuz (1986), ENAP (1988, personal communication), thinned crust.
Mukasa and Henry (1990), and YPFB (proprietary data). In southwestern Bolivia, the Upper Cambrian
(E) Mississippian paleogeography. 1, continental to Llanvirn sequence is affected by compressional deforma-
shallow marine facies; 2, slope and basinal deposits,
mainly resediments; 3, most distal facies; 4, Late Devonian
tion. In northernmost Argentina, beds of Ashgill and
or Mississippian folding; 5, structural strikes; U, uplifted Llandovery age clearly postdate this deformation
areas (CBU, Chaco Boreal high; MU, Madidi uplift; GCU, (Isaacson et al., 1976; Benedetto et al., 1992; M. C. Moya,
Guapor craton uplift). Although they varied with time, all 1992, personal communication). In the Tarija area, the
Mississippian mean paleocurrents are integrated here, Ashgill age Cancairi Formation directly overlies
including data from Helwig (1972) and Breitkreuz (1986). ArenigLlanvirn strata. This Ocloyic deformation
Note reorganization of paleocurrents and arches compared increases from east to west, and a strong cleavage
with previous period. Hercynian deformation is indicated appears in southwestern Bolivia. It therefore documents
only where it is reliably identified; it was confused with
a significant compressional event and uplift of
Late PermianEarly Triassic or Late Triassic deformation in
some areas. Paleogeography in eastern Peru is tentative; it LlandeiloCaradoc age (Sempere, 1989, 1990, 1991a).
probably included Mississippian uplifts, but deposits in However, this deformation is unknown north of 20 S lat
this area were locally eroded prior to the Cretaceous in Bolivia. The orientations of cleavages (Figure 2B)
(Mgard, 1978). (F) PennsylvanianEarly Permian paleo- suggest that the compressional fabric may continue
geography. Wavy line, erosional boundary of basin in sub- below the present-day volcanic Cordillera Occidental,
Andean-Beni-Chaco region; 1, shallow marine carbonates, that is, along an inferred northeastern boundary of the
Bolivia-Peru basin; 2, fluvioeolian sandstones; 3, limestone Arequipa massif. Although no direct evidence supports
intercalations in Cangapi Formation; 4, marine deposits this interpretation west of Lake Titicaca (Laubacher,
(carbonates) of northern Chile and southern coastal Peru
(after Breitkreuz, 1991); 5, marine or fluvial paleocurrents; 1978), the intensely deformed Llanvirn strata of central
6, progradation of eolianites; 7, axis of Bolivia-Peru basin; Peru (Mgard, 1978) may be a counterpart.
8, magmatism; 9, interstratified volcanics; 10, Late Penn- North of 20 S lat in Bolivia, the exposed succession
sylvanian intraarc lake of Breitkreuz (1991). consists of a thick sequence of LlanvirnCaradoc shal-
212 Sempere

Figure 3Reconstruction of late Proterozoic superconti-


nent (Bond et al., 1984; Murphy and Nance, 1991). AF,
Africa; APT, Appalachian-Patagonian terranes (Figure 6);
B, Baltica; NA, North America; SA, South America; thick
line, outline of cratonic Gondwana; dashed line, margin of
displaced terranes (Ramos, 1988a); circle, paleoequator;
central cross, late Proterozoic south pole.

lowing-upward sedimentary rocks, which is conform-


ably capped either by the upper Caradoc Tokochi
Formation (Sempere et al., 1991) or by the Ashgill Figure 4Map depicting present-day relationships
Cancairi Formation. Thus, deposition of thick shallow between known Ordovician deformation (black) and the
related basin (hachured). Deformation in northwestern
marine strata during LlandeiloCaradoc time coincided Argentina and southwestern Bolivia is due to closure of the
with the final development of the Ocloyic deformation in Puna aulacogen (Figures 2A, B). Black areas in the
southwestern Bolivia and northwestern Argentina northern Andes correspond to rocks deformed during this
(Figure 2B). This thick sequence of interbedded episode and again during a Permian event (after Dalmayrac
sandstone and siltstone is interpreted as the fill of a et al., 1980). Note offshore data at 9 S lat (Bourgois et al.,
marine foreland basin related to this deformation 1990) and probable deformation in coastal southern Peru
(Sempere, 1989, 1991a). (Shackleton et al., 1979). Map suggests current truncation
There is no consensus about the geodynamic system at ~12 S lat of Ordovician structures by trench and higher
that was active during this time, but the data are best Andean shortening in Bolivian orocline. eAd, edge of
explained if eastern North America and western South Andean deformation.
America were united in the late Proterozoic (Figure 3)
and only rifted apart in latest ProterozoicEarly Arenig-Llanvirn boundary according to data docu-
Cambrian time (Bond et al., 1984; Murphy and Nance, mented by Bahlburg (1990). The Atico or Marcona events
1991). The western margin of South America was thus a on the coast of southern Peru (Shackleton et al., 1979)
passive margin in Early CambrianEarly Ordovician probably correspond to this major deformation. A Cale-
time. The north-northeast trending, partially ophiolite- donian (i.e., Taconic) deformation is recorded from
floored trough in the Puna and southern Bolivian northwestern Peru up to northern Colombia (Dalmayrac
Altiplano is here interpreted as an aulacogen (the Puna et al., 1980). These data suggest that a major tectonic
aulacogen) which formed as a result of dextral-normal crisis affected the western margin of Gondwana during
strain along the southern Iapetus passive margin of the Middle Ordovician (Figure 4). However, interpreta-
Gondwana. The beginning of closure and deformation of tions of the early Paleozoic magmatic record of the
the Puna aulacogen, and thus the onset of active central Andes are contradictory (e.g., Breitkreuz et al.,
behavior along the nearby margin, occurred near the 1989; Damm et al., 1990; Rapela et al., 1992).
Phanerozoic Evolution of Bolivia and Adjacent Regions 213

LATE CARADOCMIDDLE FAMENNIAN


The late Caradocmiddle Famennian interval is repre-
sented by two sets of units that are believed to span the
late Caradocearly Llandovery and late Llandovery
middle Famennian time intervals, respectively (Figure 5).
The Devonian organic-rich shales are the most important
hydrocarbon source rocks in Bolivia.

Late CaradocEarly Llandovery


The first stratigraphic set includes the Tokochi,
Cancairi, and Llallagua formations, all of which reach
their maximum thickness in the Ulloma-Coipasa-Uyuni
domain. They thin out rapidly to the southeast and east
(Gagnier et al., in press).
The Tokochi Formation (Sempere et al., 1991) consists of
black shales up to 200 m thick that overlie shallow
marine Caradoc strata. It has yielded only an open
marine fauna, including cephalopods and Caradoc
pseudoplanktonic brachiopods (L. Branisa, 1986,
personal communication). The Tokochi is thought to
mark a sea level high (maximum flooding surface) and to
have been deposited in a relatively deep, anoxic environ-
ment. Because the thickness of these black shales
increases toward the west, they should be present at
depth in the central Altiplano.
The sharp or gradational contact between the Tokochi
and Cancairi formations indicates a drop in sea level.
The latter consists mainly of greenish gray, generally Figure 5Stratigraphic section for the late Caradoc
unstratified, resedimented sandy mudstones and diamic- middle Famennian interval. Lithologic symbols are
tites (mass flows) and minor sandstones (basinal common use.
turbidites and local shelf clastics). The Cancairi
Formation is of Ashgill age (Sempere, 1990; Benedetto et
al., 1992; Toro et al., 1992), but its top is possibly early the east and has yielded a late Llandovery graptolite
Llandovery. Thickness varies up to 1.5 km or more. The (Branisa et al., 1972).
millimeter to meter size subrounded to angular clasts The thickness of the Tokochi-Cancairi-Llallagua
consist of a variety of sedimentary, igneous, and meta- sequence increases dramatically in the eastern and
morphic rock types (Rodrigo et al., 1977). Some quartz southern Ulloma-Coipasa-Uyuni domain (Figure 2C),
and igneous rock clasts are clearly faceted and striated, and might have reached 3.5 km in depth before Andean-
indicating that they were derived from a glaciated area. age thrusting used the Cancairi Formation as the main
Large sandstone blocks, up to 1000 m3 in volume, are dcollement level there. These values and the age
more abundant near the Khenayani-Turuchipa paleo- constraints indicate maximum known decompacted
structural corridor, suggesting that structural activity sedimentation rates of about 250 m/m.y. for this time
during deposition of the Cancairi produced fault scarps span, which is very high. The predominance of sediment
and block sliding (Sempere et al., 1991). The Cancairi gravity flows in the Cancairi and Llallagua formations
Formation is interpreted as a lowstand deposit related to strongly suggests synsedimentary tectonic activity. It is
regional extension and associated high rates of resedi- possible that the Tokochi anoxic facies represent a
mentation while a large part of Gondwana was glaciated. starved basin stage in this evolution. Paleocurrents
The Llallagua Formation overlies the Cancairi (Figure 2C) indicate sediment transport toward the west
Formation with a sharp contact. It consists of turbidite and northwest, that is, in the direction of thickening and
beds, one to several meters thick, and subordinate toward the interior of the Ulloma-Coipasa-Uyuni
siltstone and mudstone; it can reach thicknesses of 1.5 domain.
km or more. These facies are attributed to lowstand These characteristics suggest that deposition of the
deposition following the Cancairi maximum lowstand. late Caradoc-Llandovery strata occurred in a fault-
The Llallagua Formation grades transitionally into the controlled basin, the principal normal faults being reacti-
Unca (Kirusillas) Formation, forming a thinning- and vated and inverted during the Andean orogeny to form
fining-upward succession as a result of sea level rise. The the Khenayani fault and Main Altiplanic thrust systems
Llallagua is attributed to the Llandovery because of its (Figure 2C) (Sempere et al., 1991). Resedimented facies
stratigraphic position. Its upper part is equated with the appear to have evolved from mass flows that reworked
basal Kirusillas (Unca) Formation, which occurs more to glaciomarine deposits (Cancairi Formation) to proximal
214 Sempere

(lower Llallagua Formation) and distal turbidites (upper


Llallagua Formation), suggesting an overall sea level rise.
This transition of the Llallagua Formation to the Unca
(Kirusillas) Formation also indicates waning of exten-
sional activity.

Late LlandoveryMiddle Famennian


The second stratigraphic set includes several units
that occur in the Bolivian Andes, sub-Andean belt, and
Beni-Chaco plains, albeit with different names. Shallow
marine, relatively warm water faunas were present as
early as Llandovery (Isaacson et al., 1976) and early
Wenlock time (Branisa et al., 1972; Merino, 1991; Gagnier
et al., in press). Although the precise ages of the Devonian
units are not firmly established, the chronostratigraphic
data of Vavrdov et al. (1991) and Racheboeuf et al. (1993)
are used here. Three overall thickening- and shallowing-
upward megasequences are recognized (Figure 5): late
LlandoveryLochkovian, Pragianearly Givetian, and late
Givetianmiddle Famennian. The first of these intervals is
generally divided into two lower order sequences: late
LlandoveryPridoli and Lochkovian (Figure 5). Paleoen-
vironments include relatively shallow marine, distal to
proximal shelf, deltaic, littoral, and locally even fluvial
deposits (e.g., in the Lochkovian age Santa Rosa
Formation).
The lithologies and stratigraphic relationships are
illustrated in Figure 5. Between Cochabamba and Santa
Cruz, a thin limestone unit of early Wenlock age
(Merino, 1991; Gagnier et al., in press) occurs in the basal
part of the late LlandoveryLudlow shales (Figure 2D),
which postdates a Llandovery hiatus. Slumped sand-
stones and siltstones are common in the Pridoli portion
of the series in several areas of Bolivia (e.g., Cocha- Figure 6Simplified Late CambrianLate Devonian
bamba-Oruro road, Tarija, and Sierras Chiquitanas) and evolution of the central Andean region. APT, Appalachian-
possibly indicate contemporaneous tectonic instability. Patagonian terranes (see Figure 3) (Ramos, 1988a;
Two prominent transgressions are recognized. The first Mpodozis and Kay, 1992); PPO, paleo-Pacific ocean.
is recorded by the lower Kirusillas Formation
(Wenlockearly Ludlow) and is probably a continuation patterns from Late Silurian to Late Devonian time
of the transgression that started in Llandovery time. The suggest large-scale tectonic modifications in the architec-
second is reflected by the Icla Formation (Pragian ture of the basin. The Late Devonian reorganization of
Emsian). Two less important transgressions also paleocurrents toward the southeast may be related to the
occurred. The earliest Lochkovian transgression is Hercynian deformation in Peru (Laubacher, 1978;
recorded by a shale unit between the Tarabuco and Santa Mgard, 1978).
Rosa formations (Chululuyoj Formation) (Racheboeuf et Onlap toward the northeast is conspicuous and
al., 1993), whereas the late Givetian transgression is possibly has exploration significance. Whereas there is
reflected by the Los Monos Formation (Sempere, 1990; no obvious stratigraphic gap in the Ordovician
Racheboeuf et al., 1993). However, the latter has been Devonian succession in the Cordillera Oriental north of
assigned an age as early as earliest Eifelian (Lobo-Boneta, 20 S lat, the Ashgill diamictites, which disconformably
1975; McGregor, 1984; Prez, 1990). These four transgres- overlie Caradoc sandstones, are disconformably overlain
sions were each followed by progradation of shallower by upper Llandovery shales or by the lower Wenlock
facies (Figure 5). The Lochkovian Santa Rosa and the Sacta limestone in the mountain area between
upper Eifelianlower Givetian Huamampampa forma- Cochabamba and Santa Cruz. The Sacta limestone
tions reflect important lowstand periods. overlies Eocambrian rocks in the nearby Andean
Figure 2D shows paleocurrent data for the Pridoli foothills. Eocambrian rocks in the Sierras Chiquitanas are
Lochkovian, upper Eifelianlower Givetian, and overlain by Upper Silurian sandstones (Lpez et al.,
Frasnianmiddle Famennian deposits. Late Silurian 1982). In contrast, in the northwestern sub-Andean belt,
Middle Devonian paleocurrents do not sustain Isaacsons Upper Silurian shales and sandstones overlie deeply
(1975) assumption that coeval sediments were derived weathered Caradoc sandstones. In the Pando well in
from the Arequipa massif. Changes in paleocurrent northern Bolivia, Lower Devonian sandstones rest on
Phanerozoic Evolution of Bolivia and Adjacent Regions 215

Figure 7Simplified stratigraphic cross section of the Carboniferous and Permian showing chronology used in this paper.
Localities (shown in Figure 2E): CT, Copacabana peninsula and Tiquina straits; RK, Ro Kaka; CH, Charagua area; AP, Angosto
del Pilcomayo. Black indicates carbonates; other lithologic symbols are common use. Information partially from YPFB propri-
etary data and Daz-Martnez (1991). Rapid transition between Chuquisaca and Villamontes supersequences (RK and CH
sections) is probably a characteristic of the basin axis (Figure 2E).

Precambrian gneisses. Onlap of this magnitude in the magmatism in the Paleozoic. Mukasa and Henry (1990)
Beni-Chaco basin is expected to have formed potential believe that the San Nicols batholith of southern Peru
stratigraphic traps. was probably emplaced in two short-lived magmatic
Maximum known sedimentation rates are high: they events at about 425 and 394388 Ma. These U-Pb dates
average about 70, 85, and 100 m/m.y. for the late Llan- indicate late Wenlockearly Ludlow and late Pragian
doveryLochkovian, Pragianearly Givetian, and late early Emsian ages, respectively, broadly corresponding
Givetianmiddle Famennian intervals, respectively. to the Kirusillas-Unca and Icla transgressions.
Subsidence was rapid and more intense to the southwest In northern Chile, the first known onlap of the Ocloyic
and increased through time. In contrast to the late basement is of Early Devonian age and the upper part of
CaradocLlandovery interval during which sedimenta- the succession is of Early Carboniferous age (Breitkreuz,
tion rates were much higher (~250 m/m.y.) in the 1986). Intensely deformed Upper DevonianMississip-
Ulloma-Coipasa-Uyuni domain, thickness variations are pian flysch occur west of this shallow marine belt and are
small and the principal units widespread. This broad- thought to represent either fore-arc deposits deformed in
scale regional subsidence is attributed to continuous accretionary wedges (Bell, 1987) or the fill of a turbiditic
tectonic loading of the craton margin transition by trans- trough that was closed in Mississippian time and subse-
pressional processes (Figure 6). Such a marine foreland quently intruded by posttectonic granitoids (Breitkreuz
basin interpretation is supported by the Late Devonian et al., 1989).
Mississippian epeirogenic deformation of the Bolivia-
Peru basin.
In summary, the overall stratigraphic succession LATE FAMENNIANMISSISSIPPIAN
reflects extensional subsidence of the trough that
straddled the Ulloma-Coipasa-Uyuni domain in late The Carboniferous succession of Bolivia is located
Caradocearly Llandovery time and transpressional mainly in the Beni-Chaco basin and adjacent sub-
foreland basin subsidence afterward. Andean belt, in the Tarija-Teoponte belt, and in the
In northwestern Argentina and southern coastal Peru, Huarina fold and thrust belt northeast and southeast of
numerous granitoids were intruded during the Late Lake Titicaca (Figure 1). A large portion of Bolivian oil
OrdovicianEarly Devonian (Bahlburg, 1990; Mukasa and gas reserves occurs in Carboniferous strata.
and Henry, 1990; Rapela et al., 1992). They are believed The chronostratigraphy of this supersequence is still
to be the earliest evidence of subduction-related being refined (Figure 7). There have been some recent
216 Sempere

significant biostratigraphic studies (Merino, 1987; Surez- Huarina Fold and Thrust Belt
Riglos et al., 1987; Rsler et al., 1989; Merino and Blanco,
1990; Vavrdov et al., 1991). In contrast with the under- Carboniferous strata in the Lake Titicaca area sharply
lying stratigraphy, Carboniferous facies are highly overlie the Devonian. Daz-Martnez (1991) have divided
variable and inhibit lithologic correlation. The literature these strata into two units. First is the Ambo Group
is replete with conflicting and chaotic stratigraphic (latest FamennianVisean or early Serpukhovian) (Rsler
schemes, although Helwig (1972) has addressed some of et al., 1989; Vavrdov et al., 1991), which includes a basal
these issues. Furthermore, the Hercynian metamor- glaciomarine horizon and dark shales grading upward
phism and deformation of Bard et al. (1974) and into cross-bedded sandstones with intercalated diamic-
Martinez (1980) is now known to be of Late Triassic age tite beds, coals, and a thin cap of reddish fluvial sand-
(Farrar et al., 1990). This deformation is not related to the stones. The second unit is the Titikaka Group, which
Late Devonian or Early Mississippian deformation consists of estuarine sandstones and supratidal
described in west-central Peru (Mgard, 1978) and west dolomites overlain by transgressive fossiliferous lime-
of Lake Titicaca (Laubacher, 1978), although there is stones and marls (Middle Pennsylvanian conodonts) (D.
evidence that some Late Devonian or Mississippian Merino, 1992, personal communication). The Titikaka
uplift and local folding did occur in Bolivia. The Late Group, which in this area spans the Early Pennsyl-
Mississippian U-Pb age obtained on the Amparaes vanian-Early Triassic interval, corresponds to the
syntectonic granite (Dalmayrac et al., 1980) is considered Copacabana and Bopi formations of the northwestern
to be dubious because three concordant Late Triassic U- sub-Andean belt and the Tarma, Copacabana, and Mitu
Pb ages have been obtained on similar plutons. (partly) groups of Peru. It is capped either by Triassic
sandstones or by Upper Cretaceous or upper Oligocene
Northwestern Sub-Andean Belt and strata.
Adjacent Beni Basin Because the Ambo Group disconformaby overlies
Lower Devonian strata in several outcrops forming a
In the northwestern sub-Andean belt and in adjacent northwest-striking alignment, it postdates a local uplift
Beni basin, the Carboniferous transitionally overlies the of Late Devonian or earliest Mississippian age. Further-
Upper Devonian and is capped by the major erosional more, PennsylvanianLower Permian limestones onlap
unconformity at the base of the Jurassic Beu Formation. It topography formed during the Late Devonian or Missis-
comprises a lower succession characterized by black sippian in the Huarina fold and thrust belt and
mudstones and common plant fossils and an upper Cochabamba area. The disconformable relationships at
succession of carbonates, eolianites, and subsurface evap- the base of these limestones contrasts with the
orites. Hercynian deformation interpretation of Martinez
The lower succession (Retama Group, latest (1980). However, on a regional basis, angular unconfor-
DevonianVisean or early Serpukhovian) is equivalent to mities represent a SilurianTriassic hiatus, reflecting a
the Ambo Group of Peru. It contains conglomerates, compressive deformation of inferred Late Devonian
thick mudstone slumps, some stratified sandstone slump Mississippian age. This local deformation merges
blocks, and shallow marine black laminated shales, laterally into an uplifted area spanning much of the
which grade upward into sandstones and coals. Dark present-day Cordillera Oriental, upon which Pennsyl-
diamictites occur in the upper part. vanianUpper Permian strata onlap. This uplift was
The upper succession (Copacabana Formation, latest probably related to a dextral motion of the Khenayani-
Mississippian or earliest Pennsylvanian to earliest Turuchipa paleostructural corridor (Sempere et al., 1991).
Permian and locally Late Permian) consists of fossilif- The eroded detritus was probably deposited in the
erous limestones and marls, white calcareous sand- FrasnianMississippian basin.
stones, thin green tuffaceous beds, dolomites, some
anhydrite in the subsurface, some dark to mottled shales, Chaco Basin, Sub-Andean, and
and a few red siltstones, sandstones, and paleosols. It Tarija-Teoponte Belts
corresponds to the Tarma and Copacabana groups of
Peru and is reviewed in the next section. In the Chaco basin, sub-Andean, and Tarija-Teoponte
The contact between these two successions is sharp belts, the Mississippian locally overlies the Upper
and, on the basis of conodont biostratigraphy, would lie Devonian gradationally. It consists of the Macharet and
within the Serpukhovian or at the MississippianPenn- Mandiyuti subgroups (Figure 7). In the southernmost
sylvanian boundary (D. Merino, 1992, personal commu- sub-Andean belt, there is a gradational contact between
nication). This contact marks a significant paleoclimatic the two subgroups. Acritarchs in the Macharet, marine
change in northwestern Bolivia, from cool-temperate to faunas in the lower Mandiyuti, and sedimentary facies
warm and arid conditions. The signature of this climatic suggest that deposition occurred in marine environ-
change should facilitate regional correlation. Further- ments. There are marked lithologic contrasts between
more, the major global regressive-transgressive event of mud-dominated and sand-dominated Macharet facies,
Serpukhovian to earliest Pennsylvanian age (Saunders whereas the Mandiyuti is predominantly sandy and
and Ramsbottom, 1986; Veevers and Powell, 1987) is shows nearshore to shore zone facies affinities near the
apparently recorded in facies near the RetamaCopaca- top. In seismic sections, the lower part of the Mississip-
bana contact and should be regionally widespread. pian interval is characterized by discontinuous reflection
Phanerozoic Evolution of Bolivia and Adjacent Regions 217

patterns, whereas the upper part is mainly featureless erosional topography indicates tectonic instability in the
because it is sand prone. This twofold lithologic partition Chaco basin. Reorganization of paleocurrents in the
represents a major change in sedimentary dynamics, latest Devonianearliest Carboniferous (compare Figures
interpreted as a rapid overall shallowing of the deposi- 2D and 2E) demonstrates the growth of the northeast-
tional systems. trending Chaco Boreal high. As in other parts of Bolivia,
There is also a conspicuous difference in color large-scale slumping shows that sediments were fed
between the two subgroups: the clay fraction in the from tectonically active highs and hence documents
Macharet is dark gray to purplish and locally contains periods of tectonic instability. Regional mapping shows
plant debris. In contrast, it is bright brown-red in the the distribution of these paleogeographic highs: Chaco
uppermost Macharet and the entire Mandiyuti. This Boreal, Cordillera Real, Madidi, and Guapor craton
color change is attributed to the Serpukhovian climatic uplifts (Figure 2E), the last two of which were probably
event recorded in northwestern Bolivia. Rapid shal- controlled by a Cambrian crustal heterogeneity. The
lowing of depositional environments in the Chaco basin distal downwarped areas are identified by the occur-
suggests a contemporaneous sea level drop related to a rence of a transition between the Chuquisaca and Villam-
worldwide Serpukhovian regression (Veevers and ontes supersequences, a lower thickness of the Macharet
Powell, 1987). The MacharetMandiyuti contact thus subgroup, or a thicker and well-defined Charagua
appears to be of early Serpukhovian age. member.
The late Paleozoic stratigraphy shown in Figure 7 is Available data suggest that crustal deformation,
based on these lithostratigraphic criteria and recent bios- involving uplift and minor folding, progressed from
tratigraphic data. I equate the Saipur Formation of the west to east during the Late Devonian and Mississippian.
Charagua area (Surez-Soruco and Lpez-Pugliessi, While structural highs and slopes formed in parts of the
1983), which has late FamennianVisean palynomorphs, Beni-Chaco basin in the late FamennianEarly Mississip-
to the lower Macharet (Itacua, Tupambi, Itacuami, and pian, other areas were downwarped and received
Tarija formations) of the Villamontes area. Marine fossils slumps and laminated black shales. Compared to the
of the Levipustula levis zone, the base of which coincides underlying Frasnianearly Famennian shallow marine
with the ViseanSerpukhovian boundary in Australia, strata, there was a significant increase in depth. Slopes
occur in a regularly stratified, reddish siltstone were tectonically maintained during the Mississippian,
mudstone intercalation that is traditionally described as controlling sigmoidal geometries of progradational units
the local Taiguati Formation. I prefer, however, to call and substantial variations in thickness. The thicknesses
it the Charagua member because it appears to be located of the Macharet and Mandiyuti sections crossing the
within the lower part of the Mandiyuti (Figures 2E, 7). Chaco basin display such relationships. The Chaco basin
The Macharet and Mandiyuti facies are broadly received a considerable sand supply from the southeast
similar and are dominated by resedimented deposits, and east, that is, from the Chaco Boreal high.
including rarely stratified diamictites, muddy mass Evolution of the Carboniferous climate in Bolivia,
flows, slumped sandstonemudstone intercalations, from cool and humid to warm and arid, matches the
grain flows, debris flows, cross-bedded sandstones migration of the south pole toward the southeast or east-
(commonly conglomeratic), slumped sandstones masses, southeast (Veevers and Powell, 1987). South American
thick to thin regularly bedded Bouma-type turbidites climatic gradients were probably high and oriented
interbedded with siltstones and mudstones, regularly northeast to north-northeast, defining climatic strips
bedded fossil-bearing fine-grained sandstones and where different facies assemblages were able to accumu-
mudstones, rarer laminated mudstones, and local intra- late at the same time. Current interpretation of palyno-
basinal olistoliths that reach several meters in thickness. zones suggests that the Paran basin was probably
The diamictites, debris flows, grain flows, mud flows, subjected to subglacial erosion during the Mississippian
and slumps are clast bearing. The clasts are millimeter to and to glacially influenced marine sedimentation during
meter size, are well-rounded to angular, and consist of a the PennsylvanianEarly Permian glacial retreat (Frana
variety of sedimentary, igneous, and metamorphic rock and Potter, 1991). The Amazon basin, Bolivia, and
types. Some of these are faceted and striated, suggesting western South America experienced cool to temperate
transport from a glaciated area. Ice rafting explains the and subsequently subtropical conditions during these
presence of some of the larger clasts. Although reworked periods.
glacial clasts occur in the Mandiyuti, the most convincing The average sedimentation rate for the Mississippian
glaciomarine features are seen only in the Macharet. of Bolivia is estimated to be about 3540 m/m.y., indi-
Most of the resedimented facies were deposited in cating a significant decrease compared to the Devonian.
channels, ranging in width from a few meters to a few However, because of the inferred increase in depth, it is
kilometers. Several large channels and valleys (several possible that the maximum subsidence rate increased in
hectometers or kilometers wide and up to several the late Famennianearliest Mississippian before
hundred meters deep) are clearly visible in many seismic decreasing after that.
sections in the Chaco basin. The apparent depth of these The Late DevonianEarly Mississippian deformation
channels or valleys increases upsection. Channel stacking known in Peru (Mgard, 1978; Dalmayrac et al., 1980)
and amalgamation of their channel fills has resulted in and northern Chile (Sempere, 1989) and the inferred
valley complexes locally more than 500 m deep. uplifts in the present-day Huarina fold and thrust belt,
The abundance of resedimented facies filling an Chaco Boreal, Cordillera Real, Madidi, and Guapor
218 Sempere

Figure 9Late PermianEarly Triassic paleogeography.


General legend same as in Figure 2. Key: 1, erosional
boundary of basin in sub-Andean belt; 2, known remnants
of Upper Permian strata in Andean area; 3, presence of
clasts of cherty limestones typical of Upper Permian
deposits in Miocene strata, suggesting proximity of Upper
Permian outcrops in the early Neogene; 4, inferred deposi-
tional area before Mesozoic erosion; 5, Late Permian basic
magmatism; 6, areas of felsic magmatism; 7, interstratified
Figure 8Map of western Gondwana showing approxi-
tuffs; 8, known extent of the Late PermianEarly Triassic
mate distribution of main Paleozoic geotectonic features.
deformation, including granitoids; 9, fold strikes suggest-
Key: 1, Middle Ordovician deformation; 2, Late Ordovician
ing southward continuation of structures. Gondwana-age
Devonian granitoids; 3, Late DevonianMississippian
folds are generally roughly parallel to Andean structures. In
deformation; 4, CarboniferousPermian basins; 5, Late
the Cordillera Oriental of southern Peru and western
PermianEarly Triassic deformation. In Bolivia, basin is
Bolivia, some structures traditionally assigned to the Late
bounded by structural highs that formed during late
PermianEarly Triassic may be of Late Triassic age.
FamennianMississippian sedimentation (Figure 2E). CBH,
Chaco Boreal high; GCU, Guapor craton uplift; MU,
Madidi uplift. Data mainly from Dalmayrac et al. (1980), of carbonates and mudstones (Figures 2F, 9). This
Archangelsky (1987), Ramos (1988a), and Rapela et al. interval was also characterized by a substantial decrease
(1992).
in tectonic activity and by the establishment of a subtrop-
ical climate. Depositional environments were predomi-
craton uplifts provide evidence of synsedimentary insta- nantly shallow marine and continental, with subordinate
bility. This evidence clearly points to a time of consider- eolianites in the Pennsylvanian and Early Permian. Penn-
able tectonic activity in the Bolivia-Peru basin in the Late sylvanianUpper Permian limestones onlap the older
Devonian and especially the Mississippian. The Missis- topography in several areas of the present-day Cordillera
sippian sedimentation was the culmination of the Oriental. The PennsylvanianPermian deposits in the
SilurianDevonian cycle of evolution. The uplifts and Amazon basin resemble those in northern Bolivia.
downwarps observed in Bolivia probably represent the The stratigraphy used in this paper implies that Penn-
foreland response to the transpressional and compres- sylvanianLower Permian carbonate deposits in the
sional processes that affected the western margin of northwest grade laterally into fluvial and eolian deposits
Gondwana (Figure 8). in the southeast (Figure 7) and may even be in facies with
the underlying shallow marine clastics. However, the
boundary between the Villamontes and Cuevo super-
PENNSYLVANIANEARLY TRIASSIC sequences in the Chaco basin is not as sharp as it is in
northwestern Bolivia. Note that the fluvial-eolian
The PennsylvanianEarly Triassic interval is charac- Cangapi Formation at the base of the Cuevo super-
terized in western Bolivia by the establishment of a sequence in the southern sub-Andean belt was originally
carbonate platform and in southeastern Bolivia by north- included in the Mandiyuti Subgroup. Its lower strata
westward progradation of littoral and continental depo- may be continental equivalents of the shallow marine
sitional systems. This was followed by a MiddleLate upper San Telmo Formation in some parts of the Chaco
Permian restricted marine transgression and deposition basin. In this sense, the Mandiyuti Subgroup may span
Phanerozoic Evolution of Bolivia and Adjacent Regions 219

this deformation is known from outcrops north of Lake


Titicaca (Martinez, 1980) and west of Cochabamba
(Kennan, 1993). It might correspond to the so-called
Eohercynian deformation in the Cordillera Real; it did
not affect the rest of the country. It was partially coeval
with intense compressional deformation in the Andes
north of 7 S lat (Dalmayrac et al., 1980), the San Rafael
deformation of west-central Argentina (Ramos, 1988b),
and the transpressional deformation of the Sierra de la
Ventana and Cape foldbelts (Cobbold et al., 1986) (Figure
8). Accretion of terranes along the southwestern margin
of South America ended in Late Permian time (Mpodozis
and Kay, 1992).
This widespread Gondwana-age deformation is
postdated by a postorogenic calc-alkaline magmatism of
EarlyMiddle Triassic age (Laubacher, 1978; Soler and
Bonhomme, 1987; Mpodozis and Kay, 1992). By the late
Middle Triassic, continental tholeiitic compositions were
more common, indicating regional extension (Soler and
Sempere, 1993). Onset of tensional conditions in Bolivia
in the late Middle Triassic initiated the Serere sedimen-
tary supersequence.
In northern Chile, metamorphism related to terrane
accretion ended in the Pennsylvanian (Cordani et al.,
1988). This was superseded by widespread magmatism
of PennsylvanianTriassic age (Breitkreuz et al., 1989;
Breitkreuz, 1991; Mpodozis and Kay, 1992). It is this
magmatic episode that is believed to have generated the
numerous tuffs observed in the Bolivian and eastern
Figure 10Simplified stratigraphic sections for the Middle Peruvian basin.
TriassicMiddle Jurassic. See Figure 11 for locations. ER,
Entre Ros area; R, Ravelo; S, Sayari; T, Tarabuco. Key:
1, carbonates; 2, mudstones; 3, medium to finegrained
sandstones; 5, conglomerates; 6, mainly alluvial facies; 7,
mainly eolian facies; 8, evaporites; 9, basalts. MIDDLE TRIASSICMIDDLE JURASSIC
The Middle TriassicMiddle Jurassic stratigraphy of
the Late Mississippianearliest Pennsylvanian. More Bolivia has recently been revised by Oller and Sempere
biostratigraphic work will decide this issue. (1990) (Figure 10). Late Middle Triassic rifting developed
Sedimentation rates for the PennsylvanianEarly in several areas as a prelude to the breakup of Pangea. It
Permian interval were less than 25 m/m.y. in Peru, about led to the formation of numerous grabens (e.g., Mitu
1520 m/m.y. in the Lake Titicaca area of western Group of Peru and Sayari and Ipaguaz formations of
Bolivia, and less than 15 m/m.y. in the Chaco basin. Bolivia) accompanied by alkaline and tholeiitic
Sedimentation rates were also low for the Late magmatism (Mgard, 1978; Kontak et al., 1985; Soler and
PermianEarly Triassic interval, generally less than 1015 Sempere, 1993). In Bolivia, tholeiitic basalts from Entre
m/m.y. The intensity of subsidence had thus decreased Ros and Tarabuco cannot be geochemically distin-
dramatically after the Mississippian. guished (Soler and Sempere, 1993) and have apparent
Although mainly a restricted marine deposit, the K-Ar ages of 233 Ma (Sempere et al., 1992) and about 171
Vitiacua Formation marks a major transgression that Ma, respectively.
started in late Kungurian or Kazanian time and covered The Mitu Group of Peru (Figure 11) spans the Late
a wide region. Toward the west, it was the most PermianMiddle Triassic interval and appears to consist
extensive transgression since the Devonian. The lower of two successions containing poorly understood
Vitiacua includes black limy mudstones resembling magmatic suites. Mafic volcanism developed as early as
those of the Irati Formation in the Paran basin and the Late Permian time in Peru. Basaltic flows, yielding
Whitehill Formation in the Karoo basin. These facies PermianEarly Triassic apparent K-Ar ages, are exposed
represent a highstand that flooded Gondwana as far along the southeastern shore of Lake Titicaca (S.
away as Australia (Sempere et al., 1992). McBride, in Kontak et al., 1985), whereas mafic rocks
Deposition of the middle PermianEarly Triassic south of Lake Titicaca have Late PermianEarly Triassic
strata was partially coeval with the intense compres- apparent K-Ar ages. Swarms of mafic sills and dikes
sional or transpressional intracratonic deformation in the (0.11 m thick) intruding OrdovicianDevonian strata in
Cordillera Oriental of central and southern Peru the Cordillera Oriental (Figure 11) are associated with
(Laubacher, 1978; Soler and Bonhomme, 1987). In Bolivia, paleograbens preserving Permian strata. These intrusives
220 Sempere

Figure 11Middle Triassic


Middle Jurassic paleogeography.
General legend same as Figure
2. Locations of stratigraphic
sections in Figure 10: ER, Entre
Ros; R, Ravelo; S, Sayari; T,
Tarabuco. Key: 1, Late Triassic
Jurassic marine deposits; 2,
EarlyMiddle Jurassic deepest
facies; 3, Middle Jurassic
turbidite fan progradations;
4, Jurassic arc magmatism; 5,
edge of Jurassic marine basin in
northern Chile; 6, Mitu volcanic
deposits; 7, EarlyLate Triassic
plutons, including the Abancay
(a) and Zongo-Yani (zy) cata-
clastic Late Triassic plutons; 8,
Middle TriassicEarly Jurassic
fluvial deposits in Bolivia; 9,
fluvial paleocurrents; 10, occur-
rences of evaporites; 11, Middle
Late Triassic continental tholei-
ites; 12, altered mafic dike and
sill swarms (Triassic?); 13, extent
of Jurassic desert in Bolivia;
14, Jurassic eolian prograda-
tions. Jurassic eolianites in
Paran basin crop out just east
of map limit. Data for Peru and
Chile from Dalmayrac et al.
(1980), Vicente et al. (1982), Soler
and Bonhomme (1987),
Grschke et al. (1988), and
Muoz and Charrier (1993).

are believed to have been emplaced during the Middle Mild extension persisted in Peru after this event,
Triassic extension. Other probable counterparts are the resulting in transgression of the Norian sea from the
Cerro Sapo nepheline syenite and kimberlite-like rocks northwest into the Cordillera Oriental area (Mgard,
west of Cochabamba, mafic dikes intruding a Neopro- 1978; Dalmayrac et al., 1980; Jaillard et al., 1990). The
terozoic pluton and Lower Ordovician strata south of Marine NorianLiassic limestones locally overlie the
Tarija, and an alnoite dike described in northwestern Mitu Group with a slight angular unconformity
Argentina by Meyer and Villar (1984). (Mgard, 1978) and are in turn overlain in eastern Peru
The grabens were filled by fluvial and lacustrine red by fluvial and eolian deposits (Sarayaquillo Formation)
beds and evaporites (Figure 11), which permitted local that correlate with those of Bolivia (Serere super-
preservation of upper Paleozoic strata. Clasts of Upper sequence). In southern coastal Peru, extension in the
PermianMiddle Triassic granitoids are known in the Yura trough culminated in BajocianBathonian time; this
upper Mitu Group. Extensional structures and magmatic trough was eventually filled in by the latest Jurassic
rocks are widespread in western South America (e.g., (Vicente et al., 1982; Jaillard et al., 1990).
Caputo, 1991; Ramos and Kay, 1991; Soler and Sempere, The termination of rifting in Bolivia was probably a
1993). consequence of tectonic reorganization that induced local
Reorganization of the regional stress field at 225220 transpressional deformation. Subsequent Late Triassic
Ma produced local intracratonic transpression, as Middle Jurassic fluvial and eolian sedimentation was
evidenced by the pervasive cataclasis of the Zongo-Yani controlled by postrift thermal subsidence, reflected in
(Bolivia) and Abancay (southern Peru) Triassic intrusions onlapping relationships (Oller and Sempere, 1990)
and deformation of their enclosing strata (Bard et al., (Figure 11). Eolian deposits were widespread in western
1974; Dalmayrac et al., 1980; Farrar et al., 1990). At least Gondwana during the Jurassic, including the Karoo and
some grabens or transtensional structures were inverted Paran basins and sub-Andean central Peru (Boquern
during this event; examples are found in western Brazil Formation) (E. Bosc, 1986, personal communication), in
(Caputo, 1991), eastern Peru (M. de Barros and E. agreement with global climatic models (Chandler et al.,
Carneiro, 1992, personal communication) and northern 1992). In Bolivia, the overall pattern of Jurassic eolian
Bolivia (YPFB, proprietary data). A 220200 Ma regional progradations (Figure 11) suggests that many dunes
metamorphic-plutonic event is also recognized in south- were longitudinal and produced by south-southwesterly
western Ecuador (Tahuin Group) (Aspden et al., 1992). winds.
Phanerozoic Evolution of Bolivia and Adjacent Regions 221

LATEST JURASSICTURONIAN
The KimmeridgianPaleocene Puca Group of Bolivia
(Figures 12, 13) records the evolution of the distal
Andean back-arc basin. Bolivia was part of cratonic
South America in the Jurassic, but was subsequently
captured by the Andean system in the latest Jurassic
during a period of large-scale rifting. This episode is
attributed to onset of extensional and transtensional
conditions in northern Chile and coastal Peru and was
contemporaneous with the South Atlantic extension. In
Bolivia, extensional reactivation of the CPKT formed the
highly fragmented Potos basin, which was filled with
continental red beds (Figures 13, 14) (Sempere, 1994).
Deposits of this age are known only from the Andean
domain. Silcretes developed contemporaneously on the
Jurassic sands in the sub-Andean-Beni-Chaco region.
Rifting created a tilted block structural relief that was
onlapped by fine-grained younger sediments. Rifting
was accompanied by local alkaline basic volcanism (Soler
and Sempere, 1993). The oldest and most important
extensional episode took place during deposition of the
Condo conglomerates in the lowermost Puca Group.
Local inhomogeneities between the Kimmerigian
Hauterivian and BarremianTuronian paleocurrent
patterns (Figure 14) suggest that topography was
modified by extensional processes during the Early
Cretaceous. The intensity of extension appears to have
waned through late NeocomianAptian time. By Albian
time, fault-controlled subsidence was abandoned. Wide-
spread regional subsidence is reflected in red mudstones
that blanketed the old rift depocenter and onlapped the
Paleozoic basement beyond. A postrift phase of thermal
subsidence is inferred.
The shallow marine Miraflores carbonates were
deposited during the CenomanianTuronian interval
(Jaillard and Sempere, 1991) (Figure 14). Organic-rich
bioturbated limestones are conspicuous at the base (early
Cenomanian) of the Miraflores Formation. This unit is
thin and has a fairly uniform thickness, reflecting low
sedimentation rates and the absence of synsedimentary
tectonism. It is believed that this transgression had a
global eustatic origin (Sempere, 1994).

SENONIANPALEOCENE
Figure 12Stratigraphic section for the Late
Jurassicearly Eocene from the Miraflores syncline near The SenonianPaleocene supersequence is best
Potos, overlying gently folded Upper Ordovician strata. developed in Andean Bolivia, although some
Key: 1, conglomerates; 2, sandstones; 3, predominantly CampanianMaastrichtian strata are known from the
mudrocks; 4, limestones; 5, red color; 6, green or gray
northwestern and central sub-Andean belt and the
color; 7, channels; 8, evaporites; 9, roots. Ranges of thick-
nesses in the basin are 01200 m for the Kimmerigian adjacent Beni-Chaco basin. In the Andean depocenter,
Albian, 1525 m for the CenomanianTuronian, and this muddy supersequence is up to 1.2 km thick (Figure
2001200 m for the Senonianmiddle Paleocene. 12). It was deposited in the external part of a wide,
underfilled foreland basin (Figure 15A) (Sempere, 1994).
The Andes were beginning to form at that time,
An important consequence of this evolution is that producing only a minor flexure of the South American
Late Triassic tectonic reorganization not only terminated lithosphere; this constitutes the earliest record of a true
this period of widespread intracontinental extension but Andean-age foreland basin.
also apparently resulted in resumption of subduction This record consists of two sequences that span the
along the Pacific margin. ConiacianCampanian (about 15-m.y. duration) and
222 Sempere

Figure 13Restored sedimentary cross sections (after Sempere, 1994). See Figure 15A for locations. Strata are decom-
pacted below the middle Selandian unconformity as reference level. Thicknesses are mainly from YPFB proprietary data and
personal observations, but those of nonoutcropping rock units are hypothetical. (A) CPKT, Khenayani-Turuchipa paleostruc-
tural corridor (see Figure 1). (B) Thicknesses from Sevaruyo-Tambillo area are possibly exaggerated and need confirmation.
This tentative restoration assumes a gross 50% shortening as a minimum and more shortening at major thrusts. Symbols: 1,
mostly mudrocks (and limestones in the El Molino Formation); 2, mostly sandstones; 3, evaporites; eO, Early Ordovician; lO,
Late Ordovician; eS, Early Silurian; lS, Late Silurian; eD, Early Devonian; lPz, late Paleozoic; R, Ravelo Fm (Late
TriassicJurassic); Cn, Condo Fm; Ko, Kosmina Fm; LP, La Puerta Fm; Su, Sucre Fm; Ta, Tarapaya Fm; , basic lava flows;
M, Miraflores Fm; Ar, Aroifilla Fm; Ch, Chaunaca Fm; To, Torotoro Fm; Cr, Coroma Fm; EM, El Molino Fm; SL, Santa Luca
Fm (u, upper); Im, Impora Fm; Cy, Cayara Fm.

Maastrichtianearly Selandian (about 16-m.y. duration) The base of the supersequence records a period of
intervals, respectively. There are a few organic-rich levels intense rifting, accompanied by relatively widespread
that are among the best hydrocarbon source rocks in the alkaline basaltic volcanism (Soler and Sempere, 1993)
Bolivian Altiplano. and locally high subsidence rates. This extension is inter-
Phanerozoic Evolution of Bolivia and Adjacent Regions 223

An increase in the sedimentation rate and an influx of


clean sand from the east near the CampanianMaas-
trichtian boundary are attributed to reactivation of
thrusting in the coastal orogen, accompanied by uplift of
a forebulge in the sub-Andean-Beni-Chaco region. This
episode coincides with a major transgression from the
north or northwest in the Maastrichtian (Gayet et al.,
1993). Restricted marine to lacustrine marl and carbonate
facies locally rich in organic matter were deposited until
the early Paleocene. Lacustrine and alluvial depositional
environments characterized much of the Paleocene.
Apart from early Selandian reactivation of the CPKT,
there was no major tectonic activity. Uniform and steady
subsidence suggests that flexural loading by thrust sheet
encroachment was important in the coastal orogen
(Sempere, 1994). A major modification of this tectonic
system occurred in the middle Selandian when a wide-
spread unconformity was formed in Andean Bolivia
(Marshall and Sempere, 1991).

LATEST PALEOCENERECENT
The Cenozoic evolution of Bolivia is subject to
conflicting interpretations. However, there is a growing
consensus that the central Andes were built by consider-
able crustal shortening (Roeder, 1988; Sempere et al.,
1990a). All the basins that developed in the Bolivian
mountains or plains during this time were related to the
building of the Andes. Cenozoic basin evolution is
expressed in a twofold record that contains latest
Paleoceneearly Oligocene (about 30-m.y. duration) and
late OligoceneRecent (about 27-m.y. duration)
Figure 14KimmeridgianTuronian paleogeography in sequences of unlike characteristics.
Bolivia (after Sempere,1994). The main depositional area is During the latest Paleoceneearly Oligocene interval,
the Potos basin. Tight lines are CPKT. Circled numbers: 1, Andean Bolivia was a foreland basin east of the Andes
Tawarreja-Thokori graben; 2, San Lorenzo-Tupiza graben; (Figure 15B). Coarsening- and thickening-upward conti-
3, Sevaruyo subbasin; 4, Norte del Lago basin. Kimmeri- nental red beds, 25 km thick, accumulated in what is
gianHauterivian interval: 1, conglomeratic facies; 2, sandy
facies; 3, basaltic flows; 4, paleocurrents. Barremian
now the Altiplano and Cordillera Oriental. This basin
Albian interval: 5, paleocurrents; 6, gypsum in Tarapaya probably resembled the present-day Beni-Chaco plain.
Fm. CenomanianTuronian interval: 7, extent of Miraflores The existence of a continuously subsiding basin indicates
Fm; 8, paleocurrents obtained from orientation of that thrusting and tectonic loading were active west of
gastropods or ripple marks. the Altiplano. The base of the sequence records a signifi-
cant change in sedimentation, approximately coinciding
preted as the foreland reaction to the onset of eastward with the end of emplacement of the Peruvian coastal
encroachment of the Andes. In Bolivia, it reactivated Late batholith, marking a turning point in the pattern of
JurassicEarly Cretaceous extensional structures and subduction (Sempere et al., 1989; Sempere, 1991b;
produced new normal faults. Coniacian playa lakes Marshall and Sempere, 1991). This change is related to
formed in the structural lows. the onset of orthogonal convergence along the margin. A
SantonianCampanian deposits include red mud- late Eocene or Oligocene uplift in the Cordillera Oriental
stones and subordinate sandstones and evaporites; there is reflected in deposition of the Camargo conglomerates
are two thin, organic-rich, restricted marine carbonate (Figure 15B).
intercalations. This stratigraphy onlaps toward the A tectonic, sedimentary, and magmatic upheaval
margins, reflecting the distal foreland response to developed in Bolivia during the late Oligocene and early
Peruvian age deformation affecting the coastal regions Miocene (Figure 16) (Sempere et al., 1990a), involving a
at that time (Jaillard, 1994, for Peru; Aspden et al., 1992, jump of the foreland basin from its Altiplano
for Ecuador). A decrease in the rate of sedimentation and Cordillera Oriental position to the present-day sub-
rapid progradation of immature sands from the west in Andean domain and Beni-Chaco plain. This major reor-
the middlelate Campanian suggest erosion and relative ganization of geologic domains occurred only in the
quiescence of the orogen. Bolivian orocline area (Figure 15B), suggesting that at
224 Sempere

Figure 15Late CretaceousCenozoic paleogeography. General legend same as in Figure 2. (A) Synthetic SenonianDanian
paleogeography. 1, rift axes; 2, marine-influenced MaastrichtianDanian basin (maximum extent of SenonianDanian basin);
3, areas undergoing deformation; 4, magmatic arc; 5, continental foreland sedimentation; 6, approximate extent of the mainly
fluvial Cajones Fm (yielded dinosaur bones and an endemic fish known otherwise in MaastrichtianPaleocene strata, but
might be partially older); 7, Maastrichtian paleocurrents; 8, locations of Figure 13 cross sections. Bolivia-Peru basin was
connected to open sea through Venezuela and possibly southern Ecuador (Gayet et al., 1993). Data partially from Marquillas
and Salfity (1988) and Jaillard et al. (1993). (B) Distribution of main geologic features after the middle Paleocene. External
foreland basins: hatchured, latest Paleoceneearly Oligocene, gray, late OligoceneRecent. Paleocurrents: 1, latest
PaleoceneEocene; 2, late Eocene or Oligocene (Camargo conglomerates); 3, late Oligoceneearly Miocene. Magmatic arcs:
4, Eoceneearly Oligocene; 5, late OligoceneRecent.

least its eastern part formed at that time (Sempere et al., dental, with some periods of increased activity (Sempere
1990a). Emergence of thrust deformation in the et al., 1990a,b). Preexisting structures have largely
Cordillera Oriental split the basin into the coeval controlled the geometry and style of Andean deforma-
Altiplano (intermontane) and sub-Andean-Beni-Chaco tion (Sempere et al., 1989, 1991; Hrail et al., 1990).
(foreland) basins (Figure 16). The Bolivian Andes grew as Ordovician deformation and thickening in the Puna
a mountain belt from that time onward. This first major aulacogen area created a significant crustal hetero-
tectonic episode lasted about 8 m.y., starting at about 27 geneity, which probably influenced the formation of the
Ma. After a period of relative tectonic and magmatic Bolivian orocline (Sempere, 1990).
quiescence between about 19 and 11 Ma, shortening and
magmatism resumed. This produced accelerated subsi-
dence of the sub-Andean-Beni-Chaco basin. In the late PETROLEUM GEOLOGY
Miocene, restricted marine waters flooded the axial
groove of the underfilled Chaco basin from the south. Oil and gas are produced in the central and southern
This transgression is reflected in deposition of the locally sub-Andean belt and adjacent lowlands of Bolivia. Many
organic-rich Yecua Formation (Marshall et al., 1993). The oil seeps are known in the northwestern sub-Andean
late MioceneRecent terrestrial fill of the sub-Andean- belt, as well as in the Altiplano. The stratigraphy of the
Beni-Chaco foreland basin is 35 km thick and coarsens productive sub-Andean and Chaco regions is divided
and thickens upward. Seal rocks occur near its base and into a lower PaleozoicJurassic pericratonic Gondwana
above the Yecua Formation. section and an upper OligoceneRecent Andean section.
The present-day structure of the Andean fold and The sub-Andean fold and thrust belt is of late Neogene
thrust belt of Bolivia involved dcollement levels in Recent age.
Ordovician, Silurian, and Devonian shale-dominated The principal source rocks in Bolivia are Devonian
units. Within the mountain belt, sedimentation shales and Mississippian mudstones. Other potential
continued in a variety of piggyback basins. The largest source rocks include Silurian shales, Pennsylvanian
and most famous of these basins is the Altiplano, which Permian shales and limestones, Upper Cretaceouslower
has a complex structure (Sempere et al., 1990b). Paleocene limestonemudstone units in the Altiplano,
Magmatism has been active since the late Oligocene in and possibly dark shales of the upper Miocene Yecua
the Cordillera Occidental, which is a subduction-related Formation in the central and southern sub-Andean
arc, and in the Altiplano and western Cordillera Occi- foothills and adjacent Chaco basin.
Phanerozoic Evolution of Bolivia and Adjacent Regions 225

Figure 16Synopsis of data concerning the late Oligoceneearly Miocene tectonic crisis in Andean Bolivia (after Sempere et
al., 1990a). BI, Illimani batholith; BQC, Quimsa Cruz batholith; CANP, main Andean thrust; CCR, Cordillera Real thrust; FCC,
Coniri thrust front. (A) 1, Composite schematic section for northern Altiplano; arrows, azimuths of mean paleocurrents; 2,
litho- and magnetostratigraphic sections at Salla, correlation to absolute time scale, and ages of nearby intrusions in BI and
BQC; 3, favored age of Cenozoic heating of Zongo-Yani pluton; 4, stratigraphic section 60 km west-northwest of Santa Cruz,
evidencing age of onset of sub-Andean-Beni-Chaco foreland basin. (B) Cross section illustrating activity of main Andean
thrust system (CANP) and synchronous development of sub-Andean and Altiplano basins at ~20 Ma. Vertical arrows show
location of data displayed above. Wavy pattern, pre-Ordovician rocks.

Reservoir intervals occur in numerous Silurian, of the sub-Andean-Beni-Chaco region, subsidence was
Devonian, Carboniferous, Permian, Triassic, Jurassic, and negligible from the Late Jurassic to Paleogene, resumed in
upper Oligocenemiddle Miocene sandstone units in the the late Oligocene, and increased considerably in the late
sub-Andean-Beni-Chaco region. Lower Silurian, Penn- Miocene, resulting in rapid burial of Paleozoic source
sylvanian, and Permian fractured limestones are also rocks beneath the Andean foreland basin.
possible reservoirs. In the Altiplano, where no discov- Tectonic burial occurred during propagation of
eries have been made, possible reservoirs include Upper Andean thrusting. However, there are differences in the
SilurianLower Devonian sandstones, Upper Creta- style of foreland basin development in latest Paleocene
ceouslower Paleocene limestones, and numerous early Oligocene and late OligoceneRecent time.
Tertiary sandstone units. Migration and trapping of hydrocarbons in the produc-
Although several seal rocks are interstratified in the tive areas is believed to be related to these phenomena.
DevonianMississippian succession, the thick and Furthermore, earlier hydrocarbon accumulations may
extensive upper Miocene mudstone unit is the important have been remobilized. Hydrocarbon generation may
seal horizon in the sub-Andean-Beni-Chaco region. In have occurred in PermianTriassic time due to locally
the Altiplano, potential seals include several Upper high geothermal gradients.
CretaceousEocene mudstone units. In the still poorly explored Altiplano where only five
The timing of hydrocarbon maturation, migration, and exploration wells have been drilled, the burial of
trapping is poorly constrained because of the absence of Paleozoic and Upper Cretaceous source rocks occurred
an established burial history. Sediment accumulation was during development of the latest Paleoceneearly
nearly continuous from the early Paleozoic to Jurassic in Oligocene foreland basin, that is, earlier than in the sub-
most of the Bolivia-Peru basin, albeit with Late Andean-Beni-Chaco region. Neogene magmatism
DevonianMississippian uplift in some parts of the Beni- probably produced locally high geothermal gradients in
Chaco basin and PermianTriassic uplift in Peru. In most this area.
226 Sempere

DISCUSSION Stromatolites are common in strata that span the


PermianTriassic (Sempere et al., 1992) and Creta-
ceousTertiary (Gayet et al., 1993) boundaries.
The Phanerozoic geology of Bolivia preserves a long
record that includes the effects of the dispersal of the Late The main aspects of the evolution of Bolivia and the
Proterozoic supercontinent, the initiation and evolution central Andes can be grouped into three periods: Late
of the western Gondwana margin, the aggregation and CambrianMississippian, PennsylvanianMiddle
consolidation of Pangea, its subsequent fragmentation, Jurassic, and latest JurassicRecent.
and the convergence effects of the westward motion of
South America. In this history, the building of the central Late CambrianMississippian
Andes mountains appears as a comparatively subordi- ( 205 m.y. duration)
nate phenomenon, albeit of impressive magnitude,
related to the dispersal of Pangea elements. Interpretation of the evolution of the central Andean
There are several milestones that summarize the region is illustrated in Figures 4, 6, and 8. While North
evolution of the Bolivia-Peru basin: America separated from South America in the latest
ProterozoicEarly Cambrian (Bond et al., 1984), the
Extension: Late CambrianEarly Ordovician, western margin of Gondwana evolved as a passive
AshgillLlandovery, late Middle Triassic, latest margin that was subjected to large-scale dextral shear.
JurassicEarly Cretaceous, early Senonian This led to the MiddleLate Cambrian formation of the
Compression and uplift (regional or local): Puna aulacogen above a thinned crust. Regional
LlanvirnCaradoc, Late DevonianMississippian, extension also favored the formation along the proto-
Late Permian, Late Triassic, late Paleocene, late Andean domain of a wide marine epicontinental basin
OligoceneNeogene which was the locus of thick sedimentation. Closure and
Magmatism (in specific areas only): EarlyLate deformation of the Puna aulacogen propagated from the
Triassic, late OligoceneNeogene west in the Middle Ordovician, probably reflecting the
High subsidence rates: Late Cambrian onset of subduction along the western margin of
Mississippian, SenonianRecent Gondwana after an inferred plate reorganization. The
Transgressions are grouped into three time rest of the basin evolved as the marine foreland basin of
intervals, with a decrease in amplitude of inunda- this deformation.
tion: OrdovicianMississippian, Pennsylvanian High subsidence rates during latest Ordovician
Late Permian, Late CretaceousDanian earliest Mississippian time resulted from the transcurrent
state of a wide tract running along the margin (Figures 6,
Some points of interest for global comparisons are as 8), induced by oblique subduction. This foreland-type
follows: basin was filled with southwest-thickening, northeast-
onlapping sequences. It was eventually affected by local
deformation and uplift in the Late DevonianMississip-
The Middle Ordovician deformation in the Andes pian. The source rocks associated with known hydro-
was coeval with the Taconic deformation of the carbon reserves in Bolivia were deposited during this
inferred conjugate margin of eastern North period.
America (Keppie, 1993). Development of approxi-
mately synchronous compressional deformation on
both sides of the southern Iapetus ocean and in the PennsylvanianMiddle Jurassic
western United States (Cotkin et al., 1992) suggests (~165 m.y. duration)
that a major plate reorganization occurred in the The consolidation and subsequent fragmentation of
Middle Ordovician. Pangea occurred during the PennsylvanianMiddle
The SilurianDevonian to Jurassic stratigraphies of Jurassic interval, with a pivotal event in the Carnian
Bolivia and the Paran basin are similar. when rifting developed worldwide (Veevers, 1988). In
The Late DevonianMississippian orogenic- the central Andean region, active subduction is believed
epeirogenic activity in the central Andean area was to have ceased in the Permian with consolidation of
contemporaneous with the Antler orogeny in Pangea (Kay et al., 1989). The central Andean geologic
western North America (Smith et al., 1993). evolution includes a complex suite of tectonic, magmatic,
Eolianites occur in the Pennsylvanian, Early and sedimentary events.
Permian, and Jurassic of Bolivia, similar to the Amalgamation of Pangea formed an enormous
western United States (Blakey et al., 1988), which thermal blanket, inducing heat anomalies in the mantle.
occupied a symmetric geotectonic location relative Deep-rooted thermal anomalies probably resulted in
to the equator. fracturing the relatively nonrefractory young Andean
The Senonianearly Paleocene underfilled foreland crust along preexisting zones of weakness (Mpodozis
basin of the Andes shows similarities with the and Kay, 1992). This dislocation was expressed in exten-
Western Interior seaway of the United States, sional and transtensional structures at the surface.
reflecting similarities in early Cordilleran-type Voluminous silicic magmatism started as early as
evolution. Pennsylvanian time in northern Chile (Breitkreuz et al.,
Phanerozoic Evolution of Bolivia and Adjacent Regions 227

1989). In the Peruvian Cordillera Oriental, a still poorly for their help and constant support. The manuscript benefited
understood but probably transpressional deformation of from comments by N. Eyles and P. E. Isaacson, and from
Late Permian age created an intracratonic foldbelt that editing work by A. J. Tankard. Thanks to J. Oller, E. Aguilera,
was subsequently intruded by calc-alkaline, posttectonic Y. Ardaya, G. Beccar, M. G. Bonhomme, R. F. Butler, G.
granites of EarlyMiddle Triassic age. This magmatism Carlier, L. Coudert, E. Daz-Martnez, G. Hrail, E. Jaillard, O.
evolved to tholeiitic composition at the same time as Lpez, R. Marocco, L. G. Marshall, D. Merino, G. Monte-
several grabens formed, indicating the onset of exten- murro, M. Prez, E. Requena, A. Sadud, E. Salinas, P. Soler, H.
sional stress fields in the late Middle Triassic. A subse- Toledo, E. Vargas, D. Zubieta, and many others for fruitful
quent reorganization of regional stresses, attributed to a discussions or additional information.
resumption of subduction beneath western Gondwana,
terminated rifting in the Late Triassic and resulted in
local transpression and basin inversion. This period was
followed by renewed extension in western Andean Peru REFERENCES CITED
and a Norian transgression in eastern Andean and sub-
Andean Peru. Allmendinger, R. W., V. A. Ramos, T. E. Jordan, M. Palma,
and B. L. Isacks, 1983, Paleogeography and Andean struc-
tural geometry, northwest Argentina: Tectonics, v. 2,
Latest JurassicRecent ( 145 m.y. duration) p. 116.
Archangelsky, S., ed., 1987, El sistema Carbonfero en la
Mild extensional conditions prevailed in Bolivia from Repblica Argentina: Academia Nacional de Ciencias,
Middle Triassic to Coniacian time, although local trans- Crdoba, 383 p.
pressional deformation occurred in the Late Triassic. Aspden, J. A., S. H. Harrison, and C. C. Rundle, 1992, New
Along the margin, TriassicAlbian transtensional condi- geochronological control for the tectonomagmatic
tions were subsequently replaced by transpression evolution of the metamorphic basement, Cordillera Real
(Jaillard, 1994). Emplacement of the coastal batholith of and El Oro Province of Ecuador: Journal of South
Peru began in middlelate Albian time (Soler and American Earth Sciences, v. 6, p. 7796.
Avila, W. A., 1992, El magmatismo cmbrico-ordovcico en
Bonhomme, 1990). Bolivia, in J. G. Gutirrez Marco, J. Saavedra, and I.
Six transgressions of marine to restricted marine Rbano, eds., Paleozoico Inferior de Ibero-Amrica:
waters are recorded: CenomanianTuronian, early Universidad de Extremadura, Spain, p. 241253.
Santonian, middle Campanian, early Maastrichtian, Bahlburg, H., 1990, The Ordovician basin in the Puna of NW
Danian, and late Miocene. The first five entered Bolivia Argentina and N Chile: geodynamic evolution from back-
from the north and the last one from the south. Develop- arc to foreland basin: Geotektonische Forschungen, v. 75,
ment of the last five was related to foreland basin 107 p.
dynamics. Bard, J. P., R. Botello, C. Martinez, and T. Subieta, 1974,
Compressional building of the central Andes started Relations entre tectonique, mtamorphisme et mise en
in the late Turonian or early Coniacian in the west and place dun granite ohercynien deux micas dans la
Cordillre Real de Bolivie (massif de Zongo-Yani): Cahiers
subsequently propagated discontinuously toward the Orstom, srie Gologie, v. 6, p. 318.
east. Thrust deformation reached the Paleozoic basin of Bell, C. M., 1987, The Late Paleozoic evolution of the Gond-
Bolivia in late Oligocene time and has since propagated wanaland continental margin in northern Chile, in G. D.
great distances through this succession in the south- MacKenzie, ed., Gondwana six: structure, tectonics, and
western part of the basin. Andean evolution is a conse- geophysics: American Geophysical Union, Geophysical
quence of convergence directly related to the break-up of Monographs, v. 40, p. 261270.
Pangea and subsequent plate motions. The style of defor- Benedetto, J. L., T. M. Snchez, and E. D. Brussa, 1992, Las
mation in the central Andes appears to be controlled by cuencas silricas de Amrica Latina, in J. G. Gutirrez
preexisting structures and by the thickness and litholo- Marco, J. Saavedra and I. Rbano, eds., Paleozoico Inferior
gies of basin fills. de Ibero-Amrica: Universidad de Extremadura, Spain,
p. 119148.
The productive Beni-Chaco basin of Bolivia is charac- Blakey, R. C., F. Peterson, and G. Kocurek, 1988, Synthesis of
terized by a twofold history: first, an early Paleozoic late Paleozoic and Mesozoic eolian deposits of the Western
Jurassic pericratonic Gondwana evolution during which Interior of the United States: Sedimentary Geology, v. 56,
time the principal source and reservoir rocks were p. 3125.
deposited, and second, a late OligoceneRecent Andean Bond, G. C., P. A. Nickeson, and M. A. Kominz, 1984, Breakup
episode of tectonic evolution and burial. of a supercontinent between 625 Ma and 555 Ma: new
evidence and implications for continental histories: Earth
and Planetary Science Letters, v. 70, p. 325345.
Bourgois, J., P. Huchon, and G. Pautot, 1990, Geologa de la
margen activa del Per entre los 3 y 12 de latitud Sur:
Acknowledgments This work is a result of an extensive Bulletin de lInstitut Franais dEtudes Andines, v. 19,
program on the geology of Bolivia initiated in 1983 and funded p. 241291.
by UR 1H of Orstom. Field and stratigraphic studies have been Branisa, L., G. A. Chamot, W. B. N. Berry, and A. J. Boucot,
conducted since 1984 in close collaboration with Yacimientos 1972, Silurian of Bolivia, in W. B. N. Berry and A. J. Boucot,
Petrolferos Fiscales Bolivianos (YPFB). I particularly thank the eds., Correlations of the South American Silurian Rocks:
Gerencia de Exploracin and staff and field geologists of YPFB GSA Special Paper 133, p. 2131.
228 Sempere

Breitkreuz, C., 1986, Das Palozoikum in den Kordilleren Central Andes: Lecture Notes in Earth Sciences: Berlin,
Nordchiles (2125S): Geotektonische Forschungen, SpringerVerlag, v. 17, p. 199208.
v. 70, 88 p. Grschke, M., A. von Hillebrandt, P. Prinz, L. A. Quinzio, and
Breitkreuz, C., 1991, Fluvio-lacustrine sedimentation and H. G. Wilke, 1988, Marine Mesozoic paleogeography in
volcanism in a Late Carboniferous tensional intra-arc northern Chile between 2126S, in H. Bahlburg, C.
basin, northern Chile: Sedimentary Geology, v. 74, Breitkreuz, and P. Giese, eds., The southern central Andes:
p. 173187. Lecture Notes in Earth Sciences, Berlin, SpringerVerlag, v.
Breitkreuz, C., H. Bahlburg, B. Delakowitz, and S. Pichowiak, 17, p. 105117.
1989, Volcanic events in the Paleozoic central Andes: Helwig, J., 1972, Stratigraphy, sedimentation, paleogeog-
Journal of South American Earth Sciences, v. 2, p. 171189. raphy, and paleoclimates of Carboniferous (Gondwana)
Caputo, M. V., 1991, Solimes megashear: intraplate tectonics and Permian of Bolivia: AAPG Bulletin, v. 56, p.
in northwestern Brazil: Geology, v. 19, p. 246249. 10081033.
Chandler, M. A., D. Rind, and R. Ruedy, 1992, Pangean Hrail, G., P. Baby, J. Oller, M. Lpez, O. Lpez, R. Salinas, T.
climate during the Early Jurassic: GCM simulations and Sempere, G. Beccar, and H. Toledo, 1990, Structure and
the sedimentary record of paleoclimate: GSA Bulletin, v. kinematic evolution of the sub-Andean thrust system of
104, p. 543569. Bolivia, in First International Symposium on Andean
Cobbold, P., A. C. Massabi, and E. A. Rossello, 1986, Geodynamics, Grenoble: Paris, Editions de lOrstom,
Hercynian wrenching and thrusting in the Sierras p. 179182.
Australes foldbelt, Argentina: Hercynica, v. 2, p. 135148. Isaacson, P. E., 1975, Evidence for a western extracontinental
Cordani, U., K. Kawashita, L. Baeza, M. Daz, and D. Evange- land source during the Devonian period in the central
lista, 1988, Geocronologa de la Sierra Limn Verde, Andes: GSA Bulletin, v. 86, p. 3946.
Antofagasta, Chile: Fifth Congreso Geolgico Chileno, Isaacson, P. E., B. Antelo, and A. J. Boucot, 1976, Implications
Santiago, v. 2, p. E63E74. of a Llandovery (Early Silurian) brachiopod fauna from
Cotkin, S. J., M. L. Cotkin, and R. L. Armstrong, 1992, Early Salta Province, Argentina: Journal of Paleontology, v. 50,
Paleozoic blueschist from the schist of Skookum Gulch, p. 11031112.
eastern Klamath Mountains, northern California: Journal Jaillard, E., 1994, Tectonic evolution of the Peruvian margin
of Geology, v. 100, p. 323328. between Kimmeridgian and Paleocene times, in J. A.
Dalmayrac, B., G. Laubacher, and R. Marocco, 1980, Carac- Salfity, ed., Cretaceous tectonics in the Andes: Earth
tres gnraux de lvolution gologique des Andes pru- Evolution Sciences Monograph Series, Wiesbaden, Vieweg
viennes: Travaux et Documents de lOrstom, Paris, v. 122, Publications, p. 101167.
501 p. Jaillard, E., and T. Sempere, 1991, Las secuencias sedimenta-
Damm, K. W., S. Pichowiak, R. S. Harmon, W. Todt, S. Kelley, rias de la Formacin Miraflores y su significado cronoes-
R. Omarini, and H. Niemeyer, 1990, Pre-Mesozoic tratigrfico: Revista Tcnica de YPFB, Santa Cruz, v. 12,
evolution of the central Andes; the basement revisited, in S. p. 257264.
M. Kay and C. W. Rapela, eds., Plutonism from Antarctica Jaillard, E., P. Soler, G. Carlier, and T. Mourier, 1990, Geody-
to Alaska: GSA Special Paper 241, p. 101126. namic evolution of the northern and central Andes during
Daz-Martnez, E., 1991, Litoestratigrafa del Carbonfero del early to middle Mesozoic times: a Tethyan model: Journal
Altiplano de Bolivia: Revista Tcnica de YPFB, Santa Cruz, of the Geological Society of London, v. 147, p. 10091022.
v. 12, p. 295302. Jaillard, E., V. Carlotto, J. Crdenas, R. Chvez, and W. Gil,
Dyck, M., 1992, Stratigraphy, fauna, sedimentology and 1993, La Nappe des Couches Rouges de Cuzco (Sud du
tectonics in the Ordovician and Silurian of east-Paraguay Prou): mise en vidence stratigraphique, interprtations
and the comparison with the Argentinian and Bolivian tectoniques et palogographiques: Comptes Rendus
Andes (abs.): Conferencia Internacional sobre el Paleozoico lAcadmie des Sciences de Paris, srie II, v. 316, p. 379386.
Inferior de Ibero-Amrica, Publicaciones del Museo de Kay, S. M., V. A. Ramos, C. Mpodozis, and P. Sruoga, 1989,
Geologa de Extremadura, Spain, v. 1, p. 6465. Late Paleozoic to Jurassic silicic magmatism at the
Farrar, E., A. H. Clark, and S. M. Heinrich, 1990, The age of Gondwana margin: analogy to the middle Proterozoic in
the Zongo pluton and the tectonothermal evolution of the North America?: Geology, v. 17, p. 324328.
ZongoSan Gabn Zone in the Cordillera Real, Bolivia, in Kennan, L., 1993, Cenozoic evolution of the Cochabamba area,
First International Symposium on Andean Geodynamics, Bolivia, in Second International Symposium on Andean
Grenoble: Paris, Editions de lOrstom, p. 171174. Geodynamics, Oxford: Paris, Editions de lOrstom,
Frana, A. B., and P. E. Potter, 1991, Stratigraphy and p. 199202.
reservoir potential of glacial deposits of the Itarar Group Keppie, J. D., 1993, Synthesis of Paleozoic deformational
(CarboniferousPermian), Paran basin, Brazil: AAPG events and terrane accretion in the Canadian
Bulletin, v. 75, p. 6285. Appalachians: Geologische Rundschau, v. 82, p. 381431.
Gagnier, P. Y., A. Blieck, C. C. Emig, T. Sempere, D. Vachard, Kontak, D. J., A. H. Clark, E. Farrar, and D. F. Strong, 1985,
and M. Vanguestaine, in press, New data on the Ordovi- The rift associated Permo-Triassic magmatism of the
cian and Silurian of Bolivia: Journal of South American eastern Cordillera: a precursor to the Andean orogeny, in
Earth Sciences. W. S. Pitcher, M. P. Atherton, J. Cobbing and R. D. Beckin-
Gayet, M., T. Sempere, H. Cappetta, E. Jaillard, and A. Lvy, sale, eds., Magmatism at a plate edge: the Peruvian Andes:
1993, La prsence de fossiles marins dans le Crtac London, John Wiley, p. 3644.
terminal des Andes centrales et ses consquences Laubacher, G., 1978, Gologie de la Cordillre Orientale et de
palogographiques: Palaeogeography, Palaeoclimatology, lAltiplano au nord et nord-ouest du lac Titicaca (Prou):
Palaeoecology, v. 102, p. 283319. Travaux et Documents de lOrstom, Paris, v. 95, 217 p.
Gtze, H. J., S. Schmidt, and S. Strunk, 1988, Central Andean Lobo-Boneta, J., 1975, Sobre algunos palinmorfos del
gravity field and its relation to crustal structures, in H. Devnico superior y Carbnico inferior de la zona
Bahlburg, C. Breitkreuz, and P. Giese, eds., The Southern subandina sur de Bolivia: Revista Tcnica de YPFB, v. 4,
p. 59176.
Phanerozoic Evolution of Bolivia and Adjacent Regions 229

Lpez, O., G. Montemurro, and H. Trujillo, 1982, Estratigrafa de la Regin de Antofagasta: 4th Congreso Geolgico
del Paleozoico medio en las Serranas de San Jos y Chileno, Antofagasta, v. 1, p. 439457.
Santiago de Chiquitos, Bolivia: 5th Proc. Congreso Lati- Oller, J., and T. Sempere, 1990, A fluvio-eolian sequence of
noamericano de Geologa, Buenos Aires, v. 1, p. 283292. probable Middle TriassicJurassic age in both Andean and
Marquillas, R. A., and J. A. Salfity, 1988, Tectonic framework sub-Andean Bolivia, in First International Symposium on
and correlations of the CretaceousEocene Salta Group, Andean Geodynamics, Grenoble: Paris, Editions de
Argentina, in H. Bahlburg, C. Breitkreuz, and P. Giese, lOrstom, p. 237240.
eds., The southern central Andes: Lecture Notes in Earth Prez, M., 1990, Miospores du Dvonien moyen et suprieur
Sciences, Berlin, SpringerVerlag, v. 17, p. 119136. de la coupe de Bermejo-La Angostura (sud-est de la
Marshall, L. G., and T. Sempere, 1991, The Eocene to Pleis- Bolivie): Annales de la Socit Gologique de Belgique, v.
tocene vertebrates of Bolivia and their stratigraphic 113, p. 373389.
context: a review, in R. Surez-Soruco, ed., Fsiles y Facies Racheboeuf, P. R., A. Le Hriss, F. Paris, C. Babin, F. Guil-
de Bolivia: Revista Tcnica de YPFB, Santa Cruz, v. 12, locheau, M. Truyols-Massoni, and R. Surez-Soruco, 1993,
p. 631652. Le Dvonien de Bolivie: Biostratigraphie et chronostrati-
Marshall, L. G., T. Sempere, and M. Gayet, 1993, The Petaca graphie: Comptes Rendus lAcadmie des Sciences de
(late Oligocenemiddle Miocene) and Yecua (late Miocene) Paris, srie II, v. 317, 795802.
formations of the sub-AndeanChaco basin, Bolivia, and Ramos, V. A., 1988a, Late Proterozoicearly Paleozoic of
their tectonic significance: Documents du Laboratoire de South Americaa collisional history: Episodes, v. 11,
Gologie de Lyon, France, v. 125, p. 291301. p. 168174.
Martinez, C., 1980, Structure et volution de la chane hercyni- Ramos, V. A., 1988b, The tectonic of the central Andes,
enne et de la chane andine dans le nord de la Cordillre 3033S latitude, in S. P. Clark, Jr., B. C. Burchfiel, and J.
des Andes de Bolivie: Travaux et Documents de lOrstom, Suppe, eds., Processes in continental lithospheric deforma-
Paris, v. 119, 352 p. tion: GSA Special Paper 218, p. 3154.
McGregor, D. C., 1984, Late Silurian and Devonian spores Ramos, V. A., and S. M. Kay, 1991, Triassic rifting and associ-
from Bolivia: Academia Nacional de Ciencias de Crdoba, ated basalts in the Cuyo basin, central Argentina, in R. S.
v. 69, p. 157. Harmon and C. W. Rapela, eds., Andean magmatism and
Mgard, F., 1978, Etude gologique des Andes du Prou its tectonic setting: GSA Special Paper 265, p. 7991.
central: Travaux et Documents de lOrstom, Paris, v. 86, Rapela, C. W., B. Coira, A. Toselli, and J. Saavedra, 1992, El
310 p. magmatismo del Paleozoico inferior en el Sudoeste de
Merino, D., 1987, Los conodontos permocarbonferos de la Gondwana, in J. G. Gutirrez-Marco, J. Saavedra, and I.
Formacin Copacabana en Yaurichambi, Departamento de Rbano, eds., Paleozoico Inferior de Ibero-Amrica:
La Paz: Tesis de Grado de la Universidad Mayor de San Universidad de Extremadura, Spain, p. 2168.
Andrs, La Paz, 135 p. Rodrigo, L. A., A. Castaos, and R. Carrasco, 1977, La
Merino, D., 1991, Primer registro de conodontos silricos en Formacin Cancairi, sedimentologa y paleogeografa:
Bolivia: Revista Tcnica de YPFB, Santa Cruz, v. 12, Revista de Geociencias de la Universidad Mayor de San
p. 271274. Andrs, La Paz, v. 1, p. 122.
Merino, D., and J. Blanco, 1990, Conodontos de la Formacin Roeder, D., 1988, Andean-age structure of eastern Cordillera
Copacabana (Carbonfero superior-Prmico inferior) en la (Province of La Paz, Bolivia): Tectonics, v. 7, p. 2339.
seccin de Huarachani-Pacobamba, Departamento de La Rsler, O., R. Ianuzzi, and R. Surez-Soruco, 1989, A flora
Paz, Bolivia: Revista Tcnica de YPFB, Cochabamba, v. 11, carbonfera da Formao Kasa em Beln, pennsula de
p. 109116. Copacabana, Altiplano boliviano, e a importancia das
Meyer, H. A. O., and Villar, L. M., 1984, An alnoite in the formas trifoliadas: 11th Congresso Brasileiro de Paleon-
Sierras Subandinas, northern Argentina: Journal of tologia, Curitiba, Abstracts, p. 4546.
Geology, v. 92, p. 741751. Saunders, W. B., and W. H. C. Ramsbottom, 1986, The mid-
Moya, M. C., 1988, Lower Ordovician in the southern part of Carboniferous eustatic event: Geology, v. 14, p. 208212.
the Argentine eastern Cordillera, in H. Bahlburg, C. Sempere, T., 1989, Paleozoic evolution of the central Andes
Breitkreuz, and P. Giese, eds., The southern central Andes: (1026 S): 28th International Geological Congress,
Lecture Notes in Earth Sciences, Berlin, SpringerVerlag, Extended Abstracts, Washington, D.C., v. 3, p. 73.
v. 17, p. 5569. Sempere, T., 1990, Cuadros estratigrficos de Bolivia: Prop-
Mpodozis, C., and S. M. Kay, 1992, Late Paleozoic to Triassic uestas nuevas: Revista Tcnica de YPFB, Cochabamba, v.
evolution of the Gondwana margin: evidence from Chilean 11, p. 215227.
frontal Cordilleran batholiths: GSA Bulletin, v. 104, Sempere, T., 1991a, Evolucin de la cuenca centro-andina
p. 9991014. (1026 S) del Cmbrico superior al Silrico inferior:
Mukasa, S. B., and D. J. Henry, 1990, The San Nicols Revista Tcnica de YPFB, Santa Cruz, v. 12, p. 221223.
batholith of coastal Peru: early Paleozoic continental arc or Sempere, T., 1991b, Cenozoic tectonic phases in Bolivia:
continental rift magmatism?: Journal of the Geological some needed clarifications: 6th Congreso Geolgico
Society of London, v. 147, p. 2739. Chileno, Via del Mar, v. 1, p. 877881.
Muoz, N., and R. Charrier, 1993, JurassicEarly Cretaceous Sempere, T., 1994, Kimmeridgian? to Paleocene tectonic
facies distribution in the western Altiplano (182130 evolution of Bolivia, in J. A. Salfity, ed., Cretaceous
S.L.)implications for hydrocarbon exploration, in Second tectonics in the Andes: Earth Evolution Sciences
International Symposium on Andean Geodynamics, Monograph Series, Wiesbaden, Vieweg Publications,
Oxford: Paris, Editions de lOrstom, p. 307310. p. 168212.
Murphy, J. B., and R. D. Nance, 1991, Supercontinent model Sempere, T., G. Hrail, J. Oller, and P. Baby, 1989, Geologic
for the contrasting character of Late Proterozoic orogenic structure and tectonic history of the Bolivian orocline: 28th
belts: Geology, v. 19, p. 469472. International Geological Congress, Extended Abstracts,
Niemeyer, H., F. Urza, F. G. Aceolaza, and C. Gonzlez, Washington, D.C., v. 3, p. 7273.
1985, Progresos recientes en el conocimiento del Paleozoico
230 Sempere

Sempere, T., G. Hrail, J. Oller, and M. G. Bonhomme, 1990a, Surez-Riglos, M., M. A. Hnicken, and D. Merino, 1987,
Late Oligoceneearly Miocene major tectonic crisis and Conodont biostratigraphy of the Upper
related basins in Bolivia: Geology, v. 18, p. 946949. CarboniferousLower Permian rocks of Bolivia, in R. L.
Sempere, T., G. Hrail, J. Oller, P. Baby, L. Barrios, and R. Austin, ed., Conodonts: Investigative Techniques and
Marocco, 1990b, The Altiplano: a province of intermontane Applications: Amsterdam, Elsevier, p. 316322.
foreland basins related to crustal shortening in the Bolivian Surez-Soruco, R., 1976, El sistema ordovcico en Bolivia:
orocline area, in First International Symposium on Andean Revista Tcnica de YPFB, v. 5, p. 111223.
Geodynamics, Grenoble: Paris, Editions de lOrstom, Surez-Soruco, R., and J. M. LpezPugliessi, 1983, Formacin
p. 167170. Saipur, nuevo nombre formacional para representar a los
Sempere, T., P. Baby, J. Oller, and G. Hrail, 1991, La napa de sedimentos superiores del Ciclo Cordillerano (Devnico
Calazaya: una prueba de acortamientos importantes superior-Carbnico inferior): Revista Tcnica de YPFB,
controlados por elementos paleoestructurales en los Andes v. 9, p. 209213.
bolivianos: Revista Tcnica de YPFB, Santa Cruz, v. 12, Toro, M., C. Vargas, and R. Birhuet, 1992, Los trilobites
p. 229234. ashgillianos de la Formacin Cancairi (regin de Milluni,
Sempere, T., E. Aguilera, J. Doubinger, P. Janvier, J. Lobo, J. Cordillera Real, departamento de La Paz): 10th Congreso
Oller, and S. Wenz, 1992, La Formation de Vitiacua Geolgico Boliviano, Extended Abstracts, La Paz,
(Permien moyen suprieurTrias ?infrieur, Bolivie du p. 188190.
Sud): stratigraphie, palynologie et palontologie: Neues Vavrdov, M., P. E. Isaacson, E. Daz-Martnez, and J. Bek,
Jahrbuch fr Geologie und Palontologie, Abhandlungen, 1991, Palinologa del lmite DevnicoCarbonfero en torno
v. 185, p. 239253. al Lago Titikaka, Bolivia: resultados preliminares: Revista
Shackleton, R. M., A. C. Ries, M. P. Coward, and P. R. Tcnica de YPFB, Santa Cruz, v. 12, p. 303313.
Cobbold, 1979, Structure, metamorphism and Veevers, J. J., 1988, Gondwana facies started when Gondwana
geochronology of the Arequipa Massif of coastal Peru: merged in Pangea: Geology, v. 16, p. 732734.
Journal of the Geological Society of London, v. 136, Veevers, J. J., and C. McA. Powell, 1987, Late Paleozoic glacial
p. 195214. episodes in Gondwana reflected in transgressive-regres-
Smith, M. T., W. R. Dickinson, and G. E. Gehrels, 1993, sive depositional sequences in Euramerica: GSA Bulletin,
Contractional nature of DevonianMississippian Antler v. 98, p. 475487.
tectonism along the North American continental margin: Vicente, J. C., B. Beaudoin, A. Chvez, and I. Len, 1982, La
Geology, v. 21, p. 2124. cuenca de Arequipa (Sur Per) durante el Jursico-
Soler, P., and M. G. Bonhomme, 1987, Donnes Cretcico inferior: 5th Congreso Latinoamericano de
radiochronologiques K-Ar sur les granitodes de la Cordil- Geologa Proceedings, Buenos Aires, v. 1, p. 121153.
lre Orientale des Andes du Prou central: implications Zubieta-Rossetti, D., P. Huyghe, G. Mascle, J. L. Mugnier, and
tectoniques: Comptes Rendus lAcadmie des Sciences P. Baby, 1993, Influence de lhritage ant-dvonien au
de Paris, srie II, v. 304, p. 841844. front de la chane andine (partie centrale de la Bolivie):
Soler, P., and M. G. Bonhomme, 1990, Relations of magmatic Comptes Rendus lAcadmie des Sciences de Paris, srie
activity to plate dynamics in central Peru from Late Creta- II, v. 316, p. 951957.
ceous to present, in S. M. Kay and C. W. Rapela, eds.,
Plutonism from Antarctica to Alaska: GSA Special Paper Authors Mailing Address
241, p. 173191.
Soler, P., and T. Sempere, 1993, Stratigraphie, gochimie et Thierry Sempere
signification palotectonique des roches volcaniques Consulting Geologist
basiques msozoques des Andes boliviennes: Comptes 13 rue Geoffroy lAngevin
Rendus lAcadmie des Sciences de Paris, srie II, v. 316, 75004 Paris
p. 777784. France

You might also like