Você está na página 1de 46

Journal of South American Earth Sciences,Vol. 1, No. 4, pp.

373-418, 1988
Printedin Great Britain

0895-9811/88 $3.00+ 0.00


PergamonPressplc

Cretaceous paleogeography and depositional cycles of


western South America
C. E . MACELLARI
Earth Sciences and Resources Institute, University of South Carolina, Columbia, SC 29208, USA*
(Received for publication J a n u a r y 1988)

A b s t r a c t - - T h e western margin of South America was encroached upon by a series of marine advances t h a t
increased in extent from the Early Cretaceous to a maximum in the early Late Cretaceous for northern
South America (Venezuela to Peru}. In southern South America, however, the area covered by the marine
advances decreased from a maximum in the Early Cretaceous to a minimum during mid-Cretaceous time,
followed by a widespread advance at the end of the period. A series of unconformity-bounded depoeitional
cycles was recognized in these sequences: five cycles in northern South America, and six (but not exactly
equivalent) cycles in the Cretaceous back-arc basins of southern South America (NeuquSn and Austral, or
Magallanes, Basins). Both widespread anoxic facies and maximum flooding of the continent in northern
South America coincide in general terms with recognized global trends, but this is not the case in southern
South America. Here, anoxic facies are restricted to the Lower Cretaceous and seem to be controlled by
local aspects of the basin evolution and configuration. The contrasts observed between northern and
southern South America can be explained by differences in tectonic setting and evolution. To the north,
sediments were deposited around the tectonically stable Guayana-Brazilian Massifs, and thus registered
global "signals" such as anoxic events and major eustatic changes. The southern portion of tim continent,
on the contrary, developed in an active tectonic setting. Here, the mid-Cretaceous Peruvian Orogeny overprinted, to a large extent, world-wide trends and only the earliest and latest Cretaceous conform to global
depositional patterns.
R e s u m e n b E l margen oeste de Sudam~rica fue cubierto per una sorie de avances maritm~ tlUe involucraron ~ireas progresivamente mhs extensas a partir del CretAcico temprano hasta alcanzar uu Inaximo en
el Cretkcico tardio temprano en el norte del continente (Venezuela a Per0). En el sur de Sudttmbrica, sin
embargo, el area cubierta per las transgresiones cretbcicas decreci6 a partir de un m6ximo en el CretAcico
m6s temprano hasta el Cret~cico medio, pero una extensa transgresibn cubri6 una gran parte del continente al finalizar este periodo. Varies cielos deposicionales soparados per sendas discordaneias regionales
se reconocieron en estas secuencias: cinco ciclos para el norte de Sudam~rica y seis ciclos (aunque no
exactamente equivalentes) en las cuencas de tras-arco cret4tcicas del sur de Sudam6rica. Tanto facies
an6xicas de gran extensibn, come la edad en que se verifie6 la m6xima inundacibn del norte de Sudam~rica
coinciden en t~rminos generales con tendencias globales, pore eete no es el case on el sur del continente.
Aqtti, las facies an6xicas estAn restringidas al Cret~cico int'erior y estAn aparentemente controladas per
aspectos pecuiiares de la evoluci6n y configuraci6n de dichas cuencas. Los contrastes observados entre el
norte y el sur de Sudam6rica pueden ser explicados per diferencias en su ubicacibn y evoluci6n tectbnicas.
En el norte, los sedimentos fueron depositaries alrededor de los macizos de Guayana y Brasil, tectbnicamonte estables, que registraron eventos globales, tales come episedios an6xieos y eustaticos. Per el contrario, la porcibn sur del continente se desarroll6 en un ambiente tect6nico active. Aqui, la orogenia
Peruviana del CretAcico medio se sobreimpuso en gran medida a los eventos globales, y solamonte durante
el Cret~cico mbs temprano y el m6s tardIo los sedimentos se depositaron de acuerdo a patrones deposicionales globales.

INTRODUCTION

THE WESTERNMARGINof South America was covered


by several successive marine advances during Cretaceous time. Sediments deposited during these events
are widely exposed along the Andes and preserved in
the subsurface of subandean basins. Cretaceous rocks
in these basins are of great economic importance as
they provide the source and, in many cases, the
reservoirs for the largest accumulations of hydrocarbons of the western margin of the continent.
The objectives of this work were to recognize common patterns of basin infill along the extensive continental margin, and to contrast and possibly explain
different depositional trends observed in northern

*Present address: Pecten International. P.O. Box 205, Houston, TX 77001 USA

.qq.q

and southern South America. These tasks were approached in two ways: the first involved the reconstruction of detailed paleogeographic maps; the
second included recognition and correlation of unconformity-bounded depositional cycles or sequences
of regional significance (sensu Vail et al., 1977).
This study was based mostly on published information, field work in Venezuela, Colombia, and
southern Argentina and Chile, and limited unpublished subsurface information for Colombia and portions of Argentina.
Various criteria were used for recognition of the
cycles. In surface outcrops, cycles usually present
three distinctive intervals or surfaces. A transgressive deposit, usually containing reworked phosphatic
fragments, iron nodules, or reworked fauna from underlying sediments, is found at the base of many
sequences. This basal unit, deposited on top of the
"transgressive surface," indicates the initiation of a

374

C.E. MACELLARt

relatively rapid rise in sea level and may be present


on top of the lowstand deposits in the deepest portion
of the basin or directly above the sequence boundary
on the updip portion of the basin (Haq et al., 1987;
Vail, 1987). The second surface is the maximum
flooding surface, which is usually associated with a
condensed zone. Since the basin is deprived at this
time of clastic input due to the increased distance
from the source of sediments, sedimentation is characterized by chemical deposits (glauconite), high
concentrations of pelagic material, or the development of hardgrounds due to the absence of sedimentation (Haq et al., 1987). The third recognizable
surface in outcrop is the sequence or cycle boundary,
which is represented by an obvious unconformity or
by more subtle changes. In the deeper portion of the
basin, the sequence boundary is conformable and is
characterized by a change from progradational deposits to more massive aggradational sequences (Haq
et aI., 1987). Other criteria used for the recognition of
these cycles include paleontological hiata, marked
changes in the map distribution of coastal onlap of
sediments, and basin-wide marked changes in lithology.
These cycles are the result of the interaction of tectonic and thermal subsidence, variations in sedimentary input, and eustatic sea-level fluctuations. Even
though no attempt has been made to evaluate the
origin of these cycles, they are comparable in magnitude to the second-order cycles of Haq et al., (1987).
The degree of certainty about the exact definition of
these cycles varies from well documented to speculative. However, it is hoped that this kind of analysis
may provide a regional framework to understand and
predict local aspects of the Cretaceous stratigraphy of
western South America.

High; Fig. 1) of major importance in the migration


and entrapment of hydrocarbons on a regional scale.
The greatest thickness of Cretaceous sediments has
been reported for the Bogot~ Trough (over 15,000

DISTRIBUTION OF CRETACEOUS
SEDIMENTS
Areal Distribution

Cretaceous sediments were deposited over a large


portion of western South America in a variety of
tectonic settings. Even though these sediments were
deposited in the Pacific (or "active") South American
margin, the Guayana and Brazilian Massifs provided
a stable platform for sediments deposited to the east
of these basins. To the west, however, sedimentation
during the Early Cretaceous took place mostly in a
back-arc setting that changed to a foreland basin
setting during the Late Cretaceous, following the
uplift and deformation of the ancestral Andes.
Depositional troughs in northern South America
are oriented NE/SW, coinciding with the distribution
of Jurassic red beds in rapidly subsiding grabens (cf.
Maze, 1984). These are the Trujillo, Machiques, Uribante, and Bogot~ Troughs (Fig. 1). Also controlling
the distribution of these sediments was a series of
NW/SE-trending paleohighs (M6rida Arch, Arauca
Arch, Santander Massif, and Vaup6s-Natagaima

Fig. 1. Generalizedisopach map of CrvLuceoussedimunLsof"the


westernmarginof SouthAmericashowingmajorpaleohighs.

Cretaceous paleogeography and depositional cycles of western South America

375

meters cited by BOrgl, 1961e). However, only half


this value was measured in the area recently (E.
Cardozo and L. Sarmiento, pets. commun., 1987).
South of the Vaup~s High, Cretaceous isopachs
trend NW/SE, parallel to the orientation of major
tectonic features. The most prominent of these tectonic features is the Marah6n Geantieline that
separated the East Peruvian and the West Peruvian
Troughs during a large portion of the Cretaceous
(Benavides, 1956).
In southern Peru, Bolivia, and northern Argentina
and Chile, Cretaceous sedimentation was restricted
to a series of relatively isolated but rapidly subsiding
basins (Reyes, 1972; Salfity, 1982). South of this
area, a major NW/SE-trending paleohigh (the "Crat6geno Central" of Bracaccini, 1960) separated the
Chaco-Paran~ Basin from the Neuqu~n Basin. In
southernmost South America, the Rio Chico or Dungeness High was also oriented NW/SE and separated
the San Jorge Basin from the Austral (Magallanes)
Basin to the south.

Classification of Basins
Cretaceous rocks are preserved in several basins of
western South America that became isolated during
the Cenozoic. These Cenozoic basins preserving Cretaceous strata are classified here as follows (Fig. 2).

Oceanic Basins. These are basins in which Cretaceous sediments were deposited on top of oceanic
crust that was later obducted onto the continent.
Included in this group are basins located in northwestern and western Colombia and in western Ecuador. These rocks have been subjected to a variable
degree of metamorphism (Bourgois et al., 1987;
among others).
Proximal Pericratonic Basins. This extensive system of basins envelops the western margin of the
Guayana-Brazilian Massifs. Because of their location to the east of the Andes, these have also been
called "subandean" basins. Cross-basinal arches
separate this area into a series of basins, namely:
Barinas-Apure, Llanos, Putumayo, Oriente, Marafi6n, Ucayali, and Madre de Dios. Sedimentation
here is characterized by the dominance of clastic
intervals. These basins have great economic significance as they contain extensive hydrocarbon reserves.
Distal Pericratonic Basins. These basins represent
the western continuation of the pericratonic basins
from which they became separated during the late
Cenozoic Andean orogeny. In general, the distal
pericratonic basins are characterized by finer
grained sediments and much thicker sequences than
those found to the east. Thick intervals of shale and

~.
2
".
:

. / ~ vo ' "
.~o

...i

"..

oooOO
ooOO

BASIN
CLASSIFICATION

PASSIVE MARGIN

~ OCEANIC
P
ROXlNJ~.
PERICRATONIC
~ D
ISTAL
PERICRATONIC
INTRA-CRATONIC

~ ACTIVE
MARGIN
~ BACK-ARC
~[ ~

SEDIMENTARYINPUT

Fig. 2. Classificationof pressnt-day majorsedimentarybasins


on the basisofCretaceousrockspresent.

:.

6001Gr

AU8TRAL

o,

376

C.E. MACELLARI

limestone are common here. The presence of excellent source rocks for hydrocarbons and the scarcity of
Cretaceous elastic reservoirs are characteristic of
these basins. Basins included within this group are
the Maracaibo, Middle Magdalena, Upper Magdalena, Santiago, and Huallaga.

buco, and Rio Negro Formations. To the west, the


basal elastics (Tambor Formation) are followed by
more distal marine sediments of the Rosa Blanca
Formation (Julivert, 1968; Zambrano et al., 1971;
Gonzalez de Juana et al., 1980; Garcia Jarpa et al.,
1980; among others).

Intracratonic Basins. A major NW/SE paleogeographic high separated the pericratonic area around
the Brazilian Massif from the Cretaceous back-arc
basins of southern South America. Cretaceous sedimentation within this high was characterized by
mostly continental deposits capped during the latest
Cretaceous by a short-lived marine transgression.
Both sources and reservoirs of hydrocarbons are
associated with this brief event. Basins included in
this group are the Titicaca Basin of southern Peru,
the Subandean Basin of Bolivia, and the Northwest
Basin of Argentina.

Ecuador and Peru. Well-documented earliest Cretaceous rocks in this portion of South America are
primarily restricted to the West Peruvian Trough. A
continuous belt of fluvial to deltaic facies was deposited adjacent to the western edge of the Marafi6n
Geanticline (Chimu, Etuancane, and Cotacucho Formations) (Dalmayrac et al., 1980; among others).
Progressively deeper water facies have been found
toward the center of the basin, with a mostly shaley
deposition in the northern portion and limestone predominant to the southwest (Santa Formation) (Benavides, 1956). Farther south, predominantly silicielastic deposition took place in a shallow marine
environment (Yura Group) (Vicente et al., 1982), and
active volcanism occurred along the coast in a strip
extending from Lima to Nazca (Myers, 1974).

"Active Margin Basins." These basins developed


at the edge of the South American plate. Because of
their position near the zone of convergence of the
Nazca and South American plates, these basins are
characterized by complex tectonics and marked facies
variations. Included in this group are the Talara
Basin of northwestern Peru (Olsson, 1934; Travis et
al., 1976) and the Navidad fore-arc basin of coastal
Chile (Cecioni, 1979; among others).
Back-Arc Basins. The Neuqu~n and Austral Basins underwent a similar tectonic evolution, both
developing as back-arc basins during Late JurassicEarly Cretaceous times. However, the Neuqu~n and
Austral Basins differ in that the former contains
thick "middle" Cretaceous continental deposits,
whereas the latter basin registers marine sedimentation throughout the Cretaceous. In both basins,
sources and the most important reservoirs of hydrocarbons accumulated during the early stages of the
Cretaceous transgression.
Passive Margin Basins. These basins developed on
the Atlantic margin of South America as the continent rifted apart from South Africa. These basins
are not considered in this study, with the exception of
brief references for comparative purposes.

PALEOGEOGRAPHY
The Cretaceous paleogeographic evolution of western South America is summarized in Figs. 3-8.
Pre-Barremian (Fig. 3)
Venezuela and Colombia. Sedimentation during
earliest Cretaceous time was restricted to the major
depocenters of the area - - namely, the BogotA, the
Uribante, and, possibly, the Machiques Troughs.
These sediments are characterized by quartz and
feldspathic arenites included in the Caqueza, Area-

Argentina and Chile. During the BerriasianHauterivian interval, the back-arc basins of Argentina and Chile were covered by an extensive marine
advance that began in the Tithonian. In the Neuqu6n Basin, deep basinal limestones and shales of the
Vaca Muerta Formation were followed by slope
sediments of the Quintuco Formation and shallower
water carbonates and elastics of the Loma Montosa
Formation (Mitchum and Uliana, 1985; among
others). Reefal facies (Chachao Formation) developed to the northeast of the basin (Legarreta et al.,
1981).
Extension in the back-arc portion of the Austral
Basin resulted in the intrusion of ophiolitic bodies
(Dalziel et al., 1974; among others). Lowermost Cretaceous sediments include a basal sandstone (Springhill Formation) followed by black shales (ltio Mayer
and Pampa Rinc6n Formations). These sediments
are replaced farther north by shallow marine sandstone (Apeleg or basal Coyhaique Formations) (Ploszkiewicz and Ramos, 1978; Skarmeta and Charrier,
1976).
At the northern part of the passive margin, the
Chaco-Paranfi Basin is characterized by deposition of
clean eolian sands (Tacuaremb6, San Crist6bal,
Botucat~ Formations), overlain by extensive tholeiitic basalt flows (Russo et al., 1980). These volcanics
are indicative of the initial extension prior to the
opening of the Atlantic Ocean. Farther south, the
San Jorge Basin recorded widespread lacustrine and
fluvial sedimentation (Pozo D-129 Formation) (Lesta
et al., 1980a). A series of smaller continental basins

Fig. 3. Pre-Barremianpalvogvographicmap. Data for northern South Americafromseveral authoru;data Ibr euuthern
South America modifiedfrom Malumi~n et al. {1983) and
Riccardi {1987).

Cretaceous paleogeography and depositionai cycles of western South America

377

associated with alkaline rocks developed along the


"Crat6geno Central Argentino."
Barremian-Cenomanian (Fig. 4)

+
/

"

../
. :
u.-,., ]
..f

/
~COTACUCHO

)
/

c, j

'I
1

PRE-BARREMIAN
MARINE
CHERT/GREYWACKE/
OPHIOLITE
~ 1
~

LIMESTONE (SHALE)
SHALE (LIMESTONE)
SANDSTONE-SHALE
SANDSTONE

CONTINENTAL
SANDSTONE-SHALE
~--]
V

NO DEPOSITION
ANDESITES
BASALTS

ALKALINE ROCKS

GRANITES

~"

OPHIOLITES

APPROXIMATE SCALE
100 0

500km.

Venezuela and Colombia. Cretaceous sediments


extended over an increasingly larger area during the
Barremian. A characteristic feature of these rocks
include the presence of evaporites interbedded in different environmental settings (Julivert, 1968; Garcia
Jarpa et al., 1980; among others). Thick elastic facies
of the Rio Negro Formation were deposited in the
Trujillo, Machiques, and Uribante Troughs. In the
Guajira Peninsula, however, quiet marine deposition
predominated (limestones of the Yuruma Formation). During this time, the M6rida Arch (which
separates the Trujillo and Uribante Troughs) and the
Santander Massif remained as positive features. In
Colombia, sandstones were deposited east of BogotA,
but they were replaced by fine-grained sediments
(Villeta and Paja Formations} to the west.
During the late Aptian-Albian, a large portion of
the continental margin was blanketed by quartz
sandstones derived from the Guayana Massif (Aguardiente, Une, and Caballos Formations). Those sandstones were deposited in a deltaic to shallow marine
environment, and covered the Putumayo and Upper
Magdalena Basins as well as the Santander Massif
and the M6rida Arch for the first time. The sandstone belt was replaced to the west by finer grained
sediments that include carbonatic intercalations
(Villeta, Simiti, and Lisure Formations). Turbidites
and basic volcanics were possibly the dominant rocks
deposited in western Colombia.
Ecuador and Peru. During Aptian-early Albian
times, the Oriente Basin orEcuador as well as a large
portion of Peruvian territory, including the MarafiSn
Geanticline, were covered by quartz sandstone derived from the Guayana and Brazilian Massifs (Fig. 4).
These sediments were flushed westward through an
extensive fluvio-deltaic system (ltolltn Formation in
Ecuador and Cushabatay, Goyllarisquizga, and Farrat Formations in Peru). In Ecuador, coeval sediments were possibly deposited to the west in the now
metamorphic portion of the Cordillera Real (Faucher
and Savoyat, 1973). Farther west, in Ecuador, an
andesitic volcanic arc was active from the Aptian to
the Late Cretaceous. This volcanic arc probably rested on oceanic crust to the north but on continental
crust to the south (llenderson, 1979}. Still farther
west, the Romeral Fault separated the rocks of the
volcanic arc from pillow basalts of the PifiSn Formation (Fig. 5). The Pifibn Formation has been
interpreted as the distal, oceanic portion of the volcanic arc (Henderson, 1979) or, alternatively, as an
accreted portion of oceanic crust (Feininger and
Bristow, 1980; Feininger, 1982; F e i n i n g e r and
Seguin, 1983). In southern Peru, non-marine red bed
deposition predominated during this time (Murco,
Huancane, and Cotacucho Formations). The only
Aptian-early Albian, well-defined marine facies were

378

C.E. MACELLARI

~.-:--:;

..... /
N

i)

i,/

V ~
V/

APTIAN-ALBIAN
MARINE

CHERT/GREYWACKE/
OPHIOLITES

! LIMESTONE (SHALE)

SHALE (LIMESTONE)

ANDSTONE-SHALE
SANDSTONE

VOLCANIC ARC

CONTINENTAL
~
SANDSTONE-SHALE
V

ANDESITES

BASALTS

GRANITES

APPROXIMATE SCALE
100 0

500 km.

deposited in the Lima area where the Farrat sandstones are replaced by marine shales and limestones
of the Pamplona Formation.
During the later part of the Albian, Cretaceous
seas invaded most of Ecuador and Peru (Fig. 5). In
the Oriente Basin, this event is reflected in the basal
portion of the Nape Formation. In Peru, the new
transgression is reflected in the marine shales of the
Esperanza-Raya Formation of the Marafibn, Ucayali,
and Huallaga Basins of Peru. F a r t h e r east and
south, however, these sediments are replaced by deltaic facies included in the Agua Caliente Formation.
Of special interest during this time interval is the
development of widespread anoxic conditions in
western Peru. These are evidenced in the bituminous limestones of the Pariatambo Formation (central
and northeastern Peruvian Andes) and the Muerto
limestone of northwestern Peru. To the south, as
well as to the east, these anoxic facies are replaced by
better oxygenated, mostly calcareous sediments.
These are the Ferrobamba (Abancay) and Arcurquina Formations to the south (Marocco, 1978) and
the Crisnejas Formation to the east in the MarafiSn
Geanticline (Benavides, 1956).
During the Albian, red bed sedimentation (Cotacucho Formation) continued in southern Peru, as
well as in the Altiplano area. These facies are replaced to the west by red shale and gypsum deposits
of the Moho Formation and by shallow marine limestones and quartzites of the Arcurquina Formation
(Newell, 1949; Audebaud et al., 1976; Dalmayrac et
al., 1980).
The late Albian marine advance was followed by a
general regression in Ecuador and Peru (Fig. 6). This
regression is clearly reflected in the East Peruvian
Trough where most of the area was covered by a elastic wedge (Agua Caliente Formation) derived from
the Guayana and Brazilian cratons. Anoxic conditions had completely disappeared by the latest
Albian. Predominantly shale and marl facies (Pulluicana Group) enclose a limestone facies (Jumasha
Formation) deposited east of the West P e r u v i a n
Trough and on top of the MarafiSn Geanticline. Marine volcanic activity continued in the coastal area
(Casma Group), but at a much less impressive pace
than during middle Albian time.
Marine sedimentation continued in southern Peru
during the late Albian-Cenomanian. This is evidenced in the shales and limestones of the Ferrobamba and Arcurquina Formations (Fig. 6). A shortlived marine incursion occurred during Cenomanian
times in the Altiplano area (Ayavacas Limestone of
the Moho Formation). Farther east, however, red bed
deposition of the Cotacucho facies continued, with
the exception of a thin dolomite intercalation that
may correlate with the Ayavacas event (Laubacher,
1978).
Fig. 4. Aptian-Albian paleogeographic map. Data for northern
South America from several author~; data for southern South
America modified from Malumi~n et aL (1983) and Riccardi
(1987).

Cretaceous paleogeography and depositional cycles of western South America

379

LATE MIDDLE A L B I A N
I

(
/
/

BRAZIL

Trujillo~
:~.Z,:=
~': "=

;.~, o~:_
=..- o ,,

"" '

0
:*%

",
i,'. ,.,.~ Shale-limestone

---,.

:;i

rr''-T"~
i
II

i
u

I n
1

~
Bituminous limestone

;:

Evaporites
Red beds

__

Andesite-volcaniclastics
Basic lavas & volcaniclastics
No deposition

80

0
I

KM
I

300
t

76

Fig.5. Late middleAlbian paleogeographyofEcuador and Peru.


A r g e n t i n a a n d Chile. In Barremian-Albian times,
the Neuqu~n and Austral Basins developed differently. The Neuqu~n records the presence of progressively shallower sedimentation that ended with the

deposition ofevaporites (Huitrin Formation) and continental red beds (Rayoso Formation). Concurrent
active andesitic volcanism took place to the west in
Chile (Aberg el al., 1984; Ramos and Ramos, 1979).

C.E. MACELLARI

380

LATE ALBIAN-CENOMANIAN
I

Gua'

iiiiiii~::

j..f: :
/

BRAZIL

It

Sandstone

"~",~" """~

;'",~
Limestone
Limestone-shale
'..'~
Turbidites
.,

=';"

Shale

Hxxet

K',."

,.~

Red beds

'E.

Andesite-volcaniclastics
Basic volcanics & volcaniclastics
No deposition
80

0
I

KM
t

I
76W

300
t
72

Fig. 6. Late Albian-Cenomanian paleogeography of Ecuador and Peru.

Marine sedimentation continued in the Austral


Basin (Rio Mayer, lower Pampa RincSn, Hito XIX,
Nueva Argentina, and Vicufia Formations), with the
exception of continental tuffaceous sediments (Divisadero Formation) deposited to the north. These
rocks were affected by a mid-Aptian tectonic event

that resulted in the first uplift of the Andes to the


west (Ramos, 1976; Aguirre Urreta and Ramos,
1981). This event coincided with extensive plutonism (Su~rez, 1979; Aguirre, 1985) and volcanic activity (Ramos, 1978) in the southern Andes.

Cretaceous paleogeography and depositional cycles of western South America


In the Northwest Basin, conglomeratic sedimentation took place in the block-faulted Alemania SubBasin where several basalt flows were extruded in
association with the earlier phases of sedimentation
(Reyes, 1972; Valencio et al., 1976). All the eastern
passive margin basins (Atlantic) are characterized by
continental sedimentation possibly initiated in Barremian-Albian times (Rio Salado Formation, Salado
Basin, Fortin Formation, Colorado Basin) (Lesta et
al., 1980b; Urien and Zambrano, 1973; Zambrano,
1980).

381

~ESOANDALO8A

~::.:::.;.:::..

~ADALUPE
MAGUGHI

./'
....... l',.'" '".

T u r o n i a n - S a n t o n i a n (Fig. 7)

.......v,. /
Venezuela a n d Colombia. The maximum extension of the Cretaceous transgression in northern
South America occurred during Turonian-Santonian
times, coinciding with a worldwide highstand of sea
level (Kauffman, 1979; Jenkyns, 1980; Haq et aI.,
1987). Uplift of the Cordillera Central of Colombia
also commenced during this interval as a result of the
mid-Cretaceous (Peruvian) orogeny (Steinmann,
1929; Btirgl, 1961a; Campbell and B0rgl, 1965;
Julivert, 1968; Irving, 1975). Three depesitional provinces are recognized between this semi-emergent
arc and the South American craton: a coastal area
composed of proximal clastic facies (Escandalosa and
eastern facies of the Guadalupe Formation); a shelf
belt composed of intercalations of sandstone, shale,
porcellanite, and phosphorite (Navay and western
facies of the Guadalupe Formation); and a pelagic
belt (outer shelf-slope), characterized by organic-rich
black shale, thinly bedded bituminous limestone, and
chert (La Luna and Villeta Formations) deposited in
anoxic conditions (Catchcart and Zambrano, 1967;
Gonz~tlez de Juana, et al., 1980; Macellari and
DeVries, 1987).
Thick turbidite packages were deposited during
Turonian-Santonian times at the edge of the South
American plate. Examples are found exposed along
the Cordillera Occidental of Colombia (Barrero,
1979) and at the northern termination of the Andes
(Barquisimete Trough, NE portion of Fig. 7; Renz,
1960b; Stephan, 1977).
Ecuador a n d Peru. Turonian-Coniacian sedimentation in the Oriente Basin is recorded in the limestones and shales of the Napo Formation (Fig. 7). In
the southwestern portion of Ecuador and in northwestern Peru, the Copa Sombrero Formation represents turbiditic sedimentation in a narrow trough
(Lancones syncline and its northward extension;
Fisher, 1956; Morris and Aleman, 1975}. A thick
prism of coeval sediment that had been winnowed
from the shelf was laid down by turbidity currents on
the deeper continental slope to the west (Feininger

600Ktl

JUMASHA-%

FERROBAMB

TURONIANSANTONIAN

TINENTAL
.
.DEPOSIT?
CHERT, OPHIOUTES
& TURBIDITES
~

,
("

BLACK SHALE (BI- ~,


TUMINOUS LIME- (

STONE, CHERT)
TURBIDITES

SANDSTONE-SHALE ~ . 1 r"
(PORCELLANITE) ~ A R o m m -

SANDSTONE

~o~'~,~, MATA I
~#'LL"
I
MO--

Fig. 7. Turonian-Santonianpaleogeographicmap. Data for


northern South America from several authors; data for
southern SouthAmericamodifiedfromMalumi6net al. (1983)
and Riccardi(1987).
SAE$ I/4--E

BASALTS

382

C.E. MACELLARI

and Bristow, 1980). The andesitic volcanic arc continued its activity during Turonian-Coniaeian times
(C61ica-Macuchi Formations), and a thick volcaniclastic sequence was deposited in the Costa region
(Cayo Formation) (Bristow, 1975).
In the East Peruvian Trough, the marine advance
is reflected in the shales and limestones of the
Chonta Formation. This unit grades toward the
craton into a deltaic system that includes pro-delta
muds and delta-top facies (proto-Amazon delta; Soto,
1979). Farther west, the Chonta Formation becomes
more limestone-rich and is replaced by the Jumasha
and the Cajamarca Formations. These limestone
facies are continued south in the Ferrobamba and
Arcurquina Formations. Limestone facies were partly replaced by shale deposition during the Santonian.
The Chonta Formation in this interval is mostly
composed of shales and is continued westward into
the marls of the Celendin Formation. A thick evaporitic sequence was deposited in southern Peru
during the Santonian (Querque and Chilcana Formations; Jenks, 1948; Vicente, 1981). Farther east,
in southern Peru, these evaporitic facies are replaced
by the red beds of the upper portion of the Moho and
Cotacucho Formations (Newell, 1949; Audebaud et
al., 1976). The Peruvian movements initiated during
the Santonian are reflected in the uplift of a large
portion of the coastal area of Peru.

Argentina and Chile. Turonian to early Campanian sedimentation in the Neuqu~n and Austral
Basins was controlled by the initial uplift of the
Andes, which provided an important western source
of sediments. The Neuqudn Basin records exclusively continental sedimentation during this time
(Neuqu~n Group). The intrusion of granitic rocks
was extensive along northern Chile; this was associated with widespread andesitic volcanism immediately to the east (cf. Aguirre, 1985).
In the Austral Basin, marine sedimentation continued after a depositional hiatus that extended
approximately from the late Cenomanian to the early
Coniacian (Malumi#~n, 1968; Malumi~n etal., 1971;
Flores etal., 1973). This hiatus is possibly related to
the extensive magmatic activity (Ramos, 1976;
Ploszkiewicz and Ramos, 1978) and intense folding
and uplift associated with the closure of the marginal
basin that gave rise to the "Paleoandes" to the west
(Cecioni, 1957; Dalziel et al., 1974; Winslow, 1980).
Sedimentation in the center of the basin, however,
seems to have been interrupted only by a much
shorter hiatus (Biddle e t a L , 1986). A rapidly
subsiding foredeep developed immediately to the east
of the newly uplifted cordillera, where a turbiditic
sequence was deposited (Punta Barrosa and Cerro
Toro Formations), but platform deposition continued
to the east (Palermo-Aike Formation). Regressive
facies are well displayed in the northern portion of
the Austral Basin (Mata Amarilla Formation), possibly replaced northward by continental pyroclastic
deposits (Cardiel Formation).

Continental sedimentation continued in the


Northwest Basin, this time extending to the rest of
the sub-basins (Pirgua Subgroup), but the Salto
Jujeiia Dorsal, located in the center of the basin,
remained as a positive feature (Salfity, 1982).
During the Coniacian-Campanian interval, the
Atlantic basins were also characterized by continental deposition. However, marine facies are recorded
in the eastern portion of the Colorado Basin (Zambrano, 1974, 1980; MalumiAn et al., 1983).

Campanian-Maastrichtian (Fig. 8)
Venezuela and Colombia. A marked regression
occurred during the Campanian-Maastrichtian interval in northwestern South America. Particularly
important was the uplift of a portion of the Cordillera
Central following the Peruvian movements. For the
first time, then, two sources of sediments are observed: an eastern source provided by the Guayana
Massif, and a more restricted western source provided by the ancestral Cordillera Central. Sediments
include sandstone to the east and south (Monserrate,
Guadalupe, and Burguita Formations) and shale to
the west and north (Umir and Col6n Formations).
The elastic influence decreases to the northwest
where a limestone-shale sequence p r e d o m i n a t e s
(Guaralamai Formation) (Renz, 1960a; Rollins,
1965). Turbidites were deposited during this time
interval in western and northwestern Colombia
(Duque-Caro, 1984).
Ecuador and Peru. In western Ecuador, uplift
associated with shortening and deformation of the
back-arc basin resulted in metamorphism in the
present Cordillera Real (Feininger, 1975, 1982). In
the Oriente Basin, sediments were derived from two
sources. To the east, a fluvial system transported
quartz sands from the Guayana and Brazilian
Massifs (Vivi~n Formation). In the southwestern
part of the Oriente Basin and in the northwestern
part of the Marafi6n Basin, these sandstones are
followed by the Maastrichtian Cachiyacu shales,
which were deposited during a brief but widely
distributed marine transgression. These facies were
replaced to the west by a red bed sequence of mostly
continental origin that was derived from the newly
uplifted Andes (Tena Formation). These sediments
truncate progressively younger beds in an eastward
direction. Thus, the red bed facies appeared in the
Santonian in the Cordillera Occidental of Peru
(Chota and Casapalca Formations), but only in the
latest Maastrichtian in the Oriente (Tena Formation) and Ucayali Basins (Huchpayacu Formation)
(Tschopp, 1953; Benavides, 1956; Wilson, 1963;
Huerta-Kohler, 1982).
Fig. 8. Campanian-Maastrichtianpaleogeographicmap. Data
fromnorthernSouthAmericafroms~veralauthors; data from
southernSouth AmericamodifiedfromMalumiAnetal. (1983)
and Riccardi(1987).

Cretaceous paleogeography and depositional cycles of western South America

In central Ecuador, a trough developed to the west


of the Cordillera Real where the M a a s t r i c h t i a n
turbidites of the Yunguilla Formation were deposited
(Bristow, 1973; Feininger and Bristow, 1980). In the
Costa area, deeper-water facies persisted during this
period (chert of the Guayaquil Member of the Cayo
Formation; Sigal, 1969; Bristew, 1975).
Possibly the same marine advance reflected in the
Cachiyacu shales of the Marafi6n Basin is recorded in
the Petacas Formation of the Talara Basin (Cruzado
Castafieda, 1980). By the end of the Maastrichtian,
most of Ecuador and Peru had become emergent.

.)

Argentina and Chile. An extensive marine advance over large portions of southern South America
took place during the Maastrichtian (cf. Uliana and
Biddle, in press). For the first time, an Atlanticderived transgression covered a large portion of
Argentina, and the Andes began to develop as the
continental divide. This marine ingression is recorded in the Neuqu~n Basin (Jaguel and Malargue
Formations) and in its eastern continuation into the
Colorado Basin (Pedro Luro Formation). Regressive
facies punctuated by oscillations and m a r i n e advances were deposited in the Austral Basin during
this time (Arbe and Hechem, 1984b; Macellari, et al.,
in review). This resulted in the deposition of deltaics
(La Anita Formation) to the northwest, and inner
and outer shelf rocks to the south (Cerro Cazador and
Fuentes Formations).
A shallow marine transgression also reached the
Northwest Basin (Lomas de Olmedo Sub-Basin)
where the oolitic and stromatolitic limestones of the
Yacoraite Formation were deposited (Marquillas et
al., 1984, among others). This short-lived transgression coincided with a similar episode observed in
southern Peru (Vilquechico Formation) and Bolivia
(El Molino Formation) where a transgression had
possibly entered from the Atlantic Trough into the
Chaco Paran~ Basin (cf. Salfity et al., 1985).
In Chile, a rapidly subsiding fore-arc basin developed along the western edge of the active (Pacific)
m a r g i n (Navidad Basin) (Cecioni, 1979; Bir6Bagoczky, 1982; Stinnesbeck, 1986). Extensive andesitic volcanism took place in the arc to the east of
this basin.
Continental sedimentation continued in the San
Jorge Basin, but the new marine advance is recorded
in other passive margin basins, the Colorado, Salado,
and Chaco-Paran~ Basins (Malumi~n et al., 1983).

CAMPANIANMAASTRICHTIAN
MARINE
m
CHERT, SHALE,
VOLCANICS

LIMESTONE-SHALE
~

SHALE

TURBIDITES

SANOSTONE-SHALE

SANOSTONE

383

e#
CONTINENTAL
~
SANDSTONE-SHALE

i CERRO
FORTALEZA

RED BEDS
I

I NON-DEPOsITION
v

LA

VOLCANICS
APPROXIMATE SCALE
100 0

500km.

CAZAOOR

C R E T A C E O U S CYCLES
Several regional unconformity-bounded cycles
have been recognized in the Cretaceous of western
South America (Figs. 9 and 10). The stratigraphy of
the pericratonic and back-arc basins is discussed for
three different major areas: northern South America
from Venezuela to Peru, the Neuqu~n Basin of
Argentina, and the Austral Basin of Argentina and
Chile. The "oceanic" basins of northwestern South

384

C.E. MACELLARI

[2 i:i':::"
HAUTERI[VlAN
IVALANGIIIIA'~

~
i: I

CHERT--PORCELLAHI~E
: i u~esTo.E
s v ~

~
~

81LTSTONE
8ANOSTONE

MULICHINC?~
-,- .-_ ----- _

I BERR/AS/A"I "l

Fig. 9. Stratigraphicchart and correlationofcyclesin the pericratonicand back-arcbasinsofwesternSouthAmerica.


America, as well as the "intracratonic" basins of central South America, are not included in this discussion because of insufficient data for this kind of
analysis in the former case, and lack of precise age
dating in the latter.
Northern South America
Five depositional cycles of regional distribution
preserved in the pericratonic to distal pericratonic
basins are recognized in northern South America.
The stratigraphy of this area is discussed in some
detail as this is the first time that these cycles are
documented on a regional scale. Regional stratigraphic sections depicting these cycles are presented
in Figs. 11-17.
Venezuela and Colombia. The five depositional
cycles found in Venezuela and Colombia are detailed
below.
Cycle I (Tithonian to Valanginian): Sedimentation during the earliest portion of the Cretaceous
was restricted to localized NE/SW-trending grabens
that originated during initial extension in a rifting
stage of evolution of the area (Stainforth, 1969;
Favre, 1983a,b). No break in sedimentation is
observed at the Jurassic/Cretaceous boundary. The
exact age and limits of this cycle are poorly defined
due to the lack of detailed biostratigraphic data to
correlate between isolated exposures.

The initial marine advance entered through the


Bogota area and progressed rapidly to the northeast
(B(irgl, 1961c; Etayo-Serna, et al., 1976; Favre, 1985).
In the eastern portion of the Cordillera Oriental of
Colombia, sedimentation started with deposition of
the Caqueza Group. This group includes up to 250
meters of a poorly sorted basal conglomerate followed
by 2500 meters of shales and siltstones with intercalations of orthoquartzites and limestones that indicate the initial transgressive episode (Hubach, 1945;
Renzoni, 1967; Miller, 1979). The group contains
diagnostic Tithonian to Valanginian ammonites.
Farther north along the NE/SW-trending graben,
in the Sierra Nevada del Cocuy, the marine advance
is recorded in the Macanal Shale, which was deposited rapidly under inner shelf conditions (Favre,
1985; Fig. 14). To the east, this sequence is replaced
by a coarser clastic sequence (Rio Ele Conglomerate).
To the west, in the Villa de Leiva area, similar clastic
facies are included in the braided fluvial and delta
plain sandstones of the Arcabuco Formation. This
unit is followed by the shallow marine shales of the
Cumbre Formation (Galvis and Rubiano, 1985; Renzoni, 1985b).
Although precise age dating is lacking, it is possible that Cycle 1 rocks may have been deposited in
the Uribante Trough of southwestern Venezuela (the
basal portion of the Rio Negro Formation; see Fig. 1).
This unit comprises arkosic sandstones of considerable thickness (Renz, 1959; Ramirez and Campos,
1972; Garcia Jarpa et al., 1980).

Cretaceous paleogeography and depositional cycles of western South America

385

DISTAL PERICRATONIC eASlHS


/

AGE

aAnACAm01UA.ALENAGmxne0T I.P.U

MAAS-

--

TR/CHTIAN

i.",~ ,',,..-.'..'~-:
--

UMIR

"-: . ' . ' . ".-.-.'-...-'."

"

__ __ __ __

_: - I

.Nc..

--LA

LUNA-

""--r'-

-r--'-r

-I

GALEMBO!-'_

~_.~..~.:~

Z-

--

,~ PUJAMA-.

~ CAPACHO

, "

.
.~i

C A P A
. . . .

"-'.

_"-_C.,~l,~ql2~l-"

-.

I..,,i ........................................

MAN/AN

....

---.

SIMITI.

:" : ' : : " " - : ' - :'.: ::: - T A

-QUILLQIJ~-

~.................. ?l ...........

' --

A~IIARDIFNTI~
.
"-'~ "'" "~':
- ~

BLAZO

.*

, , , , i L
"pin LLqCANA,

. ' " : " .'.'." : " . " " ~" .--L A ; J ' i - ~ _ ~ - :" C ~ " "
. . . . .

r'l::.: .. . . . . .

:..-.."~.IN

CA'.-~

"!:":.'.:::.:
:.~." : . : : ; . . - - . , ,

ses s,

..;,.,;.
;, .--"
~
. ; - ~ ". ;~~' .. -. .- ~
_ o . , , ~ , _ I....................... ,'

- _- -I.-:,-.-:-.,..:---~-

~---~r-----~--"-

TURONIAN

,pmN

"-...'--

::--:'.. " . ' - - . ' . :

--__--__--__--__-I~'~--__~_~__~"! :".I.';:~..'.'::.':~.."._.C~OTA.--"
.'.
--~P-I-T-~v

SANTONIAN

CENk-

--

TIE '" '" : .---..'" ".

....... ,T,,,,,T,,,,,T,,,,,,T,,,,T,,,,,T,,,,,,T,,............. ;;,;, ".,;,:.,


, , :~').
,, . . . . . .
- .

C~U~A~.

co

. . . .

- COLON--

~.~ __ __ ---

;-:-

- - - - ~

, .--:_~.;.

Fig. 10. Stratigraphic chart and correlation of cycles in the distal pericratenic basins and western Peru; see Fig. 7 for explanation
of symbols.

Cycle 2 (Valanginian to Upper Aptian): Subsidence during Cycle 2 deposition changed from being
confined and tectonically driven during the earlier
part of the cycle, to being more widespread and
thermally driven at the end of the cycle (Fabre,
1983a,b).
In Venezuela, deposition of the Rio Negro Formation continued in the Uribante Trough and was
initiated in the Machiques Trough (Figs. 1, 12, 13).
This is a time-transgressive unit that displays a
major variation in thickness - - from 2000 meters in
the Machiques Trough to only a few meters in the
platform area of the Maracaibo Basin and up to 1500
meters in the Uribante Trough. The formation is
composed of polygenic conglomerate, with sandstone
and light grayish claystone intercalations with some
evaporitic intervals, and contains a restricted shallow water bivalve assemblage (Renz, 1959; Richards,
1968; Garcia-Jarpa et al., 1980).
The flooding of the continental margin during
Cycle 2 deposition resulted in the accumulation of

limestone and shale over a large area (lower Cogollo


Group, and Ap6n Formation; Figs. 12 and 13). [n the
Sierra de Perij~ (Machiques Trough), these rocks are
followed by the Machiques Member of the Ap6n Formation, representing the maximum flooding event
during Cycle 2. These are well-laminated, bituminous, dark blue-gray limestone with abundant calcareous concretions and a predominant pelagic fauna
deposited in an anoxic environment (Renz, 1959,
1981). This unit was paleontologically dated as late
Aptian (Sutton, 1946).
The Ap6n Formation is much thinner in the Maracaibo Platform and comprises well bedded, micritic,
skeletal limestones interbedded with marly limestones deposited in a neritic,well aerated but quiet
environment (Renz, 1981; Fig. 12). Toward the east
and southeast in the Andes, the Ap6n Formation
displays a progressive terrigenous influence and a
rapid decrease in thickness from that found in the
Machiques Trough (Trump and Salvador, 1964;
Garcia-Jarpa et al., 1980). These rocks are capped by

C.E. MACELLARI

386

700

760

12 o

-12 0

...-. . . . .

-IOON

.9 ....

,,,, ~ , ."'"

~"

MARACAIBO
8A SIN

......~. -,

/:..:'

.-

~'..""
==~14

!
%
H
..-..

BA RINA

B A SIN

-.

.....
..,,

=-, /.?
m.=..

COLOMBIA""...
vAD A

Y-

~
~,, ~

%/
~

"..,

.'
Sogamoso--!
~

i-..~
32

~ ~-I

8o~ot~'~"

~ardot ~

,.,,

o ~ - , -

.:..

,.o

_..y.-'

.=

.j

~: ,~..
...

-"

- Pasto

.
. .....

~ " 4 s ' - - ~ . .

:~.'.

....-,,I'
~ ....

.., - . ' ' i : :

LOCATION OF
STRATIGRAPHIC
SECTIONS

i.~'":,..

r ~ ......

PUTUMAYO
BA -I "

'' '

~..

..

~'~
~-"

"";"

WELL

42

SURFACE

SE CTION

:__-~.
4,.= T ......
*,o .... J
2':

-~

SECTION

O GENERALIZED
SECTION
WITH AVERAGE THICKNESS
0 o_

_.~
"~o
76 o

I
Fig. 11. Major

~ \...~L . .

72 o

74 ~

70Ow

physiographic features of western Venezuela and Colombia showing the locations ofstrutigraphic sections.

upper Aptian, gray, partly micaceous, sandy shale


approximately 20 meters thick (Guaimaros Shale),
deposited during the peak of Cycle 2 (Renz, 1959,
1981; Fig. 13).
In the Sierra Nevada del Cocuy area, the basal
sand of Cycle 2 (Las Juntas Formation) thickens to

up to 2122 meters and spans the Hauterivian to


lower Aptian interval (Fabre, 1985; Fig. 14). These
sediments are followed by approximately 130-140
meters of upper Aptian calcarenites and mudstones
of the ApSn Formation (Etayo-Serna, ]985; Fabre,
1985).

APT|AN

~LBIAN~

CAMPANIAN~IAASTRICHTIAN
TURONIAN-=r
SA N TONIAI~J_~

i ~- -

in

O.

~2

_ COLON

=t

~ARACAIBO
BASIN

FK

ilACHIOUES
TROUGH
,

.~URE

'1

"t

~0

1~

20 30 40 80kin.

UMESTONE

DEPOSITIONALCYCLE

8ANDSTONE

8HALE

FM.--

MARL ""////'~/ ~/'//~


CALCARENITE

"COLON

SUCEOUS FACIESI ~
BLACK SHALE
~

MNOxlC)

:-

_1

11
f2

II

$E
~A RINA S
BA SIN

IACIANTONIAN ?
ONIAN
)MANIAN
[AIM

PANIAN;TRICHTIAN

iECTION A

TROUGH

ANDES

rENEZUEL A N

Fig. 12. S'~ratigraphic section A, modified from Renz (1981); see Fig. 11 for location of section,

"

;IERRA DE
PERIJA

SW MARACAIBO
BA SIN

c~
oo

=I
~o

>

c~e

-<

C)

IALANGINIAN" I
HAUTERIVIAN

|ARRENIAN F

kPTIANF

4LBIAN |

:ENONANIAN/"

TURONIAN-~
~ANTONIAN

28

|~

MIDDLe

FM. L

IIA G DA L ENA
VALLEY

rABLAZO

- UMIR FM.

40O

SANTANDER
MASSIF

--

-~-----:

18

LIMESTONE

I0

20

30

40

81LICEOUS FACIES
~
mm SLACK SHALE (ANOXIC) ~
SHALE
~

60km.

DEPOSlTIONAL CYCLES

SANDSTONE

MARL
CALCARENITE

MERIDA
,, ARCH L

26

; O L O N FM.-4
--- z4 .---I

~~__-__--_[---_~_-_-_-__-_-__-__-_~_____________.___
_______,~_ t3

SECTION B

- -

fENEZUELAN A N D E S

"ig. 13. Stratigraphic section B, modified from Julivert ( 19~8 ~and Renz ~1981 ); see Fig. 11 for location ofsection.

....

-I

MIDDLE
MA GDALENA VALLEY

rRUJILL 0
TROUGH

~6--2T -

:8

IE

~RE-APTIAN

kPTIAN

,LIIIAN

'~ENONANIA N

'UEONIAN o
ANTONIAN

Z
, <

:)0
~0

Cretaceous paleogeography and depositional cycles of western South America

389

East of BogotA, Cycle 2 was initiated with a trans- Uribante Trough. Four members are recognized in
gressive quartz sandstone (Caqueza Sandstone of this formation but only the lower two are included in
Miller, 1979; or the Alto de Caqueza Formation of Cycle 3. Member S, composed of a micaceous, carRenzoni, 1967; see Fig. 3). The Caqueza Sandstone is bonaceous shale, forms a regional, 6-24 meter thick
Hauterivian in age (Hubach, 1957; Biirgl, 1961a) and guide unit that represents the m a x i m u m flooding of
is followed by the shales of the Fomeque Formation Cycle 3 transgression (Russomano and Velarde,
that represent the peak of Cycle 2 sedimentation. 1982). This is followed by Member R, a glauconitic
This unit comprises 700 meters of dark gray, iron- interval with intercalations of crossbedded sandstone
rich, occasionally gypsiferous silty shale with inter- and minor dark shale (approximately 40 m thick)
calations of calcareous, micaceous siltstone, fine- that resulted from the progradation of shallow
grained, carbonaceous sandstone, and hard, fine- marine clasticsduring the final portion of Cycle 3.
In the southern Venezuelan Andes Basin, Cycle 3
grained orthoquartzite (Renzoni, 1967).
was initiated with the basal transgressive sand of the
In the Middle Magdalena Valley, sedimentation
started with the Tambor Formation (Figs. 10, 14, 15), lower to middle Albian Aguardiente Formation
which rests unconformably on Jurassic red beds. (Notestein et al., 1944; Figs. 12 and 13). South of the
M6rida Arch, this formation is composed of 300 to
This unit includes 350 to 650 meters of conglomerates and quartz sandstones interbedded with red 500 meters of light colored, hard, medium to thickly
and gray mudstones that were deposited in several bedded sandstone interbedded with carbonaceous
continental to shallow marine settings ranging from shale and siltstone, with glauconite becoming more
braided to high-sinuosity rivers, alluvial plain, and abundant up-section (Trump and Salvador, 1964).
tidal fiat (Bueno, 1979; Renzoni, 1985a). The age of North of the M6rida Arch, equivalent beds contain
thick intercalations of calcareous sandstone (Peflas
the basal deposits is generally accepted as Valanginian-Hauterivian (Morales et aL, 1958; Julivert, Altas Formation; Renz, 1959; Fig. 13). The Pe6as
Alias Formation represents a tidally dominated,
1961, 1968; Bueno, 1979; Taborda, 1979). Well
established marine conditions were achieved during highly destructive delta with long strings of sand
deposition of the overlying unit, the Rosa Blanca
perpendicular to the coast (Bartok et al., 1981). The
Formation. This unit is composed of up to 400 meters m a x i m u m flooding event of Cycle 3 in the southern
Venezuelan Andes is recorded in the anoxic facies of
of thinly bedded, medium- to dark-gray wackestones,
micrites, biomicrites, grainstones and biomicrudites, the La Grita M e m b e r of the Capacho Formation
representing shallow marine to intertidal and supra- (Renz, 1959). This member (upper Albian to lower
tidal deposits (Taborda, 1979; Cardozo and Ramirez, Cenomanian) is composed of 10-15 meters of well
1985). Ammonites reported from this formation in- bedded black limestones that may represent a condicate a Hauterivian-Barremian age (Morales et a[., densed zone at the peak of Cycle 3. The upper portion
1958).
of Cycle 3 in this area is represented by the lower
portion of the Seboruco shales of the Capacho ForThe upper portion of Cycle 2 in the Middle Magdalena Basin area is composed of 125 to 625 meters of mation (Renz, 1959, 1977; Figs. 12 and 13).
A progressive increase in limestone is observed toblack, thinly laminated, frequently micaceous and
slightly calcareous shales commonly containing gyp- ward the Maracaibo Platform during Cycle 3. These
sum intercalations (Paja Formation; Julivert, 1958; sediments are included in the Mercedes Member of
Reyment, 1981; Forero and Sarmiento, 1985). the Aguardiente Formation, in the upper ApSn,
Barremian to Aptian ammonites are reported for this Lisure, and Maraca Formations, and in the undifferentiated Cogollo Group, r e p r e s e n t i n g mostly
unit (Morales eta[., 1958; Julivert, 1968).
shallow marine sedimentation (Rod and Maync,
Cycle 3 (Upper Aptian to Mid-Cenomanian): 1954; Richards, 1968; Renz, 1977, 1981; Bartok et al.,
Deposition during this cycle occurred at a time of 1981). These are followed by the La Aguada Member
generalized subsidence controlled by the thermal of the La Luna Formation, which is composed of 60
cooling of an increasingly rigid crust (Fabre, 1983b). meters of dark gray massive limestone with large
This, combined with a high stand of sea level, resul- calcareous concretions and black shale that contain
ted in the flooding of several paleohighs that had upper Albian to Cenomanian a m m o n i t e s (Renz,
remained emergent during the earlier portion of the 1981). Cycles 3 and 4 are poorly differentiated in this
rapidly subsiding area.
Cretaceous.
In the Llanos Basin of Colombia, Cycle 3 was
Sedimentation in the Barinas Basin was initiated
with deposition of the Aguardiente Formation (Figs. initiated with the deltaic Une-Ubaque sandstones
9 and 12), a unit 55-93 meters thick composed of (Fig. 14). The Une is composed of approximately 500
glauconitic, well bedded calcareous sandstones inter- meters of quartz sandstone with siliceous cement,
calated with well laminated shales and gray, crys- quartz conglomerate, and medium- to fine-grained
talline limestone (Gaenslen, 1962). An Aptian?- clayey sandstone with carbonized debris (e.g., P~rez
Albian age has been determined for this unit in the and Bolivar, 1985). This unit is replaced in the north
Barinas Basin (Russomano and Velarde, 1982). This by the lithologically similar Ubaque F o r m a t i o n
(Miller, 1979).
sequence is overlain by the Cenomanian-Turonian
Toward the west, these sediments are replaced by
Escandalosa Formation, which is 55-262 meters
thick with m a x i m u m values increasing toward the deeper water facies preserved in the western margin

13C

ALBIAN

95" CENONANIAN

TURONIAN

ren~S
L' ~

ill

![

: ...i';.._TA___r~

. . . . .

~
"---

"

oI

:'.,......':'~'-'.::~:.:.:','.'.

,oI

: . . :.'~': : ' - . " " ' . . ' ,

:':~"~-~"._.,

,oo,..
I

e"

CANO
(3ARZA
34

SILICEOUS FACES

SECTION

(~

8HALE WITH COAL ~


8HALE

MARL

S-IOA
32

31

S'9

DEPOSITIONAL CYCLE

SANDSTONE

~
/IMFRT(~NF
I1 ~ II . . . . . . . . .

"

I ,..

RONDON
33

LLANOS BASIN

~
BLACK SHALE
~iiiiiiii~ (ANOXIC)

~-~

_ - - ~ ' " ~ ' U P I ( F M . " . " : ' ; " . ~

SANTA
MARIA
38

".'..y!E.R .N~ .S.S.. - ?.:':./~..:.-~

. . . . . . , Z'-.: ::':,..-:....

-" ~A~C_--------

II

Fig. 14. Stratigraphic section C, modified from Morales et al. (1958) and Fabre (I985); see Fig. 11 for location of section.

FM.:':.<:.:.'.:.:..

-~-.--~.~::~.-.!.:.:-.::i:-:.:

_ . . . . .

SIERRA
DEL COCUY
3B

E A S T E R N CORDILLERA

. . . . . . . . . . . . . . . .

:-----u~-__-_

5"

.......

VALLEY

MIDDLE
MAGDALENA

ESE

PICA PICO
30

C~

C~

GO
D

600

lO00m

--

"--'

GUADUAS

FM_.

8HALE

~-~

I]
~ u~lMARL
1-1--4
F_'--~Z-1DEEP WATER SHALE
(PARTLY ANOXIG)

,['t-r"1"! LIMESTONE

81LICEOUS FACIES

'ti~~ ~ ~ '~--r/:-~

--"

NEIVA SUB-BASIN
f
H

II

SECTION

-_-t_-

:~i~
;

PAJA

FM.

- -

I
37

_S_~:TTF~.__"

TABLAZO

....

MIDDL E MA GDA L E NA BA S I N

Fig. 15. Stratigraphic section D; see Fig. 11 for location of section.

lOOkm.

DEPOSITIONAL CYCLe'

60

8ANDSTONE

SANO$TONE ANO SHALE

~
~

-~SHALE AND COAL

"/'

~ ~ :

: a-~-

~CALCARENiTE

UPPER MA GDA L ENA BA S I N

VALANGINIAN

HAUTERIVIAN

BARREM|AN

APTIAN

ALBIAN

CENOMANIAN

TURONIAN

CONIACIAN

$ANTONIAN

CAMPANIAN"
MAAETRICHT|AN

AGE

~=~

~D

>

t=~

~C

=_

t=--

C3
P'I

392

C . F,. M A C E L L A R I

of the Cordillera Oriental and in the Middle Magdalena Valley. The basal unit is the Tablazo Limestone, composed of 150 to 325 meters of massively
bedded, fossiliferous limestone and marls (Figs. 13,
14, 15). The age of this formation is defined as late
Aptian-early Albian on the basis of its position in the
sequence (Morales et al., 1958). The maximum
flooding during Cycle 3 in the Middle Magdalena
Valley is recorded in the Simiti Shale, consisting of
gray to black locally calcareous shale and interstratified thinly to medium bedded, dark, fossiliferous limestone deposited in at least partially anoxic
conditions. The Simiti Shale ranges from 250 to 650
meters in thickness and contains Albian ammonites
(Morales et aL, 1958). There is a basinward progradation of shallower water deposits (Salto Limestone)
at the end of this cycle, possibly resulting from the
development of stillstand conditions. This unit is
composed of 50 to 125 meters of dark gray, argillaceous limestone with numerous thin intercalations of
black, calcareous shale (Morales et al., 1958;
Maugham et al., 1979; Taborda, 1979). Macrofauna
present in this interval indicate a late Albian to
Cenomanian age (Morales et aL, 1958).
West of BogotA, Cycle 3 sediments include a portion of the monotonous shales of the Villeta Group
(Hubach, 1957; Julivert, 1968). The base of the cycle
is marked by the turbiditic sands of the lower SocotA
Formation (Polania and Rodriguez, 1978), possibly

BURDINE-I

ACAE-I

CAIMAN-6
J ~ -----~- ?

representing a lowstand fan deposit formed after a


Type I unconformity (sensu Vail et al., 1984). Maximum flooding during this cycle is observed in the
well bedded limestones and siliceous shales of the
upper Albian Hilo Formation (Etayo-Serna, 1979).
Cretaceous sedimentation in the southern portion
of Colombia was initiated during Cycle 3 with a basal
transgressive sandstone (Caballos Formation). In
the Putumayo Basin, this unit is composed of quartz
sandstone with minor intercalations of glauconite
and scattered black shales; this is the most important
hydrocarbon reservoir of the basin (Govea and
Aguilera, 1980; C~ceres and Margfoy, 1985; Figs. 15,
16, 17). In the Upper Magdalena Valley, the
Caballos Formation - - which is also an important
producing horizon - - includes crossbedded and occasionally conglomeratic quartz sandstone with intercalations of gray shale with numerous limonitic
concretions (Canard et al., 1964; Beltr~n and Gallo,
1979). These sediments were deposited in fluvial to
deltaic to shallow marine environments. The thickness of the Caballos ranges from 60 to 200 meters in
the Putumayo Basin and from 50 to 170 meters in the
south to over 400 meters in the northern portion of
the Upper Magdalena Valley. In the northern part of
the basin, the Caballos Formation rests conformably
on top of arkosic sandstones, red mudstones, and
conglomerates deposited in a fluvial environment
(Yavi Formation; Mojica and Macia, 1983; Fig. 15).

CAIMAN-5
DATUM

TOP OF VILLETA

CAIMAN-3

CAIMAN-4

Fm

I
,

" -

1/71s'

150 "4-

;
~-~7"'...... "
1GO

/////
j/fllll"

"

"~:":'~"

-4-

CAIMAN'S~

.~

60--I-

ACA~ 1

o~

C^m^N-,

STRATIGRAPHIC

oss-sEcT,oN
c

'
I !

0-L

mDmZ-I

SECTIO. E

PUTUMAYO B A S I N

.
Depomltlonol

AN*6

oyale

Fig. 16. Stratigraphic section E (Putumayo Basin), modified from Caceres and Margfoy (1985); see Fig. 11 for location of section.

Cretaceous paleogeography and depositional cycles of western South America


Deeper water conditions were established during
deposition of the overlying Villeta Formation. These
are hard, black, locally siliceous shales with limestone intercalations. In the Putumayo Basin, three
persistent sandstone intercalations have been named
the "T," "U," and "N Sands" (Govea and Aguilera,
1980; C~ceres and Margfoy, 1985); only the T Sand is
included in Cycle 3. The coarsening upward trend
observed in the SP curve of well logs possibly indicates that this represents a prograding elastic input
at the end of Cycle 3 (Fig. 16). This unit is followed
by a new marine advance of less regional distribution, which is included here as Subcycle 3a. As all
these sandstone intervals pinch out basinward, it is
difficult to identify the different cycles within the
Villeta Shales in the Upper Magdalena Valley.
Cycle 4 (Upper Cenomanian? to Lower Campanian): This cycle coincides with the maximum
flooding of northern South America. In the subsurface of the Barinas Basin, sedimentation was initia-

- - ] ~N~- SAND

:$

~-:__--

:'i':'~' %

"A" LIMESTONE

~:-.'=:-:.~:._.'-

__

!:!:!:!:! )

100

.~

| "USAND

::::::::: ') B ; - ~ I M E S T O N E
~__~_~: . ~

N
::'~::::~:~:::::~-

-.

. I~ ~-_~--

Fig. 17. Generalized well-log signature, oil occurrence, and Cretaceous cycles of the Putumayo Basin.

393

ted with the P Member of the Escandalosa Formation


(Russomano and Velarde, 1982; Figs. 9 and 12). This
27-88 meter thick member is composed of fine to
coarse, crossbedded, locally glauconitic quartzose
sandstone intercalated with minor carbonaceous
shales; it is the most important reservoir of the
Barinas Basin (Feo-Codecido, 1972). The top of the
Escandalosa Formation is composed of up to 91
meters of gray, fossiliferous, occasionally sandy,
crystalline limestone (Guayac/m Member). This unit
is followed by the Navay F o r m a t i o n , which is
composed of two members (Renz, 1959; Gaenslen,
1962). The maximum flooding of the basin during
Cycle 4 took place during deposition of the lower
member (La Morita; Russomano and Velarde, 1982)
of Coniacian age (Ford and Houbolt, 1963). This unit
is composed of dark gray, calcareous to partially
limonitic shale and c a l c a r e o u s i n t e r c a l a t i o n s
containing phosphatic pellets and fish debris (FeoCodecido, 1972). The overlying Quevedo Member
(Coniacian to early Campanian? age) is composed of
hard, white, well bedded siliceous shales, coarsely
bedded sandstones, black shales with fish debris, and
chert intercalations. This unit has been interpreted
as resulting from the progradation of a deltaic system
into the basin (Russomano and Velarde, 1982).
Farther to the west, in the T~chira Depression,
Cycle 4 was initiated with the deposition of a calcareous sand present in the middle of the Seboruco
Shale of the Capacho Formation (Notestein et al.,
1944; Renz, 1977; Fig. 13). Sedimentation continued
with the upper portion of the Seboruco Shale and the
rhythmical intercalations of massive, light gray coquina limestones and dark gray shales of the Guayac~n Member of the Capacho Formation. A diagnostic
late Turonian ammonite was reported for the upper
part of this unit (Renz, 1959; Rod, 1959). Deep water
sedimentation was established during deposition of
the overlying Coniacian-lower Campanian? La Luna
Formation. This unit, which was mostly deposited
under anoxic conditions (11edberg, 1931; Zambrano et
aL, 1971; Ghosh, 1984; Macellari and DeVries, 1987),
is composed of a lower interval, with black shales and
well laminated limestones, and an upper m e m b e r
(T~chira Chert) that includes intercalations of thinly
bedded chert with blue-black limestone and several
phosphorite beds.
Anoxic deposition in the more basinal facies of the
Maracaibo Basin and in the northern Venezuelan
Andes took place almost continuously during Cycle 4.
Sedimentation is represented here by the l,a l,una
Formation, composed of thinly bedded, dark gray to
black bituminous limestone and shale with intercalations of black chert (Hedberg and Sass, 1937;
Sutton, 1946; Renz, 1959, 1981). The age of these
rocks is established as Turonian and Coniacian on
the basis of ammonite data in the type area (Perij/k
foothills; Sutton, 1946); however this u n i t may
include the entire Santonian and part of the Campanian as well (Ford and Houbolt, 1963).
In the Llanos Basin, Cycle 4 onlaps onto the craton
beyond previous cycles and includes a shale unit with

394

C.E. MACELLARI

minor coal intercalations (Gacheta or Chipaque For- the overlying ColSn shales (Macellari and DeVries,
mations) (Renzoni, 1967; Miller, 1979; P~rez and 1987). This has been explained by syndepositional
Bolivar, 1985; Fig. 14). Toward the northern part of deformation affecting the upper beds of the Galembo
the basin, the Gacheta Formation includes delta- Member or the T~chira Chert, which originated by
front facies as well as laterally continuous deltaic slope instability processes (Macellari and DeVries,
1987).
sands (Gabela, 1985). Microfossils from this unit
indicate a Turonian to Campanian age (Gavela,
Cycle 5 (Campanian to Maastrichtian): In the
1985).
In the Cordillera Oriental, Cycle 4 rocks are inclu- Maracaibo Basin, the base of Cycle 5 is formed by the
ded in the Guadalupe Formation or Group. In the Tres Esquinas Member of the Col6n Formation.
Bogota area, the Guadalupe is divided into the Raizal Ghosh (1984) has interpreted these rocks as a con(Dura), Plaeners, Labor, and Tierna Formations (or densed sequence; it is more likely, however, that they
members) (Hubach, 1957; Julivert, 1962, 1968; P~rez represent the transgressive or ravinement horizon of
and Salazar, 1978). The Raizal Formation is com- Cycle 5. Overlying this unit in most of the Maracaibo
posed of thick sandstones separated by shale inter- Basin are 900 meters of gray shales of the Col6n
vals. This is overlain by the Plaeners Formation, Formation, which contain Maastrichtian ammonites
which is composed of well bedded siliceous shales. and mierofauna (Cushman and Hedberg, 1941; Renz,
1977; Figs. 12, 13, 14, 15).
The upper two formations (Labor and Tierna) are
In the Middle Magdalena Valley, the La Luna
composed mostly of well bedded sandstones that in
this work are considered to represent the basal sands Formation is followed by the Campanian-Maastrichtian Umir Shale (Morales et al., 1958; Figs. 13,
of Cycle 5.
14, 15). This formation is composed of 1000 meters of
To the west, the basinal facies of Cycle 4 are
included in the Villeta and the La Luna Formations soft, dark gray shale with abundant concretions and
(Figs, 14 and 15). The Villeta Formation is composed very thin lenticular beds of sideritic ironstone. A
of dark shale, siliceous shale, siltstone, and chert. phosphatic conglomerate is found at many localities
This unit extends from the Putumayo Basin to the in the basal Umir (Maugham et al., 1979).
In the Barinas and Llanos Basins, Cycle 5 is alUpper Magdalena Valley and the central western
portion of the Cordillera Oriental. Two marker units most entirely composed of sandstone - - included in
of wide regional distribution occur within this for- the Burguita and in the upper part of the Guadalupe
mation: a 70 meter thick Lower Chert ('Lidita In- Formations, respectively (Figs. 13 and 15). These are
ferior") and a 50-100 meter thick Upper Chert ('eLi- fine-grained sandstones, usually with a large percendita Superior") (B(irgl and Dumit-TobSn, 1954; tage of clay matrix, including crossbedded intervals
B~irgl, 1961b; Julivert, 1968). Northwest of BogotA, and intercalations of shale and phosphorite (Feothe Lower Chert is included in the La Frontera Codecido, 1972; Gabela, 1985; P6rez and Bolivar,
1985).
Formation of Turonian-Coniacian age (Biirgl, 1961a;
Sandstone deposition also dominated in the BogoEtayo-Serna, 1979). The Lower Chert (Lidita Inferior), which represents the maximum flooding th area (Labor and Tierna Members of the Guadalupe
event of Cycle 4, has been considered as either latest Formation) (Julivert, 1962; Pdrez and Salazar, 1978;
among others). These sediments are followed by
Coniacian or Santonian in age (Biirgl, 1961a;
regressive facies of the Guaduas Formation (Julivert,
Julivert, 1968).
The Upper Chert is dated on the basis of micro- 1963; Figs. 14 and 15). In the Upper Magdalena
fossils as lower Campanian (Btirgl, 1961a,b). This Valley, similar facies are included in the Monserrate
unit of regional distribution is interpreted as repre- Formation - - one of the most important hydrocarbon
senting regressive facies at the end of Cycle 4. The reservoirs in the basin (Fig. 15). This unit, up to 200
presence of widespread siliceous rocks and phosphor- meters thick, is composed of quartz sandstone with
ites is indicative of upwelling conditions developed in intercalations of siltstone, porcellanite, and minor
a semi-enclosed basin bounded to the west by the shale and phosphorite (Beltrhn and Gallo, 1979;
partially emergent Cordillera Central (Macellari and Waddell, 1982). These sands were deposited in outer
neritic to nearshore, tidally influenced environments
DeVries, 1987).
(Waddell, 1982). The upper regressive facies are
In the Middle Magdalena Valley, the Villeta
Formation equivalent facies are included in the La included in the paralic Guaduas Formation, which is
Luna Formation (Morales et al., 1958; Fig. 15). The composed of light gray shale with intercalations of
La Luna Formation is divided into the Salada (50- coal lenses (Corrigan, 1979; Fig. 15).
100 m), Pujama (50-285 m), and Galembo (0-350 m)
Ecuador. Sedimentation in the Oriente Basin of
Members. These units include black shale, well laminated limestones, and chert, and range in age from Ecuador followed a pattern similar to that of the
Turonian to possibly early Campanian (Morales et Putumayo Basin of Colombia and was initiated in
Aptian-Albian time with Cycle 3 rocks (Fig. 18). The
al., 1958; Maugham et al., 1979). Maugham et al.,
(1979) noted that the upper beds of the Galembo are basal unit is the Hollin Formation, composed of 150usually folded with respect to the overlying Umir 300 meters of white to tan, coarse- to very fineShale. A similar relationship was found in the grained, crossbedded quartzose sands (Tschopp, 1953;
T~chira Depression between the Thchira Chert and Canfield et al., 1982; Fernandez Garrasino, 1982).

___ I ~ .

'

I:'_--.I:'-"'--" I ~ "

TrO~

,..-~--,:'""':

--'~

i,

~---'--'J

'-.,,

--.~

V~LANO

~':':':':-==~

COIOMStA

ORIENTE

_~

.......
....

--- ~

~i

....... ~,r.,.. - ; 7 - -

~-

.
-

8A~'41~TONE

8P.TSTONE

-..%.: LIMESTONE

~.

~,~

:-=-=-='~
~

r~:-:-::1

~:

-___-___-:-

.~:::_-:_-:

.....

"~ ~

~'~'f'~~'~';~'~:~

/UANO
~_%~.~

BASIN

.,~i

~"~'~

---=:--N

.........~;~.~,~,::~~.::;

---~

DEPOSITIONAL CYCLES

..

8ACHA-1

l:-_--_---_----:l

~-~.__

8HUSHUFINDI- 1

Fig. 18. Regional stratigraphic cross-section ofthe Oriente Basin of Ecuador showing Cretaceous cycles;data from Tschopp (1953h De] Solar (I 982), and Canfield etal.(1982).

' . . ~ ,- ". ' ~ r ~ ' ~ . , , , . ~ rI " " " :

NE-MARA~ON
BASIN
(Rmm

",v

COFANE- 1

~D
O1

3.

>

[/:

,,<

D.,
C~

,.<

'O

O
0rq

3"

'ID

o
O

P'l

396

C.E. MACELLARI

The Hollin Formation is the main reservoir of the


basin. The lower portion of this unit was deposited
by a braided river system, whereas the upper portion
represents the basal transgressive sand deposited in
a shallow marine environment (Canfield et al., 1982).
Continued deepening of the basin resulted in deposition of the Napo Formation, a thick shale,
limestone, and sandstone sequence of widespread distribution. This unit, which is Albian to Coniacian in
age (Tschopp, 1953; Sauer, 1971), has a thickness
ranging from 236 to 915 meters increasing rapidly in
a southwesterly direction. The Napo Formation can
be divided into three members. The lower member
starts with a thin unit composed of dark gray shale,
siltstone, and light gray, microcrystalline limestones
(Tschopp, 1953; Fernfindez Garrasino, 1982). The
lower member is characterized by the presence of two
sandstone intervals - - the T Sand (lower) and the U
Sand (upper). Similar to the Putumayo Basin, the T
Sand (=middle sand of Tschopp, 1953) displays a
coarsening upward trend interpreted as representing
a shoaling episode during Cycle 3 (Alvarado et al.,
1982). This unit extends throughout the basin, with
a thickness ranging from 30 to 52 meters, and is composed of a glauconitic sandstone, with shale interbeds, mostly deposited in a continental environment
(Canfield et al., 1982). The T Sand is followed by a
limestone and shale unit that originated during a
localized excursion of the sea (Subcycle 3a). These
sediments are Albian in age, as determined by the
presence of diagnostic ammonites (Tschopp, 1953).
A persistent 13-22 meter thick sandstone interval
named the U Sand (=Upper Sand of Tschopp, 1953)
overlies the previous rocks. This unit is interpreted
as forming the basal transgressive unit of Cycle 4.
The U Sand is composed of gray, very fine to mediumgrained, subrounded and moderately to poorly sorted
sands. The upper portion of the unit becomes more
glauconitic and is often argillaceous (Canfield et al.,
1982). The middle member of the Napo Formation is
composed of 78-91 meters of a laterally persistent,
thickly bedded to massive fossiliferous limestone.
The middle limestone member contains Turonian
ammonites (Tschopp, 1953), and lower Turonian
ostracods (Fernhndez Garrasino, 1982).
The upper member of the Napo Formation is composed of gray-green, dark gray, and black, occasionally siliceous, shales deposited during the maximum flooding event of Cycle 4. The thickness of the
upper member varies greatly, mostly due to pre-Tena
erosion, ranging from 73 meters to the south to zero
in the center of the basin to 170 meters to the north
(Tschopp, 1953, Campbell, 1970; Canfield et al.,
1982). A Coniacian ammonite has been reported for
these beds (Tschopp, 1953).
In the eastern and northeastern portions of the
Oriente Basin, the upper member of the Napo Formation is followed by a sandstone unit (M-1 or San
Andr6s Sandstone; Lozada and Endara, 1982), here
considered as forming the base of Cycle 5. This unit
is the lateral equivalent of the Vivi~n Formation of
the MarafiSn Basin. The M-1 Sandstone is composed

of light gray, fine-grained, moderately sorted, angular to subangular friable sandstone with siliceous
cement and intercalations of dark gray shales (Lozada and Endara, 1982). The sandstone is followed in
some places by a shale unit equivalent to the Cachiyacu Shale of the Marafi6n Basin that provides an
excellent seal to the M-1 reservoir. The maximum
thickness of this unit is 68 meters in the southeastern margin of the basin near the P e r u v i a n
border; however, this unit was eroded to the west
during latest Cretaceous time (Del Solar, 1982).
In the central and western portions of the basin,
the Napo Formation is disconformably followed by
the Tena Formation. In general, the amount of
truncation decreases to the east where the Tena is
deposited on top of the sands of the M-1 or Vivi~in
Formation. The Tena Formation is a fine-grained
red bed sequence, with thicknesses ranging from 760
meters in the west to 30 meters in the east. The few
microfossils present here indicate a Late Cretaceous
(possibly Maastrichtian) to Paleocene age (Tschopp,
1953). The Tena Formation was deposited as a response to the uplift and erosion of the proto-Andean
chain to the west.

Peru. Cretaceous rocks are present over most of


the Peruvian territory, but the major depocenters are
the East Peruvian and the West Peruvian Troughs,
which are separated by the NW/SE-trending MarafiSn Geanticline (Benavides, 1956; Cobbing et al.,
1981). Because of space limitations, the following
discussion is limited to these depocenters.
Cycle 1 (Tithonian to Lower Valanginian?):
Sediments are restricted to the West P e r u v i a n
Trough. Sedimentation started during the Tithonian
with deposition of up to 1500 meters of dark gray and
black shales of the Chicama Formation (Figs. 10 and
19). This unit is overlain by the OySn Formation,
which is composed of a few hundred meters of shales
and sandstones with intercalated coal beds (Wilson,
1963). These sediments are mostly continental, but
they contain marine intercalations with Berriasian
ammonites (Cobbing et al., 1981). To the southwest,
in the Lima area, Berriasian sediments are represented by the Puente Piedra G r o u p - - approximately
2000 meters of andesitic flows with intercalations of
limestone and shale (Rivera et al., 1975; Rivera,
1979).
Cycle 2 (Upper Valanginian? to Lower Aptian?): Sediments, also restricted to the West Peruvian Trough, were deposited over a larger area than
those of the preceding cycle. Sedimentation to the
north of the trough, in the Cajamarca area, is believed to have started during Valanginian time with
the Chimu Formation (Benavides, 1956; Fig. 20).
This unit has a gradational lower contact with the
OySn Formation in central Peru, but lies with a
minor erosional d i s c o r d a n c e on the C h i c a m a
Formation in northern Peru (Cobbing et al., 1981).
The Chimu is composed of 700 meters of quartz sand-

Cretaceous paleogeography and depositional cycles of western South America

UNIT

AGE
I_ CAMPANIAN

LITHOLOGY

CHOTA FM.

; RELATIVE

397

"~i(:!,~:~.'.'::.'.~'::

METERS
SANTONIAN
CONIACIAN

TURONIAN

CELENDIN FM.
(25s)

O
N

soft, yellowish
lOSS. shale

~.__-___. ~

CAJAMARCAFM.
(528)

massive end

lithographic Ires.

QUILLQUINAN G R . ~
(140)
CENOMANIAN

shale and marls

MUJARRUN F
(370)
= [ , , 1,
~

UPPER

nodular Ims.
marls

argillaceous Ims.-coquinas

YUMAGUAL FM ~
(496)

massive Ires.
i

3---t

silty marl, ss
PARIAT.~ MBO

bituminous Ires

1120~

thinly bedded Ires.

MIDDLE

LOWER
APTIAN
BARREMIAN
HAUTERIVIAN

CHULEC FM.
(525)

soft brownish shale


yellowish marls

massive Ires.
ferruginoul Oolltl Ires..
sl txtone r Is.

INCA FM. (90)


OYLLARISQ~SGAFM
(FARRAT FM3
(250-500)

quartz

SS.
t

CARHUAZ FM.
(1300)

gray-brownish
s h a l e , . s i l t s o n e & ss.

SANTA FM.

dark g r a y Ims
shale, chert

CHIMU FM.

massive qtz. ss.


(minor s h a l e s )

(195)

VALANGINIAN

F---.....----.~

(655)

BERRIASIAN

OYON FM.
(100)

TITHONIAN

CHICAMA FM.
(1500)

shale, sandstone
coal beds
black shale

Fig. 19. Generalized stratigraphic column of the West Peruvian Trough.

stone, with minor intercalations of plant-bearing


carbonaceous shales as well as minor coal horizons
(Benavides, 1956).
Overlying the Chimu is approximately 200 meters
of limestones and shales of late Valanginian age
the Santa Formation (Benavides, 1956; Wilson,
1963). These sediments were deposited in brackish
as well as shallow marine environments that become
deeper in a southwesterly direction (Benavides,
1956). In the Lima area, the Santa Formation is
replaced by approximately 500 meters of marine
sandstones and shales (Morro Solar, Herradura, and
Marcavila Formations) (Rivera et al., 1975). This
transgressive event was followed by regressive deposits of the Carhuaz Formation, which comprise up
SAES I/4--F

to 1300 meters of non-marine to brackish, varicolored


shales, siltstones, and quartzites. This unit passes
gradationally upward into the quartz sandstones of
the Farrat Formation (Wilson, 1963). The age of the
strata overlying the Santa Formation is poorly constrained but is tentatively included in the IIauterivian-Aptian interval (cf. Cobbing et al., 1981). Near
the coast of Lima, these units are replaced by marine,
thinly bedded limestones and shales of the Pamplona
Formation and the massive limestones of the Atocongo Formation (Rivera et al., 1975).
Cycle 3: A new transgression covered most of
the Peruvian territory during late Aptian or early
Albian time. The basal transgressive unit in the

:...'--:_:_-cAe

0or

. ss

,~

'
,

+"

.~< \ \ \ \ \ \ ,
"

x~,\\\\\""

SHALE

[ ~

CONGLOMERATE

SANDSTONE

SILTSTONE

MARL

LIMESTONE

OOLITIC LIMESTONE

GEANTICLINE;

BALSAS

++

CAJAMARCA

YES TERN CORDIL L ERA

rURA3@I C

-CYCLES

~'. R O S A

VUALLAGA

BASIN

.~

" \

-0. \

\ x,.,.
"~

f8
i:_~

t..=_ . . ~ r

" ' ~\c~m,, \.:/ . R A Z , L

~-,./' ,~RO 1-"'~

;ACHIYACU

kLBIAN
ItPTIAN

3ENOMANIAN

rURONIAN

3ANTONIAN

~AUPAN[AN

~AASTRICHTIAN

Pig. 20. East-west stratigraphic cross-section of Peru showing proposed correlation of cycles across the Mara~dn Geanticline; data from Benavides (I956), Rodrlguez and Chalco (1975), and
3obbing et al. (I981).

meters'

II

~ooo

BARREMIAN
VALANGINIAN

APTIAN

ALBIAN

~ENOMANIAN

TURONIAN

SENONIAN

i,V

C R E T A C E O U S N O R T H E R N PERU

>,

l>

tD
~0

Cretaceous paleogeography and depositional cycles of western South America


West Peruvian Trough is the Inca Formation, which
is composed of 90 meters of interbedded, brownish
gray, oolitic, arenaceous, and ferruginous limestone
and yellowish-green fossiliferous shale with minor
intercalations of quartz sandstone and ferruginous
siltstone (Benavides, 1956; Figs. 19 and 20). This
unit is possibly replaced to the east by the Goyllarisquizga Formation. Farther south, the unit becomes
less ferruginous and increasingly calcareous, and it
passes into the Pariahuanca Formation (Wilson,
1963}. To the southwest, in the Lima area, widespread volcanic activity started at this time and
extended to the end of the Albian or possibly the
Cenomanian. The rocks originated during this event
are included in the Casma Group (Myers, 1974;
Guevara, 1980; Atherton et al., 1983).
The basal transgressive unit in the Mara~6n
Geanticline is formed by the quartz sandstone of the
Goyllarisquizga Formation. However, it is unclear
whether this unit represents only the basal portion of
Cycle 3 or if it includes older intervals as well
(McLaughlin, 1924; Benavides, 1956; Wilson, 1963).
In the northern portion of the West Peruvian
Trough, the Chulec Formation was deposited on top
of the Inca Formation as deeper water conditions
began to develop. The Chulec is composed of 500
meters of fossiliferous, light gray marls with interbedded massive gray limestones that contain a lower
middle Albian macrofauna (Benavides, 1956). These
rocks are followed by a lower interval of friable
brownish shale with intercalations of massive
limestone-bearing abundant thick-shelled bivalves,
and an upper interval of thinly bedded limestones.
The maximum flooding of the West Peruvian Trough
during Cycle 3 resulted in the deposition of anoxic,
black bituminous shales with thin intercalations of
fetid limestones and chert (Pariatambo Formation;
Cobbing et al., 1981; see Fig. 5).
In the northern portion of the Marafi6n Geanticline, the Chulec and Pariatambo Formations are
replaced by the Crisnejas Formation, which consists
of 200-300 meters of calcareous shale and sandstone
with intercalations of limestone (Benavides, 1956;
Fig. 2O).
A normal oxygenated environment was re-established during the late Albian concurrent with emergence at the margins of the trough (Cobbing, et al.,
1981). These regressive sediments are included in
the Pulluicana Group, which is composed of approximately 900 meters of wavy bedded, nodular, argillaceous limestone with intercalations of sandstone and
calcareous shale (see Fig. 6). The Pulluicana Group
was deposited in a shallow marine environment and
contains late Albian to mid-Cenomanian ammonites
(Benavides, 1956).
In the East Peruvian Trough, as now preserved in
the Marafi6n, Ucayali, and Huallaga Basins, the
basal unit of Cycle 3 is included in the Cushabatay
Formation (Figs. 20 and 21). This formation is
composed of 100-450 meters of crossbedded, white
quartz sandstone with intercalations of siltstone and
micaceous shale with plant remains and coal hori-

399

zons (Huerta-Kohler, 1982; Zegarra, 1982). These


sandstones are interpreted to have been deposited in
fluvial to deltaic environments in the Marafi6n Basin
(Kummel, 1948; Pardo and Zt~fiiga, 1976; Soto, 1979).
In the Huallaga Basin, the lower portion of the unit
was deposited in a fluvial environment, but the upper
portion represents coastal barrier sands (HuertaKohler, 1982}. Only the upper portion of the unit has
yielded age-diagnostic palynomorphs, which indicate
an Albian age (sominario and Guizado, 1976). The
maximum flooding event of Cycle 3 resulted in the
deposition of the Esperanza Formation (see Fig. 5).
This unit is composed of 20-200 meters of black fossiliferous shale, with occasional glauconitic as well
as sandstone and limestone intercalations. These
shales contain ostracods, foraminifers, bivalves, and
gastropods which indicate deposition in lagoonal to
shallow marine environments (Soto, 1979}. Palynomorphs present here indicate an Albian age (sominario and Guizado, 1976). To the east and southeast,
the Esperanza is replaced by sandstones and shales
containing plant fragments and a b u n d a n t mica
(Raya Formation).
The upper part of Cycle 3 is represented by the progradation of a deltaic system (Agua Caliente Formation; Soto, 1979; see Fig. 6). This unit is composed of
50-500 meters of medium- to coarse-grained to conglomeratic massive to crossbedded quartz sandstone,
with occasional intercalations of black shale with
abundant iron concretions and p l a n t r e m a i n s
(Martinez, 1980). This unit is considered Albian in
age on the basis of limited diagnostic palynomorphs
(Seminario and Guizado, 1976).
Cycle 4 (Upper Cenomanian to Lower CampanJan?): The boundary between Cycle 3 and Cycle 4 is
not clearly expressed in the West Peruvian Trough,
but a new deepening trend was established with
deposition of the Quillquih~n Group (Figs. 19 and
20). This unit consists of 100 meters of brown shales
and marls interbedded with rusty yellowish limestones, followed by bluish-gray limestones and marls
(Benavides, 1956). The Quillquifi/m Group represents deposition in a neritic environment transitional between the shallower deposits of the Pulluicana Group and the deeper sediments of the overlying Cajamarca Formation. These sediments conrain an upper Cenomanian to lower Turonian fauna
(Benavides, 1956).
The overlying Cajamarca Formation was deposited during the peak of Cycle 4 and consists of about
300-500 meters of gray-brown, very fine grained
limestones and well stratified bluish marls with very
thin shaley intercalations. A rich macrofauna was
collected in the Cajamarca Formation that indicate a
mid- to late Turonian age (Benavides, 1956). Toward
the Marafi6n Geanticline, the Pulluicana and Quillquifi~n Groups and the Cajamarca Formation are
replaced by the Jumasha Formation (Fig. 20). This
unit consists of 1000 meters of bioclastic and pelletal
limestone and dolomite, with intercalations of very
fine grained carbonate and siltstone (Wilson, 1963).

400

C.E. MACELLARI

The Jumasha Formation was deposited in a relatively shallow water environment devoid of clasticinput.
The Cajamarca Formation of the West Peruvian
Trough and the Jumasha Formation of the Marafi6n
Geanticline were followed by the fossiliferousConiacian to lower Santonian Celendin Formation, consisting of up to 300 meters of gray and yellowish, calcareous shale and siltstone with intercalations of
nodular limestone (Benavides, 1956). The Celendin
Formation was deposited in a shallower water setting than was the preceding unit (Benavides, 1956).
In the East Peruvian Trough, Cycle 4 was initiated either in the upper part of the Agua Caliente
Formation or in the overlying marine shales depo-

sited d u r i n g the new marine a d v a n c e (Fig. 21).


These are dark gray shales with interbeds of sandstone, siltstone, and limestone that are included in
the Chonta Formation. The C h o n t a F o r m a t i o n
varies from a thin (100 m), p r e d o m i n a n t l y s a n d y
sequence to the southeast to a thick (1600 m), predominantly shale and limestone sequence to the
northwest in the ancient Amazonian delta of the
Marafl6n Basin (Soto, 1979). The Chonta is divided
into three members: a lower shale interval, followed
by a middle, mostly micritic, limestone interval, and
an upper gray shale and micritic limestone u n i t
(Fuentes, 1980). The Chonta contains an abundant
macro- and microfauna that indicate an Albian to

LI T E R T I A R Y .
;HIYACU FM.
550
meters

5OO

~,N FM.

)NTA FM.

~SNALE
~
400

r~

SILTSTONE
LIMESTONE
SANDSTONE W. WATER

3OO

SANDSTONE W. OIL

)NTA s a n d s t o n e

al CHONTA s a n d s t o n e
2OO
JA CALIENTE FM.

100

'A FM.

3 H A B A T A Y FM.

Fig. 21. Generalizedwell-logsignature, oil occurrence,and Cretaceouscyclesof the northeastern Mara56n Basin (modifiedfrom
Del Solar, 1982).

Cretaceous paleogeography and depositional cycles of western South America


Santonian age (Knechtel et al., 1947; Kummel, 1948;
Seminario and Guizado, 1976; Pardo and Zfifiiga,
1976; Huerta-Kohler, 1982).
Cycle 5 (Upper C a m p a n i a n to Maastrichtian):
By the end of the Santonian, the western portion of
the West Peruvian Trough had been uplifted and
provided the source for a continental red bed molassic
sequence known as the Chota Formation in northern
Peru and as the Casapalca Formation in central Peru
(Benavides, 1956). The age of these red beds is not
clearly established, but a Campanian to Paleocene
age is generally accepted (Cobbing et al., 1981).
In the East Peruvian Trough, Cycle 5 commenced
with the transgressive Vivi6n Formation, which
forms the most important reservoir of the Marah6n
Basin. This unit is composed of 50-300 meters of
yellow-brown to white, coarse- to fine-grained, crossbedded sandstone with minor intercalations of black
shale with plant remains. These sediments represent
an extensive braided fluvial system developed in an
arid climate, followed by coastal barriers deposited at
the initiation of the new transgression (Del Solar,
1982). The continuation of the Cycle 5 transgression
resulted in deposition of the Cachiyacu Formation.
This unit is composed of black shales, marly claystone and siltstone containing a brackish to marine
fauna (Kummel, 1948; Kohl and Blissenbach, 1962}.
The sequence is poorly dated due to the absence of
diagnostic fossils, but a Maastrichtian age is accepted for the Cachiyacu Formation and a CampanianMaastrichtian age for the underlying Vivi6n Formation (Kummel, 1948; Kohl and Blissenbach, 1962).
The Cachiyacu Formation is capped by red beds
(Huchpayacu Formation; Huerta-Kohler, 1982) that
represent the first important clastic influx derived
from the uplifted Andes.
Neuqu~n Basin

Because of its economic importance, its excellent


exposures, and its relatively easy access, the Neuqu6n is one of the best known basins of South
America. The early stage of development took place
in an intra-arc setting with the extrusion of volcanics
together with the development of extensive horsts
and grabens (Digregorio et el., 1984). The Neuqu6n
Basin, with its complexly block-faulted paleogeography, was covered by up to 6000 meters of Jurassic
and Cretaceous, mostly marine sediments. These
sediments represent three major sedimentary cycles
(Jur6sico, Andico, and Riogrfindico) as defined in the
works of Groeber (1929, 1946, 1953; Fig. 22). The
cycles document transgressive/regressive events separated by two major tectonic pulses: the late Kimmeridgian Araucanian phase, and the "mid-Cretaceous" Peruvian phase - - the latter associated with
the intrusion of large granodioritic bodies (Stipanicic
and Rodrigo, 1970; Ramos and Rgmos, 1979).
Extension initiated during the early Mesozoic was
followed by thermal subsidence in a back-arc setting
during the Jurassic to the earliest Cretaceous ("Jut-

401

~sico" and lower "Andico" cycles). During the later


portion of the Cretaceous (upper Andico and Riogr~ndico cycles), sedimentation took place in a foreland basin setting (Digregorio et el., 1984).
Several sequence-stratigraphic syntheses of the
basin have been presented recently (Legarreta et el.,
1981; Gulisano et el., 1984; Mitchum and Uliana,
1985; Gulisano and Legarreta, 1987). In the present
work, the Cretaceous stratigraphy is subdivided into
six second-order cycles, following the scheme outlined by Gulisano and Legarreta (1987).
Cycle 1 (~Lower Mendocian" - - K i m m e r i d g i a n to
Lower Valanginian). The basal unit is represented
by the continental Tordillo Formation, which is composed of red and gray-green conglomerate and sandstone, followed by green siltstone (Digregorio and
Uliana, 1980; Allen et el., 1985). This unit rests with
angular unconformity on top of the "Jur~sico" cycle
(Leanza et el., 1977; Leanza, 1981; Figs. 22, 23, 24).
The basal clastics are followed by a series of progradational depositional sequences, with fluvial to
shallow marine facies to the east-southeast and
basinal facies to the west-northwest (Gulisano et el.,
1984; Mitchum and Uliana, 1985). The basinal facies
are included in the Vaca Muerta Formation. These
are dark gray and black shales, locally highly
organic and bituminous (Weaver, 1931), deposited
under at least partially anoxic conditions. The Vaca
Muerta increases rapidly in thickness from east to
west, from approximately 20 to more than 600
meters. The top of this unit becomes younger from
south to north, ranging from early to late Tithonian,
to Berriasian, to early Valanginian in successive
sections (Leanza, 1973, 1981; Leanza and Hugo,
1977; Mitchum and Uliana, 1985). The outer shelf
facies, which laterally replace and overlie the basinal
facies, are included in the Quintuco Formation.
These are dark gray and black, partly bituminous
shales that become more silty toward the basin
margins. Thicknesses range from 400 meters to the
south to 750 meters near the center of the basin
(Gulisano et al., 1984). To the east and north of the
basin, the Quintuco grades, respectively, into the
Loma Montosa Formation (oolitic grainstones), and
the Chachao Formation (bioclastic calcirudites,
packstones, and calcarenites), which represent midshelf sedimentation (Zilli et el., 1979; Uliana et al.,
1977; Mombrfi et al., 1979; Carozzi et al., 1981;
Legarreta et el., 1981; Legarreta and Kozlowski,
1981; Ploszkiewicz and Orchuela, 1982).
The more proximal facies of Cycle 1, which laterally replace and overlap the previous facies, are
included in the Mulichinco Formation (Fig. 22). This
unit consists of 230-450 meters of sandstones and
conglomerates with thin limestone intercalations
that represent shallow marine, deltaic, and continental environments (Leanza et al., 1977; Leanza, 1981;
Gulisano et al., 1984). The Mulichinco Formation
ranges in age from late Berriasian in the southeast,
to Valanginian in the northwestern portion of the
basin (Leanza and Hugo, 1977; Leanza, 1981).

402

C.E. MACELLARI

~,~o~
JAGUEL
uJ
~AASTRICH- nTIAN
<
ALLEN
<
2

REwATERLATIVE
DEFrH

,oo1 8.,ev.oori,e
,

SO0

Greenish mudstone ~ 6

!~

CAMPANIAN
NEUQUEN
Gp.

SANTONIAN

300/

Yellowish mudstone

1300

& sandstone
(red sandstone
&cnglmerate)

~
Lu3
5 ~

~1
~1

~,
1

CONIACIAN
ALBIAN

APTIAN

250

-"~""~:'):.:l

Red claystone

O RAYOSO
1 H~ I ~ - ~ f l
and sandstne
n
Fm.
O
>,<
n"^'-'^'.'K"A
""^ ".'A"'7~"[ G y p s i f e r o u s
HUlTRIN 1201
390
Fm.
IIIIP : ._.hi sandstone, sa,t
. A . . ^ . .

t__Jl.
....

- w A. . . . . . . . . . .

AGRIO

BARREMIAN

Marls, limestone,
AVILE Mbr.

15o/
Fm.

shale, sandstone

1600

HAUTERIVIAN
L~
<
N

VALANGINIAN O
(3

Sandstone,
sandy limestone,
shale

MULICHINCO 25/
450
Fm.

iJ.i
BERRIASIAN

TITHONIAN
KIMMERIDGIAN

QUINTUCO 2OOl ~
5oo ~
Fm.

VACA
MUERTA too/
2oo
Fm.

Limestone & shale


(sandstone)

"~'~--'-

TORDILLO
400
Fm.

Shale, marl,
bituminous
limestone
siltstone

Fig. 22. Generalizedstratigraphiccolumnofthe Neuqu~nBasin showingdistributionofcyclesand hydrocarbonproducingunits;


Groeber'scyclesshownon the right.

Cycle 2 CMid-Mendocian" - - Upper Valanginian


to Lower Hauterivian). The second cycle was deposited on top of the Intra-Valanginian unconformity.
Sedimentation in Cycle 2 was initiated with lowstand deposits of the upper part of the Mulichinco
Formation (Gulisano et al., 1984), and the transgressive facies are represented by the Agrio Formation (Fig. 22). This unit is divided into upper and
lower members of similar lithology, separated by the

Avil~ Sandstone. In the southern part of the basin,


the Agrio is composed of micritic limestone, grainstone, coquina, and calcareous sandstone, which are
replaced by carbonaceous shale toward the center of
the basin. On the margin of the basin, the Agrio
represents a shallow marine to fluvial and possibly
lacustrine environment (Dellap~ et al., 1979). In the
northern portion of the Neuqu~n Basin, the Agrio
ranges from dark shales and micritic limestone to the

Cretaceous paleogeography and depositional cycles of western South America

403

S C H E M A T I C STRATIGRAPHIC SECTION
M E N D O Z A GROUP
SOUTHERN NEUQUEN BASIN

A,

AGRIO Fm
~
?

QUEBRADA DEL SAPO

".

l- t~-~ -k

~s.-

" ....

---;L'."- -~-

~ "" -.........

"..... i:.'"~'."__

VACA
I

'. . .

I SHALE

lli'...~NE=UIICl~N
II;i~7 ~-

C/

CONGLOMERATE
SANDSTONE

TORD

SILTSTONE
~

LIMESTONE

(~

DEPOSITIONALCYCLE

Fig. 23. North-south stratigraphic section of the Mendoza Group (Cycles 1-3) in the southern portion of the Neuqu6n Basin
(modified from Leanza et aL, 1977).

STRATIGRAPHIC SECTION
MENDOZA GROUP
NORTHERN N E U Q U E N BASIN

CPx-2

LCos-1

LOx-2

LPx-3

C'

Sa-7

Sa-6

Sa-3

HUAuTIiIIVIAN = L
,,~

"--.:-~--,~

L.

<

,.,
i

U<
><

TITHONIAN
o

i . .. . I .,. .~ .
,.-,

I:

,
:'"

:
'"""

.. . ~
_

/ i / ...~;,~'-~00

~,~_~/

o~-5

~
"

...........
,~,:~:,-'.c,.~'."

~ , ~,!~?.:; >
~ ,;,:.; 7 : "

~l'~ L~%4/
~(~
J

'1 black shale I limestone

fossillferous limestone
gray-green calcareous clastlcs

~,~

'-200

/"

gray-green clastlcs

"

"

,~'~/"

""".'"'""

q:'

V'ALANOINI&NBERRIAIAN ~'i! .

,.

"~'-~

~"~--~_
|NEUQUEN ) \
k BA~ilN I.~.~.~ll'-

,.,oo

KIMMERIDGIAF ~ u.
(~

deposltionel cycle

Fig. 24. East-west stratigraphic section of the Mendoza Group (Cycles 1-3) in the northern portion of the Neuqu~n Basin (after
Legarreta et al., 1981 ).

404

C.E. MACELLARI

west, to skeletal and oolitic limestones grading into


gray calcareous sandstone to the east (Legarreta et
al., 1981; Mombrfl et al., 1979). West of this area, in
Chile, facies encompassing Cycles 1 and 2 are included in the Lo Vald6s Formation. This unit shows
a marked shallowing event at the cycle boundary
(Hallam et al., 1986).
Cycle 3 ("Upper Mendocian" - - Upper Hauterivian
to Lower Aptian?). This cycle was initiated with the
Avil6 Member of the Agrio Formation, which is a
lowstand deposit (Figs. 22, 23, 24). This unit is well
represented in the interior (western) portion of the
basin, but is absent to the east. These sands are
followed by the marine shales of the upper Agrio Formation. Regression at the end of Cycle 3 is defined by
the appearance of coarser clastic facies (Troncoso
Sandstone; Dellap6 et al., 1979). To the east, Cycle 2
and 3 rocks are replaced by the proximal clastics of
the Centenario Formation (Digregorio, 1972).
Cycle 4 (~Huitriniano" - - Lower Aptian? to MidCenomanian?). Deposition during this cycle took
place in more restricted marine conditions. The base
of the cycle is represented by the Huitrin Formation,
composed of up to 500 meters of gypsum and carbonate with minor conglomerate, sandstone and shale.
The depositional environment has been interpreted
as sabkha, lacustrine, and fluvial. An increase in
clastics occurs toward the east and southeast where
this unit is replaced by the upper Centenario
Formation. Even though no diagnostic fossils have
been reported, an Aptian-Albian age is generally
accepted for the Hutrin Formation (Uliana et al.,
1975; Volkheimer and Salas, 1975; Digregorio and
Uliana, 1980; Malumi~n et al., 1983).
The overlying Rayoso Formation (=Ranquiles and
Cafiad6n de la Zorra Formations of Uliana et al.,
1975} is composed of up to 900 meters of conglomerate and reddish sandstone and shale, with minor
amounts of gypsum and salt. The depositional environment ranges from fluvial to lacustrine and
possibly deltaic (Digregorio, 1972). Fossils include
dinosaur bones, wood fragments, and freshwater
bivalves (Digregorio, 1972). The upper boundary of
Cycle 4 coincides with a regional unconformity
associated with a major diastrophic face (Stipanicic
and Rodrigo, 1969; Uliana et al., 1975). There is no
agreement, however, on how much time is missing
between Cycles 5 and 6. According to Riccardi
(1987), this hiatus extended from the early Cenomanian to the Coniacian, whereas Gulisano and
Legarreta (1987), considered sedimentation to have
been almost continuous.
Cycle 6 ("Neuqueniano" - - Mid-Cenomanian? to
Lower Campanian?). This cycle was deposited in a
continental setting, and is included in the Neuqu6n
Group (Digregorio, 1972; Fig. 22). These sediments
lap on top of older rocks in an easterly and southeasterly direction (Uliana et al., 1975). The Neuqu6n
Group is up to 1400 meters thick and is divided into

three formations m the Rio Limay, Rio Neuqu6n, and


Rio Colorado m representing three successive sedimentary cycles. Each cycle starts with sandstone
and conglomerates and is capped by red mudstone
(Cazau and Uliana, 1972).
Cycle 7 ("Malalhueyan" - - Upper Campanian? to
Paleocene). These sediments herald the onset of the
first Atlantic-derived transgression into the basin (cf.
Wichmann, 1927; Bertels, 1979; Uliana and Dellal~,
1981). The basal unit of Cycle 7 is the Allen Formation, approximately 65 meters thick, which is
composed of light gray sandstone and conglomeratic
sandstone, followed by claystone with intercalations
of sandstone, and finally by tuff and stromatolitic
limestone (Andreis et al., 1974; Uliana and Dellap6,
1981). The Allen Formation is interpreted as a sand
flat environment close to the shoreline (lower member), followed by a mixed flat setting in a tidal
environment (middle member), and finally capped by
a restricted and hypersaline marginal sabkha environment (Andreis et at., 1974).
Clearly defined marine conditions were established during deposition of the overlying Jaguel Formation, which represents an open shelf environment
distant from the source of clastic input (Bertels,
1970). This unit is composed of 95 meters of green to
yellowish green, monotonous, massive, partly friable
siltstone and well laminated clays (Uliana and
Dellap6, 1981).
Austral Basin

The Austral (or Magallanes) Basin originated


during a Triassic to Jurassic extensional episode
(Gust et al., 1985), which was immediately followed
by the formation and infill of a marginal basin
("Rocas Verdes") to the west of the basin (Dalziel et
al., 1974; Dalziel, 1981). This event was followed by
a time of thermally driven subsidence just after the
emplacement ofophiolitic bodies in the Rocas Verdes
Basin. During the Late Cretaceous, a thick sedimentary package was deposited in a foreland setting
following uplift of the Andes and the formation of a
foredeep adjacent to the uplifted orogen.
The Cretaceous sequence of the basin has recently
been subdivided into depositional sequences by
Biddle et al. (1986) on the basis of seismic, stratigraphic, and well data, and by Arbe (1987) based
mostly on outcrop data. Here, the Cretaceous sequence is divided into six second-order cycles. However, the lack in many cases of diagnostic fauna or
correlation between ambiguously dated isolated outcrops makes it difficult to establish with certainty
the age of several of the cycle boundaries. The cycles
presented here agree in general with those outlined
previously.
Cycle 1 (Tithonian to Lower Aptian). This cycle
was initiated with the Springhill Formation, which is
the main hydrocarbon reservoir of the basin. This
unit is comprises a lower, continental, member com-

Cretaceous paleogeography and depesitional cycles of western South America


posed of subangular, medium- to coarse-grained
quartz sandstone with a tuffaceous matrix, and an
upper portion which is composed of well sorted, crossbedded, glauconitic quartz sandstone with clay
matrix and shell fragments and represents beach and
shallow marine deposition during the transgressive
phase of Cycle 1 (Thomas, 1949; Cecioni, 1955;
Riccardi, 1977; Robles, 1982, 1984; Miller et al., 1982;
Hinterwimmer et al., 1984). Macro- and micropaleontologic data indicate that this unit is timetransgressive, becoming younger toward the north of
the basin (Fig. 25). Thus, these sediments are
Oxfordian to Kimmeridgian to the south, in Chile
(Sigal et al., 1970; Natland et al., 1974), Tithonian to
Berriasian in the Lago Argentino area (Blasco et al.,
1979), Berriasian in the Lago San Martin area
(Riccardi, 1977), and up to Valanginian in the Lago
Belgrano area (Aguirre Urreta and Ramos, 1981).
The lack of more precise age control does not allow
the determination of whether this sand represents a
single unit or if, as proposed by Biddle et al. (1986), it
includes a series of backstepping intervals.
A thick package of basinal sediments (Rio Mayer
Formation) was deposited as the basin continued to
subside (Figs. 9 and 25). These are dark gray to black
laminated shales with minor limestone intercalations that were deposited under partially anoxic conditions (Riccardi, 1971; Riccardi and Rolleri, 1980;
Nullo et al., 1981). The Rio Mayer shales thicken
considerably from a few meters to the north (Aguirre
Urreta and Ramos, 1981) to over 700 meters in the
Lago San Martin area to the south (Riccardi and
Rolleri, 1980). In the southwestern portion of the
basin, similar facies are included in the Zapata Formation (Katz, 1963), whereas in the subsurface they
are included in the Palermo Aike Formation (Russo
and Flores, 1972) or into the Pampa RincSn Formation (= Lower Inoceramus Shale; Flores et al., 1973).
The maximum flooding during Cycle 1 took place
in the late Hauterivian (Riccardi, 1987). After this
event, the northern portion of the basin was prograded by a coarsening upward clastic sequence (Rio
Belgrano Formation). This unit, which ranges from
the Barremian to the lower Aptian, is composed of
green sandstone with intercalations of shale, and it is
replaced by coarser crossbedded sandstone toward
the top, deposited in shallow marine to intertidal
settings (Ramos, 1979). In the Lago Cardiel area,
this interval is represented in the lower portion of the
Cerro Pelado Member of the Piedra Clavada Formation (Ramos, 1982).
Cycle 2 (Mid-Aptian to Mid-Cenomanian? ). A new
sedimentary cycle was initiated in the mid-Aptian.
To the north of the basin, the Rio Tarde Formation
was deposited unconformably on top of previous rocks
(Ramos, 1979). This sequence includes 320 meters of
conglomerates at the base and purple mudstones and
tufts to the top deposited in a continental setting
(Ramos, 1979). Farther south, these sediments grade
into the sandstones of the Piedra Clavada Formation,

405

which include a marine intercalation near its base


(Ramos, 1982}. In the Lago San Martin area, similar
sediments are included in the Kachaique Formation,
with a well developed marine fauna near its base
containing lower to mid-Albian ammonites (Riccardi
et al., 1987}.
The basal transgressive units are followed by the
north to south progradation of beach, tidal, and fluvial sandstones and finally red mudstones and tufts
(Cardiel Formation). In the south of the basin, however, the Rio Mayer Formation continued to be deposited without noticeable break. Because of the lack of
diagnostic fossils in the north and uncertainties in
the exact provenance of the fauna to the south, the
upper limit of Cycle 2 remains poorly constrained.
However, the youngest fossils found in this cycle are
possibly Cenomanian in age (Riccardi 1979}.
Cycle 3 (Mid.Cenomanian? to M i d - C o n i a c i a n ? ).
The initiation of this cycle coincided with a major
tectonic event that resulted in the closure and
deformation of the marginal basin located to the west
of the basin. This event gave rise to the "Paleoandes"
and the concurrent formation of a foredeep immediately adjacent to the east (Dalziel et al., 1974;
Winslow, 1980; Wilson, 1983}. This event is reflected
in a paleontological hiatus extending from the upper
Cenomanian to the lower Coniacian present in large
portions of the basin (Malumi~n, 1968; Malumi~n et
al., 1971; Flores et al., 1973; Riccardi, 1984). A thick
turbidite sequence was deposited in this new depocenter (Punta Barrosa Formation in Ultima Esperanza, and basal Cerro Toro Formation in Lago
Argentino; Cecioni, 1957; Katz, 1963; Arbe and
Hechem, 1984a). Toward the northern basin edge,
these turbidites are possibly replaced by gray and
green siltstone and shale of the Mata A m a r i l l a
Formation (Russo and Flores, 1972; Leanza, 1972).
This unit contains diagnostic Coniacian ammonites
(Leanza, 1969}.
Cycle 4 (Mid-Coniacian? to M i d - C a m p a n i a n ? ) .
The contact between Cycle 3 and Cycle 4 was recognized in the Lago Argentino area by Arbe and
Hechem (1984a), and it is also represented in the
subsurface (Biddle et al., 1986). Large amounts of
sediment were dumped below wave base by turbidity
flows on a deep sea fan system in the western margin
of the depositional trough adjacent to the uplifted
cordillera (Scott, 1966; Winn and Dott, 1979; Dott et
al., 1982). These sediments are included in the Cerro
Toro Formation (Katz, 1963} and comprise more than
2000 meters of turbidites transported southward
along a major trough (Scott, 1966; Vilela and Csaky,
1968; Winn and Dott, 1979}. The Cerro Toro and
equivalent units contain Santonian to lower Campanian ammonites (Katz, 1963; Leanza, 1963, 1967;
among others). At the northern basin edge, these
sediments were possibly replaced by greenish calcareous sandstone (El Alamo Formation) bearing a
similar ammonite fauna (Leanza, 1967, 1972).

,.

- -_-

~ -

74"

7(yw

_ zAp,I TA - - -

--TsE_s~s_o_s--

CRETACEOUS

_~

- ---

..................

Lago
Argentlno

CYCLES

'

'~,_

4,

RIO MA YER

~SANDSTONE

"~TUFF

MUDSTONE

GRAY MUDSTONE

Lago
BeIgrano

CONGLOMERATE

APPROXIMATE HORIZONAL SCALE


m
m
m
J
0
50
lOOkm.

SANDSTONE, SHALE

~RED

~SILTSTONE,

"

Lego
Card/el

BASIN

F~-~SHALE(PARTLY ANOXIC)

TURBIDITES

"

Lago
San Martin

AUSTRAL

~
~

Lego
Pledra
V l e d m a Clavada

- WESTERN
N

....

Lago
Pueyrreddn

Fig. 25. Generalized stratigraphic section of the western margin ofthe Austral, or Magallanes, Basin based on outcrop data; data from Ferugtio (1949J. H~nicken (1955), Katz (1963), Riccardi
(1971,1977|, Blasco etal. (1979), Ramos (1979,1982), Aguirre Urreta and Ramos (1981), Nullo et al. (1981), Arbe and Hechem (1984a, b), and Macellari etaL ~in rexiew).

BERRIASIAN

VALANGINIAN

HAUTERIV~N

BARREMIAN

APTIAN

ALBIAN

CENOMANIAN

TURONIAN

CONIACIAN

8ANTONIAN

CAMPANIAN

MAASTRICHTIAN

Ultima
Esperenza

C3
r~
r~

>

C3

O
O~

Cretaceous paleogeography and depositional cycles of western South America


Cycle 5 (Mid-Campanian? to Lower Maastrichtian?). The base of this cycle is defined by a pronounced basinward shift in the depositional facies,
possibly related to a sea-level fall. Cycle 5 is marked
by the north to south progradation of alluvial plain
muds (Cerro Fortaleza Formation), delta front sands
(La Anita Formation), and delta slope turbidites
(Alta Vista and Tres Pasos Formations) (Arbe and
Hechem, 1984b; Feruglio, 1949; Katz, 1963; Smith,
1977; Nullo et al., 1981; and Macellari et al., in review). Deposition of turbidites ceased near the end of
the cycle when the Tres Pasos Formation was replaced by the shallow marine sandstones and shales
of the Cerro Cazador Formation and deeper water
shales of the Fuentes Formation (Charrier and Lahsen, 1969). These sediments contain late Campanian
ammonites (Htinicken et al., 1980; Maeellari, in
press).
Cycle 6 (Lower Maastrichtian? to Paleocene). A
basinward migration of continental facies is observed
at the base of this new cycle, which is only preserved
south of Lago Viedma (Fig. 25). In the northern
portion of the basin, sedimentation started with
braided river deposits (La Irene Formation), followed
upward by meandering river mudstones and shallow
marine sandstone and shale (Chorrillo Formation).
This unit interfingers toward the south with fossiliferous marine sandstone and shale (Cerro Cazador
Formation; Feruglio, 1949; Hiinicken, 1955; Nullo et
al., 1981; Macellari et al., in review). The age of this
cycle is poorly constrained, excepting to the south
where Maastrichtian ammonites are present (Hiinicken, 1965; Charrier and Lahsen, 1969}. This cycle
possibly spans the Cretaceous/Tertiary boundary
(Macellari, 1985).

DISCUSSION
Cycles
A series of marine advances that resulted in successive second-order cycles is reflected in the Cretaceous sediments of the western margin of South America (Figs. 9, 10, 26). Five cycles are recognized from
western Venezuela to southern Peru. The maximum
flooding events during these cycles took place in the
Berriasian? (Cycle 1); in the late Valanginian in
western Peru, Hauterivian in the Middle Magdalena
Basin, and early to mid-Aptian in the Maraeaibo
Basin (Cycle 2), late Albian (Cycle 3), late TuronianConiaeian (Cycle 4), and in the mid-Maastrichtian?
(Cycle 5). Cycle 1 is poorly constrained as it is restricted to only a few poorly dated depocenters. Uncertainties in the maximum flooding event of Cycle 2
may be due to the presence of more than one transgressive event or, more likely, to different tectonic
subsidence histories. The maximum flooding events
of Cycles 3 and 4 are well constrained and they seem
to be synchronous throughout the area. Uncertain-

407

ties in the timing of maximum flooding during Cycle


5 stem from the lack of diagnostic fossils.
Recognized cycles in the back-are basins of
southern South America have similarities as well as
differences with their northern counterparts. In the
Neuqu~n Basin, the six cycles recognized have maximum flooding events during the Tithonian (Cycle 1),
early Hauterivian? (Cycle 2), Hauterivian-Barremian (Cycle 3), and possibly mid-Maastrichtian for
Cycle 6. Cycles 4 and 5 represent mostly non-marine
sedimentation and usually lack diagnostic fossils;
thus, they are poorly constrained in time. The boundary between Cycles 4 and 5 coincides with a well
documented tectonic event (ttuantraico Phase of
Stipanicic and Rodrigo, 1970}.
In the Austral Basin, maximum flooding events
took place during the Hauterivian (Cycle 1), during
the late Aptian-early Albian (Cycle 2), in the midCampanian? (Cycle 5), and possibly in the midMaastriehtian (Cycle 6). Cycles 3 and 4 in the Austral Basin, extending from the mid-Cenomanian to
the mid-Campanian, lack detailed paleontological
information. The boundaries between Cycles 1 and 2
and Cycles 2 and 3 coincide with diastrophic movements (early Patagonian and Main Patagonian or
Peruvian phase, respectively) (Stipanieic and Rodrigo, 1970; Malumi~n and Ramos, 1984; among many
others).
Only two unconformities have a continent-wide
distribution: the intra-Cenomanian, and the intraCampanian. Of these, only the mid-Campanian
event correlates with a major unconformity in the
Haq et al. (1987) chart of coastal onlap. The intraValanginian unconformity, represented in northern
South America and in the Neuqu~n Basin, has been
reported also for the subsurface of the Austral Basin
(Biddle and Uliana, 1976). It may also represent a
regional event, well expressed as a Type I unconformity in the global chart of coastal onlap.
The recognized cycles are compared in Fig. 26 to
the latest of Vail's charts (in Haq et al., 1987). A
precise comparison is difficult as the cycle boundaries
are usually poorly dated because both the regressive
facies of the underlying cycle and the transgressive
facies of the overlying cycle are normally poorly
fossiliferous. Thus, the reliable dates are provided by
the marine interval (when present) of each cycle.
Comparison of the maximum flooding events with
the condensed zones in the same chart is not possible
since these are plotted for the third-order cycles only
(thus beyond the limit of resolution of this work). In
northern South America, the contacts between
Cycles 1 and 2, 2 and 3, and 4 and 5 coincide with
major unconformities in Vail's chart. In the Neuqu~n Basin, only the contacts between Cycles 1 and 2
and 5 and 6 bear some resemblance to those predicted
globally. In the Austral Basin, three cycle boundaries do not coincide, while two agree with those
predicted by the global model.
In conclusion, it appears that the recognized cycles
have some elements in common with global patterns,
but overall, within the areas considered the northern

408

C.E. MACELLARI

portion of South America, the Neuqu6n Basin and


the Austral Basin each record a somewhat different
stratigraphie evolution.
Relative Changes in Sea Level

The recognized cycles are the result of several


interactive factors such as eustatic fluctuations, thermal and tectonic subsidence, and variations in sediment input into the basins. However, it is difficult to
separate the individual contribution of these factors
in the western margin of South America. This problem has already been discussed extensively by
Burton et al. (1987), who noted the impossibility of
separating the contribution derived from each one of
these independent (or semi-independent) variables.
An attempt was made to plot areas covered by the
sea through time in relation to arbitrary standard
areas using more detailed paleogeographic maps
than those presented here. This approach posed
several problems, including: i) the difficulty of clearly distinguishing the contact between marine and
non-marine facies in the more proximal sections;
ii) the lack of precise age control in the shallower

STAGE

RELATIVECHANGESOF
COASTAL
ONLAP

AGE

NORTHERN
S. AMERICA

2lid Order
I.;I/Cle$

MA
MAASTRICHTIAN

-70

CAMPANIAN

6
5

(~ON|ACIAN

TURONIAN
CENOMANIAN

<
Z

<
N

-90-95-

n..
el
a.

-~

=E

Illl ?[llllllllll Illlllllli ?1111111111


II I ]{1[[1 [[11] [ 1 l II 1111 I1"~
r-

z6
O
O

-80 -

-85-

AUSTRAL
IMAGALLAHESi
BASIN

SANTONIAN

NEUQUENBASIN

-75-

marine facies; and iii) the erosion and truncation of


Cretaceous rocks by Tertiary sediments in several of
the proximal pericratonic basins (Barinas, Llanos,
and possibly the MarafiSn and Ucayali Basins). With
these caveats in mind, a curve of percentage of area
flooded versus time was plotted for western Venezuela and Colombia, Peru and Ecuador, and southern
South America (Fig. 27).
The Venezuela-Colombia and Ecuador-Peru curves are nearly identical, with the possible exception
of the earliest Cretaceous (Fig. 27). However, these
differences may be due to uncertainties in dating and
establishing the lateral extent of these rocks. In
general, the northern portion of the western margin
of South America was affected by a progressive encroachment of the sea up to approximately 89 Ma
(late Turonian-Coniacian), followed by a progressive
retreat. The maximum flooding events recognized in
northern South America coincide with eustatic peaks
in Vail's chart (Fig. 27). In addition, the time of
maximum Cretaceous flooding is in approximate
agreement with global estimates by Hays and Pitman (1973), Pitman (1978), and Sliter (1976), but
differs from that of Hancock and Kauffman (1979),

l
ilillll[lllllllllllllllll

r , l r ,:E,1:L ~ "~':r'J:',''': " ' ~ ' ~ I ' 1 [ " ' ' " 1 "

;~,

3a

-lOO-

ALBIAN
-105-

AP.TIAN

-110-

BARREMIAN

,-115-1

-120-

N
n-

I11I[I

I1 1 1 1 1 1 L l l l 111 11 I I

(,,)

f~'
.4

UJ

eVALANGINIAN

BI:RRIASIAN

-125-

-130-

"~
I

HAUTERIVIAN

N,

l - r J H I H J l r H [ I , [ J L , , L H J ~'{II: ~

""i'

===';

. . . . . . . .
.......

Fig. 26. Comparison of Cretaceous cycles recognized in northern South America and in the NeuquGn and Austral Basins with the
global chart of coastal onlap of Haq et al. (1987). The chart of coastal onlap has been simplified in order to show only major
unconformities: 1, Type I unconformity; 2, Type I! unconformity (sensu Vail et al., 1984). Time scale aider Haq et al. (1987).
Groeber's cycles are shown in the NeuquGn Basin. The dotted line in the Austral Basin indicates additional unconformities
recognized by Biddle et al. (1986) on the basis of subsurface information. Solid arrows indicate major paleontological discontinuities.

Cretaceous paleogeography and depositional cycles of western South America

STA6E

M AASTRICHTIAN

AGE

NORTHERNS. AMERICA SOUTHERNS. AMERICA


R e l a t i v e of a r e a f l o o d e d

=0 ,.o 6,o sps


""'"

-70-

EUSTATICCURVE

R e l a t i v e of a r e a f l o o d e d

2~

\\

4~
'

,5/

/
/

'...

CAMPANIAN

I****

s~

slos

/I

SANTONIAN

CONIACIAN
I"URONIAN
CENOMANIAN

-95

409

<g'"

ii!i!!iiiiiiiiiiiiiiii!iiii iiiii!b~r-iiiiiii

!ii i i !i i1!i i

ALBIAN

,o ,....

HHHHi
HHHHHi
HH!HiHilHH!Hi
I
!!I

APTIAN
BARREMIAN
HAUTERIVIAN

VALANCINIAN

-125-

""*.

BERRIASIAN

Anoxic Facies
Fig. 27. Comparisonof variation in percentage of area floodedin northwestern South America (solid line, Venezuela-Colombia;
dotted line, Ecuador-Peru;numbers indicate major floodingevents within the recognizedcycles),and southern South America,
with global eustatic fluctuationsand percentagesofarea flooded(after Eyged, 1956;Sliter, 1976;and Haq et al., 1987). Large solid
dots indicate control points. Time scale after Haq et al. (1987). Stipple pattern indicates times of recognizedglobal anoxic events
(fromJenkyns, 1980),and times of majoranoxicfaciesdepositionin western South America.
who placed this maximum in the Maastrichtian. A
late Albian event is also well represented, but with a
smaller amplitude than the overlying peak of Cycle
4. These events are r e m a r k a b l y similar to those
described for Nig er i a and India (cf. Hancock and
Kauffman, 1979}.
Widespread anoxic facies are present in northern
South America in the upper Albian and in the upper
Turonian to Santonian intervals (Figs. 27, 28). The
timing of these anoxic facies coincides in general
terms with "global anoxic event s " (e.g., J e n k y n s ,
1980; De Graciansky, 1984; Jacquin, 1987; A r t hur et
al., 1987; among others).
A very different situation is found in s o u t h e r n
South America. Here, the flooding curve is almost
inverse to th at described previously, with a minim u m value during the mid-Cretaceous, and maximum floodings in the Early Cretaceous and parti-

cularly in the Late Cretaceous (Maastrichtian), when


the m a x i m u m epeirogenic submersion took place
(e.g. Uliana and Biddle, in press). This Maastrichtian peak is also well represented in northern Europe
(Hancock, 1961; Hancock and Kauffman, 1979) and
in western Africa (Reyment and Morner, 1977).
Anoxic events in the Neuqu~n and Austral Basins
are concentrated in the Early Cretaceous, also coinciding with peak floodings in both areas. These
anoxic facies, however, do not correlate with recognized "global anoxic events" (Fig. 27). This is similar
to what was observed in the South Atlantic where
Early Cretaceous anoxic black shales are restricted
to localized basins and are controlled by local, not
"global," factors (Jacquin, 1987).
It appears that the differences observed between
northern and southern South America can be explained by tectonic processes interacting with global

410

C.E. MACELLARI

trends (see also Hallam et al., 1986). In particular,


the mid-Cretaceous, or Peruvian, orogeny (Steinmann, 1929) profoundly affected the distribution of
emergent landmasses and their possible connections
with the sea. This deformation is closely related to a
time of rapid spreading rates (Larson and Pitman,
1972; Frutos, 1981) and the final opening of the
South Atlantic (Malumi~n and Ramos, 1984; among
others). In the stratigraphic record of northern South
America, the effect of the Peruvian orogeny is reflected in the progressive increase of the areal perc e n t a g e c o v e r e d by red beds d u r i n g the L a t e
Cretaceous (Fig. 29). However, the incipient uplift of
the western margin of the continent did not effectively isolate the different basins from a marine
connection. Thus, presumably global eustatic fluctuations and anoxic events continued to influence
this portion of South America. In southern South
America, in contrast, the incipient Andean uplift
resulted in the isolation of the Neuqu~n Basin from
the Pacific margin during the mid-Cretaceous m the
time of maximum flooding to the north. Finally, in
the Maastrichtian, whereas the Argentine Atlantic
margin - - d e v o i d of large topographic barriers - - witnessed an extraordinary increase in the areas flooded
(Malumifin et al., 1983; Uliana and Biddle, in press),
the northern portion of the continent was affected by
the progressive uplift of the Andes and the associated
retreat of marine facies.

CONCLUSIONS
Cycles observed in western South America seem to
be the result of a combination of several factors, each
one unique to each basin or major area of similar
tectonic history. Although similarities can be established with global patterns, each basin or major area
records a unique set of events. The knowledge of
these cycles may help to predict and explain the
presence of potential hydrocarbon reservoirs and
time distribution of source rocks.
The n o r t h w e s t e r n m a r g i n of S o u t h A m e r i c a ,
extending from Venezuela to Peru, followed a similar
pattern of evolution during the Cretaceous. Five
depositional cycles are observed here; regional anoxic
facies and m a x i m u m flooding events coincide in
general terms with worldwide recognized trends.
The southwestern margin of South America (Neuqu~n and Austral Basins) shares some common elements, but differs to a large extent from the patterns
recognized in the north. Here, six (but not exactly
equivalent) cycles were recognized for both basins.
Anoxic facies are restricted to the Lower Cretaceous
and seem to be controlled mostly by local aspects of
the basin evolution. Two widespread flooding events
were recorded here - - one in the Early Cretaceous,
and the other in the Late Cretaceous (Maastrichtian). This last flooding episode coincides with a
recognized global highstand of sea level.

A R E A C O V E R E D BY
P O T E N T I A L OIL S O U R C E R O C K S

E
\

1000-

500~Worid-wide
anoxic events

0'

TIME
Fig. 28. Area covered by potential oil source rocks in northern South America (western Venezuela to Peru) during the Cretaceous,
and comparison with anoxic events (Schlanger and Jenkyns, 1976).

Cretaceous paleogeography and depositional cycles of western South America

411

7501
AREA COVERED BY RED BEDS

\
Q

t~
O

Peruvian orogeny
r

250.

TIME
Fig. 29. Distribution of area covered by red beds during the CretAceous in northern South America (western Venezuela to Peru).

The contrasts observed between these two portions


of South America can be explained by differences in
tectonic framework and evolution superimposed on
global trends. To the north, sediments were deposited around the tectonically stable GuayanaBrazilian Massifs and thus recorded "global signals"
such as anoxic events and major eustatic changes.
T h e b a c k - a r c b a s i n s o f s o u t h e r n S o u t h A m e r i c a , on
the other hand, represent a tectonically active environment. Thus, the effect of the Peruvian orogeny
o v e r p r i n t e d to a l a r g e e x t e n t w o r l d w i d e t r e n d s
during the mid-Cretaceous, and a "global signal" was
c l e a r l y r e - e s t a b l i s h e d in t h e a r e a o n l y d u r i n g t h e
Maastrichtian.

REFERENCES
Aberg, G., Aguirre, L., Levi, B., and Nystrom, J. O., 1984.
Spreading, subsidence and generation of ensialic marginal basins:
An example from the Early Cretaceous of central Chile. In:
Volcanic and Associated Sedimentary and Tectonic Processes in
Modern and Ancient Marginal Basins (Edited by Vokelaar B. P.,
and Howells M. F). Geological Society of London, Special Publication 16,185-193.
Aguirre, L., 1985. The Southern Andes. In: The Ocean Basins and
Margins, 7A: ThePacific Ocean (Edited by Nairn, A. E. M., Stehli,
F. G., and Uyeda, S.), pp. 265-376. Plenum Press, New York.
Agulrre Urreta, M. B., and Ramos, V. A., 1981. Estratigrafia y
paleontologla de la altA Cuenca del Rio Roble, Cordillera Patag6nica, Prov. de Sta. Cruz. Actas, VIII Congreso Geol6gieo Argentino, San Luis 3, 101-138.
Allen, R., Nickelsen, B., and Feehan, J., 1985. The Geology of the
Neuqu~n Basin, Argentina. Earth Sciences and Resources Institute, University of South Carolina, Technical Report 85-0010, 115
p.

Acknowledgements--This study was made possible thanks to


grants from Tenneco Oil Company, which supported the earlier
phases of this study and field work in Colombia. Field work in
southern Argentina and Chile and the final stAges of this research
were supported by the National Science Foundation (Grant EAR8418145). I would like to thank D. Wiman (Tenneco) for fruitful
discussions on the CretAceous cycles, and R. Allen, T. DeVries, and
A. Nairn (Earth Sciences and Resources Institute, University of
South Carolina) for advice and criticism on earlier versions of this
paper. I thank A. Hallam (University of Birmingham) and an
anonymous reviewer for their valuable comments.

Alvarado, G., Bonilla, G., De Rojas, G. G., and Arroyo, R. F., 1982.
An#ilisis sedimentol6gico de la Zona Arenisca T en la Cuenca del
Napo. In: Simposio Exploracidn Petrolera en las Cueocas Subandinas de Venezuela, Colombia, Ecuador y Peru, Vol. 2.
Asociaci6n Colombiana de Geblogos y Geoflsicos del Petroleo,
BogotA.
Andreis, R. A., lfiiguez Rodriguez, A. M., Lluch, J. J., and Sabio, D.
A., 1974. Estudio sedimentol6gico de los formaciones del CretAcico
Superior del area del Lago Pellegrini (Prov. de Rio Negro, Rep.
Argentina). Revista Asociacidn Geoldgica Argentina 29 (1), 85104.
Arbe, H. A., 1987. E1CretAcico en la Cuenca Austral de Argentina.
Boletin de lnformaciones Petrolera, Tercera Epoca (Buenos Aires)
4 (11),91-110.

412

C.E. MACELLARI

Arbe, H. A., and Hechem, J. J., 1984a. Estratigrafla y facies de


dep6sitos marinos profundos del CretAcico Superior, Lago Argentino, Provincia de Santa Cruz. Aetas, IX Congreso Geol6gico
Argentino, San Carlos de Bariloche 5, 7-41.
Arbe, H. A., and Hechem, J. J., 1984b. Estratigrafla y facies de
dep6sitos continentales, litorales y merinos del Crethcico Superior,
Lago Argentino, Provincia de Santa Cruz. Acres, IX Congreso Geo16gico A rgentino, San Carlos de Bariloche 7,124-158.
Arthur, M. A., Schlanger, S. O., and Jenkyns, H. C., 1987. The
Cenomanian-Turonian oceanic anoxic event, II: Palaeoceanographic controls on organic-matter production and preservation.
In: Marine Petroleum Source Rocks (Edited by Brooks, J., and
Fleet, A. J.). Geological Society of London, Special Publication 26,
401-420.
Atherton, M. P., Pitcher, W. S., and Warden, V., 1983. The
Mesozoic marginal basin of central Peru. Nature 305, 303-306.
Audebaud, E., Laubacher, G., and Marocco, R., 1976. Coupe
g6ologique des Andes du sud du P6rou de l'Oc6an Pacifique au
Buclier br6silien. Geologische Rundschau 65,223-264.
Barrero, D., 1979. Geology of the Central Western Cordillera, West
of Buga and Roldanillo, Colombia. Ingeominas (Bogota), Publicaci6n Especial 4, 75 p.
Bartok, P., Reijers, T. J. A., and Juhasz, I., 1981. Lower Cretaceous
Cogollo Group, Maracaibo Basin, Venezuela: Sedimentology,
diagenesis, and petrophysics. Bulletin of the American Association
of Petroleum Geologists 65, 1110-1134.
Beltran, N., and J. Gallo., 1979. The geology of the Neiva SubBasin, Upper Magdalena Basin, southern portion, 1968. In: Geological Field-Trips, Colombia, 1958-I 978, pp. 253-275. Colombian
Society of Petroleum Geologists and Geophysicists, BogotA.
Benavides C., V. E., 1956. Cretaceous system in northern Peru.
American Museum of Natural History (New York), Bulletin 108
(4), 359-493.
Bertels, A., 1970. Los foraminiferos planct6nicos de la cuenca
CretAcico-Terciaria en Patagonia septentrional (Argentina), con
consideraciones sobre la estratigrafia del Fortin Gral. Roca (Prov.
de Rio Negro). Ameghiniana (BuenosAires) 7 (1), 1-56.
Bertels, A., 1979. Paleobiogeografia de los foraminiferos del
CretAcico superior y Cenozoico de America del Sur. Ameghiniana
(Buenos Aires) 16 (3-4), 273-356.
Biddle, K. T., Uliana, M. A., Mitchum, R.M., Jr., Fitzgerald, M. G.,
and Wright, R. C., 1986. The stratigraphic and structural evolution of the central and eastern Magallanes Basin, southern South
America. In: Foreland Basins (Edited by Allen, P. A., and Homewood, P.). International Association of Sedimentologists, Special
Publication 8, 41-61.
Blasco, G., Nullo, F., and Proserpio, C., 1979. Aspidoceras en
Cuenca Austral, Lago Argentino, Prov. de Santa Cruz. Revista
A sociaci6n Geol6giea Argentina 34 (4), 282-293.
Bir6-Bagoczky, L., 1982. Revisi6n y redefinici6n de los "Estratos
de Quiriquina', Campaniano-Maastrichtiano,en su localidad tipo
en la Isla Quiriquina, 36 35' let. S., Chile, Sudam~rica, con un
perfil complementario en Cocholgue. Acres, Ill Congreso Geol6gieo
Chileno, Concepci6n, A29-A64.
Bourgois, J., Totmsaint, J. F., Gonzalez, H., Azema, J., Calle, B.,
Desmet, A., Murcia, L. A., Acevedo, A. P., Parra, E., and Tournon,
J., 1987. Geological history of the Cretaceous ophiolotic complexes
of northwestern South America (Colombian Andes). Tectonophysics 143,307-327.
Bracaccini, I. O., 1960. Lineamientos principales de la evoluci6n
estructural de la Argentina. Petrotecnia (Buenos Aires) 10 (6}, 5769.
Bristow, C. R., 1973. Guide to the Geology of the Cuenca Basin,
Southern Ecuador. Ecuadorian Society of Geologists and Geophysicists, Quito, 54 p.
Bristow, C. R., 1975. The age of the Cayo Formation, Ecuador.
Newsletters of Stratigraphy 4 (2), 169-173.
Bueno, R., 1979. A geologic section between Bucaramanga and La
Uribe: Middle Magdalena Valley, 1971. In: Geological Field

Trips, Colombia, 1959-1978, pp. 325-348. Colombian Society of


Petroleum Geologists and Geophysicists, BogotA.
Biirgl, H., 1961a. Geologic history of Colombia. Revista Academia
Colombiana de Ciencias Exactas Fi~icas y Naturales 1 (43), 1-81.
Btirgl, H., 1961b. Geologia de los alrrededores de Ortega, Tolima.
Unioersidad de Santander (Bucaramanga), Boletin de Geologla 8,
21-38.
Btirgl, H., 1961c. Sedimentaci6n ciclica en el geosinclinalCretaceo
de la Cordillera Oriental de Colombia. Bolet~n de Geologia
(Bogota) 8 (1-3), 85-118.
Biirgl, H., and Dumit Tob6n, Y., 1954. El Cretaceo Superior en la
regi6n de Gira rdot. Boletin de Geologia (Bogot~i) 2 (1), 23-48.
Burton, R., Kendall, C. G., and Lerche, I., 1987. Out of our depth:
On the impossibility of fathoming eustasy from the stratigraphic
record. Earth Science Reviews 24,237-277.
Caceres, H., and Margfoy, P., 1985. Cuenca del Putumayo, provincia petrolifera meridional de Colombia. In: H Simposio
Exploraci6n Petrolera en las Cuencas Subandinas, Vol. I. Asociaci6n Colombiana de Ge61ogosy Geofisicos del Petroleo, BogotA.
Campbell, C. J., 1970. Guide to the Puerto Napo Area, Eastern
Ecuador. Ecuadorian Society of Geologists and Geophysicists,
Quito, 40 p.
Campbell, C. J., and Btirgl, H., 1965. Section through the eastern
Cordillera of Colombia, South America. Bulletin of the Geological
Society of A merica 76,567-590.
Canard, C. L., Gallo, J., and Shannon, P. J., 1964. Geology of the
Southern Upper Magdalene Valley: A Progress Report. International Petroleum (Colombia) Ltd., Report 1574.
Canfield, R. W., Bonilla, G., and Rebbins, R. K., 1982. Sacha oil
field of Ecuadorian Oriente. Bulletin of the American Association
of Petroleum Geologists 66,1076-1090.
Cardozo, E., and Ramirez, C., 1985. Ambientes de dep6sito de la
Formaci6n Rosablanca: Area de Villa de Leiva. In: Proyecto Cre~cieo (Edited by Etayo-Serna, F.). lngeominas (Bogota), Publicaci6n Especial 16, XIII.1-XIII.13.
Carozzi, A. V., Bercowski, F., Redriguez, M., S~nchez, M., and
Vonesch, T., 1981. Estudio de microfacies de la Formaci6n Chachao (Valanginiano), Provincia de Mendoza. Acres, VIII Congreso
Geol6gico Argentino, San Luis 2,545-565.
Catchcart, J. B., and Zambrano, F., 1967. Roca Fosfdtica en
Colombia (con una Secci6n sobre Fosfatos de Turmeque, Boyacd,
pot Pedro Mofica). Boletin de Geologia (BogotA) 15 (1-3), 92 p.
Cazau, L. B., and Uliana, M. A., 1972. E1 Cret~cico Superior
continental de la Cuenca Neuquina. Acres, V Congreso Geol6gico
Argentino, Buenos Aires 3, 131-163.
Cecioni, G., 1955. Edad y facies del Grupo Springhill en Tierra del
Fuego [Instituto Geol6gico Publicaci6n 6]. Universidad de Chile,
Anales Facultad de Ciencias Fisicas y Matemdticas 12,243-256.
Cecioni, G., 1957. Cretaceous flysch and molasse in Departamento
Ultima Esperanza, Magallanes Province, Chile. Bulletin of the
American Association of Petroleum Geologists 41,538-564.
Cecioni, G., 1979. Darwin's Navidad Embayment, Santiago
Region, Chile, as a model of the southeastern Pacific shelf. Journal
of Petroleum Geology 2 (3), 309-321.
Charrier, R., and Lahsen, A., 1969. Stratigraphy of I,ate Cretaceotm-early Eocene, Seno Skyring-Strait of Magellan area,
Magallanes Province, Chile. Bulletin of the American Association
of Petroleum Geologists 53,568-590.
Cobbing, E. J., Pitcher, W. S., Wilson, J. J., Baldock, J. W., Taylor,
W. P., McCourt, W., and Shelling, N. J., 1981. The Geology of the
Western Cordillera of Northern Peru. Institute of Geological
Sciences (London), Overseas Memoir 5,143 p.
Corrigan, H. T., 1979. The geology of the Upper Magdalena Basin
(northern portion). In: Geological Field Trips, Colombia, 19591978, pp. 221-249. Colombian Society of Petroleum Geologists and
Geophysicists, BogotA.
Cruzado Castafieda, J., 1980. Zonaci6n del Campaniano, Maestrichtiano y Daniano en el noroeste del Perti. Boletin de la

C r e t a c e o u s p a l e o g e o g r a p h y a n d d e p o s i t i o n a l cycles of w e s t e r n S o u t h A m e r i c a

Cushman, J. A., and Hedberg, H. D., 1941. Upper Cretaceous foraminifera from Santander del Notre, Colombia, SA. Contributions
to the Cushman Laboratory of Foraminiferal Research 17, pt. 4,
(232), 79-100.
Dalziel, I. W. D., 1981. Back-arc extension in the southern Andes:
A review and critical reappraisal. Philosophical Transactions of
the Royal Society of London, Series A 300,319-335.
Dalziel, I. W. D., de Wit, M. J., and Palmer, K. F., 1974. A fossil
marginal basin in the southern Andes. Nature 250, 291-294.
Dalmayrac, C., Laubacher, G., and Marocco, R., 1980. Geolagie des
Andes Peruviennes. Travaux et Documents de I'ORSTOM (Paris)
122, 501 p.
De Graciansky, P. C., Deroo, G., Herbin, J. P., Montadert, L.,
Muller, C., Schaaf, A., and Sigal, J., 1984. Ocean-wide stagnation
episode in the Late Cretaceous. Nature 308,346-349.
Dellap6, D., Pando, G., and Volkheimer, W., 1979, Estratigrafia y
palinologta de las Formaciones Mulichinco, Agrio, y Grupo La
Amarga, al sur de Zapala. Actas, VII Congreso Geoldgico Argentino,Ncuqu~n 1,593-607.
Del Solar, C., 1982. Ocurrencia de hidrocarburos en la Formacion
Vivihn, nororiente Peruano. In: I Simposio Exploracidn Petrolera
en las Cueneas Subandinas, Vol. 1. Asociaci6n Colombiana de
Ge61ogos y Geofisicos del Petroleo, BogotA.
Digregorio, J. H., 1972. Neuqu~n. In: Geologla Regional Argentina (Edited by Leanza, A. F.), pp. 439-506. Academia Nacional de
Ciencias, C6rdoba.
Digregorio, J. H., and Uliana, M. A., 1980. Cuenca Neuquina. In:
Geologia Regional Argentina, Segundo Simposio, pp. 985-1032.
Academia Nacional de Ciencias, C6rdoba.
Digregorio, R. E., Gmisano, C. A., Gutierrez Pleimling, A. R., and
Minniti, S. A., 1984. Esquema de la evoluci6n geodin~mica de la
Cuenca Neuquina y sus implicancias paleogeogr~kficas. Actas, IX
Congreso Geoldgico Argentino, San Carlos de Bariloche 2, 147162.
Dott, R. M., Jr., Winn, F. D., Jr., and Smith, C. M. L., 1982.
Relationship of late Mesozoic and early Cenozoic sedimentation to
the tectonic evolution of the southernmost Andes and Scotia Arc.
In: Antarctic Geoscience (Edited by Craddock, C.), pp. 193-203.
University of Wisconsin Press, Madison, WI, USA.
Duque-Caro, H., 1984. Structural style, diapirism, and accretionary episodes of the Sinu-San Jacinto terrane, southwestern
Caribbean borderland. In: The Caribbean-South American Plats
Boundary and Regional Tectonics (Edited by Bonini, W. E.,
Hargraves, R. B., and Shagam, R.). Geological Society of America,
Memoir 162,303-316.

413

Colombia). In: Proyecto Creaicico (Edited by Etayo-sorna, F.).


Ingeominas(BogotA),Publicaci6n Especial 16, XIX.1-XIX.20.
Faucher, B., and Savoyat, E., 1973. Equlsse g6ologique des Andes
de l~quateur. Revue de G~ographie Physique et de G~ologie
Dynamique (Paris), S~rie 2 15 (1-2), 115-142.
Feininger, T., 1975. Origin of petroleum in the Oriente of Ecuador.
Bulletin of the American Association of Petroleum Geologists 59,
1166-1175.
Feininger, T., 1982. The metamorphic "basement" of Ecuador.
Bulletin of the Geological Society of A meriea 93, 87-92.
Feininger, T., and Bristow, C. R., 1980. Cretaceous and Paleogene
geologic history of coastal Ecuador. Geologisehe Rundschau 69,
849-874.
Feininger, T., and Seguin, M. K., 1983. Simple Bouguer gravity
anomaly field and the inferred crustal structure of continental
Ecuador. Geology 11,40-44.
Feo-Codecido, G., 1972. Contribuci6n a la estratigrafia de la cuenca Barinas-Apure. Memorias, IV Congreso Geoldgico Venezolano,
Caracas 2,773-790.
Fernandez Garrasino, C. A., 1982. Algunos rasgos geol6gicos de la
cuenca Amaz6nica Ecuatoriana. Actas, V Congreso Latinoa mericano de Geologia, Buenos A ires 1, 81-95.
Feruglio, E., 1949. Descripcidn Geoldgiea de la Patagonia I, H, lll.
Direcci6n General de YPF, Buenos Aires.
Fischer, A. G., 1956. Desarrollo geol6gico del noroeste peruano
duranto el Mesozoico. In: I Congreso Nacional de Geologia, Lima.
Boletln de la Sociedad Geol6gica del Pert~ 30,177-190.
Flores, M. A., Malumi~m, N., Masiuk, V., and Riggi, J. C., 1973.
Estratigrafla CretAcica del subsuelo de Tierra del Fuego. Revista
Asociacidn Geol6gica Argentina 28 (4), 407-437.
Ford, A., and Houbolt, J. J. H. C., 1963. The Mierofacies of the Cretaceoue of Western Venezuela. E.J. Brill, Leiden, The Netherlands,
56 p., 55 pls.
Forero, H., and Sarmiento, L., 1985. La facies evaporftica de la
Formaci6n Paja en la regi6n de Villa de Leiva. In: Proyecto Cretdcico (Edited by Etayo-Serna, F.). Ingeominas (Bogota), Publicaci6n Especial 16, XVII.1-XVII.16.
Frutes, J., 1981. Andean tectonics as a consequence of sea-floor
spreading. Tectonophysics 72, 21-32.
Fuentes, R., 1980. Tendencia regional de la porosidad en la Formaci6n Chonta, Cuenca Maraf6n. Boletin de la Sociedad Geo16gica del Peru 67, 35-51.

Egyed, L., 1956. The change of the earth's dimensions determined


from paleogeographic data. Geofisica Pura e Applicata (Milano)
33, 42-48.

Gabela, V. H., 1985. Campo Carlo Lim6n, Llanos orientales de


Colombia. In: II Simposio Exploraci6n Petrolera en Ins Cueneas
Subandinas, Vol. 1. Aseciaci6n Colombiana de Ge61ogos y Geofisicos del Petroleo. Bogota.

Etayo Serna, F., 1979. Zonation of the Cretaceous of Central


Colombia by Ammonites. Ingeominas (BogotA), Publicaci6n Geo16gica Especial 2,186 p.

Gaenslen, G., 1962. A discussion of the Cretaceous stratigraphy of


the seuth-west Barinas mountain front. Asociacidn Venezolana de
Geolagia Mineria y Petroleo, Bolet~nlnformaciones 5 (3), 65-74.

Etayo Serna, F., 1985. Paleontologta estratigr~fica del Sistema


CretAcico en la Sierra Nevada del Cocuy. In: Proyecto Cretdcico
(Edited by Etayo-Serna, F.). Ingeominas (BogotA), Publicaci6n
Especial 16,XXIV.1-XXIV.47.

Galvis, N., and Rubiano J., 1985. Redefinici6n estratigr~tfica de la


Formaci6n Arcabuco, con base en el an~lisis facial In: Proyecto
Cretdcico (Edited by Etayo-Serna, F.). Ingeominas (Bogota),
Publicaci6n Especial 16, VII.I-VII.16.

Etayo Soma, F., Renzoni, G., and Barrero, D., 1976. Conternos
Sucesivos del Mar Cretaceo en Colombia. Memorias, l Congreso
Colombiano de Geologfa, Bogotd, 217-252.

Garcia Jarpa, F., Ghosh, S., Rond6n, F., Fierro, I., Sampol, M.,
Benedetto, G., Medina, C., Odreman, 0., S~nchez, T., and Useche,
A., 1980. Correlaci6n estratigr~fica y slntssis paleoambiental del
Cretaceo de los Andes Venezolanos. Boletin de Geologta (Caracas)
14 (26), 3-88.

Fabre, A., 1983a. La subsidencia de la Cuenca del Cocuy


(Cordillera Oriental de Colombia) durante el Cretaceo y el Terciario, II: Egquema de evoluci6n tectbnica. Geologia Norandina 8,
21-2.
Fabre, A., 1983b. La subsidencia de la Cuenca del Cocuy
(Cordillera Oriental de Colombia) durante el Cretaceo y el Terciario, I: Estudio cuantitativo de la subsidencia. Geologia Norandina 8, 49-61.
Fabre, A., 1985. Dinamica de la sedimentaci6n cretAcica en la
regi6n de la Sierra Nevada del Cocuy (Cordillera Oriental de

Ghosh, S. K., 1984. Late Cretaceous condensed sequence, Venezuelan Andes. In: The Caribbean-South American Plate Boundary
and Regional Tectonics (Edited by Bonini, W. E., Hargraves, R. B.,
and Shagam, R.). Geological Society of America, Memoir 162, 317324.
GonzAlezde Juana, C., lturralde de Azorena, J. M., and Picard, X.,
1980. Geologia de Venezuela y de sue Cueneas Petrol~feras. Ediclones Foninves,Caracas, 2 vols., 1031 p.
Govea, C. R., and Aguilera, H. B., 1980. Geologia de la Cuenca del
Putumayo. Boletin Geoldgico (Buearamanga) 14 (28), 45-71.

SAES 1/4--(]

414

C.E. MACELLARI

Groeber, P., 1929. Linens Fundomentales de la Geologia del Neuqu$n, sur de Mendoza y Regiones Adyacentes. Direccibn de Minas,
Geologia e Hidrologia (Buenos Aires), Publicacidn 58,109 p.
Groeber, P., 1946. Observaciones geoldgicas a la largo del meridiane 70", I: Hoja Chos Malal. Revista Asociacidn Geoldgica
Argentina 1,177-208.
Groeber, P., 1953. Andico. In: Geografla de la Repilblica Argentina. GAEA (Sociedad Argentina de Estudios Geogr~fieos) 2,349510.
Guevara, C., 1980. E1Grupo Casma del Pert~ central entre Trujillo
y Mala. Boletin de la Sociedad Geoldgica del Peril 67, 73-83.
Gulisano, C. A., Gutierrez Pleimling, A. R., and Digregorio, R. E.,
1984. An/tlisis estratigrafico del intervalo Tithoniano-Valanginiano (Formaciones Vaca Muerta, Quintuco y Mulichinco) en el suroeste de la Provincia del Neuqudn. Actas, IX Congreso Geoldgico
A rgentino, San Carlos de Bariloche 1,221-235.
Gulisano, C. A., and Legarreta, L., 1987. Andlisis secuencial del
Jurdsico, CretAcico y Eoterciario de la Cuenca Neuquina, Argentina [resumen]. Actas, X Congreso Geoldgico Argentine, San
Miguel de Tucumdn 5, 16-17.
Gust, D. A., Biddle, K. T., Phelps, D. W., and Uliana, M. A., 1985.
Associated Middle to Late Jurassic volcanism and extension in
southern South America. Tectonophysics 116,223-253.
Hallam, A., Bir6-Bagoczky, L., and Pdrez, E., 1986. Facies
analysis of the Lo Valdds Formation (Tithonian-Hauterivian) of
the High Cordillera of central Chile, and the palaeogeographic
evolution of the Andean Basin. GeologicalMagazine 123 (4), 425435.

Hi~nicken, M., Charrier, R., and Lahsen, A., 1980. Baculites


(Lytoceratina) de la base de la Formaci6n Fuentes (Campaniano
medio-superior) de la Isis Riesco, Provincia de Magallanes, Chile.
Academina Nacional de Ciencias (Cdrdoba), Boletin 53 (3-4), 221235.
Irving, E. M., 1975. Structural Euolution of the Northernmost
Andes, Colombia. U.S. Geological Survey, Professional Paper 846,
42 p.
Jacquin, T., 1987. Analogies et differences entre les "black shales"
du Cr~tac~ inf~rieur et supdrieur de rAtlantique de Sud. Bulletin
de la Soci~t~ G~ologique de France 1987 (4), 705-713.
Jenks, W. F., 1948. Geology of the Arequipa Quadrangle. Institute
Geol6gico del Perth, Boletin 9,204 p.
Jenkyns, H. C., 1980. Cretaceous anoxic events: From continents
to oceans. Journal of the Geological Society (London) 137, 171188.
Julivert, M., 1958. Geologia de la zona tabular entre San Gil y
Chiquinquird, Cordillera Oriental, Colombia. Hniversidad de
Santander (B ucaramanga) , Boletin de Geologia 2, 33-45.
Julivert, M., 1961. Geologia de la vertiente W de la Cordillera
Oriental en el sector de Bucaramanga. Unwersidad de Santander
(Buearamanga), Boletln de Geologia 8, 39-42.
Julivert, M., 1962. Estudio sedimentol6gico de la parte alta de la
Formaci6n Guadalupe al E de Bogota (Cretacico Superior). Universidad de Santander (Bucaramanga), Boletin de Geologio 10, 2554.

Hancock, J. M., 1961. The Cretaceous system in northern Ireland.


Quarterly Journal of the Geological Society (London) 117, 11-36.

Julivert, M., 1963. Estratigrafia y sedimentologia de la parte


inferior de la Formacion Guaduas al S de la Sabana de Bogot~
(Cordillera Oriental). Uniuersidad de Santander (Bucaramanga),
Boletin de Geologia 12, 85-99.

Hancock, J. M., and Kauffman, E. G., 1979. The great transgressions of the Late Cretaceous. Journal of the GeologicalSociety
(London) 136, 175-186.

Julivert, M., 1968. Lexique Stratigraphique International: Vol. 5,


Amdrique Latine; Fasicule 4a, Colombia. Centre National de la
Recherche Scientifique (France), 650 p.

Haq, B. V., Hardenbol, J., and Vail, P. R., 1987. Chronology of


fluctuating sea levels since the Triassic. Science 235,1156-1166.

Katz, H. R., 1963. Revision of Cretaceous stratigraphy in Patagonian Cordillera of Ultima Esperanza, Magallanes Province,
Chile. Bulletin of the American Association of Petroleum Geologists 47,506-524.

Hays, J. D., and Pitman, C. S., III., 1973. Lithospheric plate motion, sea-level changes and climatic and ecological consequences.
Nature 246,18-22.
Hedberg, H. D., 1931. Cretaceous limestones as petroleum source
rocks in northwestern Venezuela. Bulletin of the American
Association ofPetroleum Geologists 15,229-244.
Hedberg, H. D., and Sass, L. C., 1937. Synopsis of the geologic
formations of the western part of the Maracaibo, Venezuela.
Boletin de Geologia y Mineria (Caracas) 1 (2-4), 71-112.
Henderson, W. G., 1979. Cretaceous to Eocene volcanic arc activity in the Andes of northern Ecuador. Journal of the Geological
Society (London) 136,367-368.
Hinterwimmer, G. A., Meissinger, V. E., and Soave, L. A., 1984.
An~ilisis de facies, porosidad y diagdnesis de una secuencia de
plays, Formaci6n Springhill, en el sondeo Pueste Barros, Provincia
de Santa Cruz. Actas, IX Congreso Geoldgico Argentine, San
Carlos de Bariloche 5, 136-145.
Hubach, E., 1945. La Formacion Caqueza, regi6n de Caqueza
(oeste de Cundinamarca, Colombia). Compilacion Estudios Geoldgicos Oficiales, Colombia (Bogota) 6, 23-26.
Hubach, E., 1957. Estratigrafia de la Sabana de Bogota y alrrededores. Boletin de Geologia (Bogota) 5 (2), 96-112.
Huerta-Kohler, T., 1982. Exploraci6n petrolffera en la Cuenca
Ucayali, Oriente Peruano. In: H Simposio Exploracidn Petrolera
en las Cuencas Subandinas, Vol. 1. Asociaci6n Colombians de
Gedlogos y Geoflsicos del Petroleo, BogotA.
Himicken, M., 1955. Dep6sitos Neocretacios y Terciarios del
Extreme SSW de Santa Cruz (Cuznca Carbonifera de Rio Turbio).
Revista Institute Nacional de Investigaciones Ciencias Naturales
Bernardino Rivadavia (Buenos Aires), Ciencias Geoldgicas 2 (1),
161 p.
Hfinicken, M., 1965. Algunos cefaldpodos supracretAcicos de Rio
Turbio (Santa Cruz). Uniuersidad Nacional de Cdrdoba, Reuista
Facultad Ciencias Exactas F~sicasy Naturales 26 (1-2), 49-99.

Kauffman, E. G., 1979. Cretaceous. In: Treatise on Invertebrate


Paleontology - - Part A, Introduction (Edited by Teichert, C., and
Robinson, R. A.), pp. A418-A487. University of Kansas Press,
Lawrence, KS, USA.
Knechtel, M. M., Richards, E., and Rathbun, M. V., 1947. Mesozoic
Fossils of the Peruvian Andes. Johns Hopkins University (Baltimore), Studies in Geology 15, 150 p.
Kohl, E., and Blissenbach, E., 1962. Las eapas rojas del Cretaceo
superior-Terciario en la regi6n del Curse Medio del Rio Ucayali,
Oriente del Perth. Boletin de la Sociedad Geoldgica del Perk 39, 137.
Kummel, B., Jr., 1948. Geological reconnaissance of the Contamana region, Peru. Bulletin of the Geological Society of America
59,1217-1266.
Larson, R. L., and Pitman, W. C., 1972. World-wide correlation of
Mesozoic magnetic anomalies, and its implications. Bulletin of the
Geoiogical Society of A merica 83(12),3645-3661.
Laubacher, G., 1978. Geologic des Andes Peruuiennes. Memoires
ORSTOM (Paris) 95,216 p.
Leanza, A. F., 1963. Patagoniceras gen. nov. (Binneyitidae) y otros
ammonites del CretAcico superior de Chile Meridional con notas
acerca de su posici6n estratigr~fica. Academia Nacional de Ciencias (Cdrdoba) , Boletln 43,203-225.
Leanza, A. F., 1967. Descripci6n de la fam~a de Placenticeras del
CretAcico superior de Patagonia Austral con consideraciones
acerca de su posicibn estratigr~fica. Unwersidad Nacional de
Cdrdoba, Revista Facultad de Ciencias Exactas F~sicas y Naturales
25 (3-4), 93-107.
Leanza, A. F., 1969. Sobre el descubrimiento de depdsitos del piso
Coniaciano en Patagonia Austral y descripcidn de una nueva
especie de Ammonites (Peroniceras santacrucense n. sp.). Academia Nacional de Ciencias (Cdrdoba), Boletin 47 ( 1), 5-20.

C r e t a c e o u s p a l e o g e o g r a p h y a n d d e p o s i t i o n a l cycles of w e s t e r n S o u t h A m e r i c a
Leanza, A. F., 1972. Andes Patag6nicos Anstrales. In: Geologia
Regional Argentina (Edited by Leauza, A. F.), pp. 689-706. Academia Nacional de Ciencias, Cbrdoba.
Leanza, H. A., 1973. Estudio sebre los cambios faciales de los
estratos limltrofss JurAsico-CretAcicos entre Loncopu6 y Pict~n
Louftl, Provincia de Neuqu~n, Reptlblica Argentina. Revista
A sociaci6n Geol6gica Argentina 28, 97-132.
Leanza, H. A., 1981. The Jurassic-Cretaceous boundary beds in
west-central Argentina and the ammonite zones. Neues Jahrbuch
ffir Gzologie und Pal6ontologie 161 (1),62-92.
Leanza, H. A., and Hugo, C. A., 1977. Sucesi6n de ammonites y
edad de la Formaci6n Vaca Muerta y sincr6nicas entre los paralelos 35* y 40 l.s., Cuenca Neuquina-Mendocina. Revista Asociaci~n
Geol6gica Argentina 32,248-264.
Leanza, H., Marchese, H., and Riggi, J. C., 1977. Estratigrafla del
Grupo Mendoza, con especial referencia de la Formaci6n Vaca
Muerta entre los paralelos 25* y 40 s., Cuenca Neuquina-Mendocina. Revista Asociaci6n Geol6gica Argentina 32 (3), 190-208.
Legarreta, L., and Kozlowski, E., 1981. Estratigrafla y sedimentelogia de la Formaci6n Chachao, Prov. de Mendoza. Actos, VIII
Congreso Geol6gico Argentino, San Luis 2,521-593.
Legarreta, L., Kozlowsky, E., and Bol, A., 1981. Esquema estratigrbfico y distribucionesde facies del Grupo Mendoza en el ~mbito
submendocino de la Cuenca Neuqtfina. Acta8, VIII Congrezo Geo16gioo Argentino, San Luis 3, 389-409.
Lesta, P., Ferello, R., and Chebli, G., 1980a. Chubut Extraandino.
In: Geologia Regional Argentina, Segundo Simposio. Academia
Nacional de Ciencias (C6rdoba) 2,1307-1387.
Losta, P., Mainardi, E., and Stubelj, R., 1980b. Plataforma Continental Argentina. In: Geologia Regional Argentina, Segundo
Simposio. Academia Nacional de Ciencias (C6rdoba) 2, 15771601.
Lozada T., F., and Endara B., P., 1982. Estudio de litofacies de la
zona Arenisca-M-I de la Formaci6n Nape en la Cuenca Oriental.
In: H Simposio Exploraci6n Petrolera en /as Cuencas Subandinas,
Vo. 2. Asociaci6n Colombiana de Ge61ogos y Geoflsicos del Petroleo, BogotA.
Macellari, C. E., 1985. El limits CretAcicofrerciario en la
Peninsula AntArtica y en el sur de Sudam~rica: Evidencias macropaleontol6gicas. Actas, VI Congreso Latinoamericano de Gzologia,
Bogot6 1,268-278.

415

Marocco, R., 1978. Geologie des Andes Peruuiennes. Memoire


ORSTOM (Paris) 94,195 p.
Marquillas, R. A., Boso, M. A., and Salfity, J. A., 1984. La
Formaci6n Yacoraite (CretAcicosuperior) en el norte Argentino, al
sur del paralelo 24. Actas, IX Congreso Geol6gico Argentino, San
Carlos de Bariloche 2,300-310.
Martinez, M. V., 1980. Sedimentaci6n del Cretaceo y ocurrencia de
petroleo en el delta del orients Peruano. Boletin de la Sociedad
Geol6gica del Per~ 67, 85-96.
Maughan, E. K., Zambrano, F. O., Mojica, G., Abozaglo, M.,
Pach6n, F. P., and Dur~n, R. R., 1979. Paleontologic and Stratigraphic Relations of Phosphate Beds in Upper Cretaceous Rocks of
the Cordillera Oriental, Colombia. U.S. Geological Survey, Open
File Report 79-1525, 97 p.
Maze, W. B., 1984. Jurassic La Quinta Formation in the Sierra de
Perij~, northwestern Venezuela: Geology and tectonic environment of red beds and volcanic rocks. In: The Caribbean-South
American Plate Boundary and Regional Tectonics (Edited by
Bonini, W. E., Hargraves, R. B., and Shagam, R.). Geological
Society of America, Memoir 162,263-282.
McLaughlin, D. H., 1924. Geology and physiography of the Peruvian Cordillera, Departments of Junin and Lima. Bulletin of the
Geological Society of A merica 35, 591-632.
Miller, J. A., 1979. Geology of the Llanos Basin and adjaceltt
Eastern Cordillera. In: Geological Field Trips, Colombia, 19591978, pp. 349-379. Colombian Society of Petroleum Geologists and
Geophysicists, Bogota.
Miller, M., Lanussol, D., and Marinelli, R. V., 1982. Paleoambiente de depositaci6n de la Formacibn Springhill en el yacimiento Cariad6n Salto - - Cuenca Austral, Provincia de Santa
Cruz. I Congreso Nacional de Hidrocarburos, Petroleo y Gas, Exploraci6n, Buenos Aires, 231-245.
Mitchum, R. M., Jr., and Uliana, M. A., 1985. Seismic stratigraphy
of carbonate depositional sequences, Upper Jurassic-Lower Cretaceous, Neuqu6n Basin, Argentina. In: Seismic Stratigraphy II
(Edited by Berg, O. R., and Woolverton, D. G.). American Association of Petroleum Geologists, Memoir 39,255-274.
Mojica, J., and Macia, C., 1983. Caracteristicas estratigrbficas y
edad de la Formaci6n Yavi, Mesozoico de la regi6n entre Prado y
Dolores, Tolima, Colombia. Universidad Nacional de Colombia
(Bogotd), Geologia Colombians 12, 7-32.

Macellari, C. E., in press. Late Cretaceous Kossmaticeratidae


(Ammonoidea) from the Magallanes Basin (Chile). Journal of
Paleontology.

Mombrti, C., Uliana, M., and Bercowski, F., 1979. Estratigrafta y


sedimentologia de las acumulacionesbiocarbonSticasdel Cretacico
Surmendocino. Actas, VH Congreso Geol(~gico A rgentino, Neuqu~n
1,685-700.

Macellari, C. E., and DeVries, T. J., 1987. Late Cretaceous upwelling and anoxic sedimentation in northwestern South America.
Palaeogeography, Palaeoclimatology, and Paloeoecology 59, 279292.

Morales, L. G., and others, 1958. General geology and oil occurrences of the Middle Magdalena Valley, Colombia. In: Habitat of
Oil: A Symposium (Edited by Weeks, L. G.). American Association of Petroleum Geologists, Special Publication, 641-695.

MaceUari, C. E., Barrio, C. A., and Manassero, M. J., in review.


Upper Cretaceous to Paleocene depositional sequences and sandstone petrography of southwestern Patagonia (Argentina and
Chile). Journal of South American Earth Sciences.

Morris, R. C., and Aleman, A. R., 1975. Sedimentation and tectonics of middle Cretaceous Cops Sombrero Formation in northwest Peru. Boletin de la Sociedad Geol6gica del Per~ 48, 49-69.

Malumi~n, N., 1968. Foraminfferos del CretAcico Superior y Terciario del subsuelo de la Provincia de SantA Cruz, Arg. Ameghiniana(BuenosAires) 5(6), 191-227.
Malurnian, N., and Ramos, V. A., 1984. Magmatic intervals,
transgression-regresssion cycles, and oceanic events in the Cretaceons and Tertiary of southern South America. Earth and Planetary Scienoe Letters 67,228-237.
Malumibn, N., Masiuk, V., and Riggi, J. C., 1971. Micropaleontologia y sedimentologia de la perforaci6n SC-1, Sta. Cruz:
Su importancia y correlaciones. Reuista Asociacidn Geoldgica
Argentina 26 (2), 175-208.
Malumi~n, N., Nullo, F. E., and Ramos, V. A., 1983. The CretaceousofArgentina, Chile, Paraguay, and Uruguay. In: Phancrozoic Geology of the World, Vol. II13: Mesozoic Geology (Edited by
Moullade, M., and Nairn, A. E. M.), pp. 265-304. Elsevier, Amsterdam.

Myers, J. S., 1974. Cretaceous stratigraphy and structure, western Andes of Peru. Bulletin of the American Association of
Petroleum Geologists 58, 424-487.
Natland, M. L., Gonz~lez, E., Cafi6n, A., and Ernst, M., 1974. A
System of Stages for Correlation of Magallanes Basin Sediments.
Geological Society of America, Memoir 139,126 p.
Newell, N., 1949. Geology of the Lake Titicaca Region, Peru and
Bolivia. Geological Society of America, Memoir 36, 111 p,
Notestein, P. B., Hubman, C. W., and Bowler, J. W., 1944. Geology
of the Barco concession, Republic of Colombia, South America.
Bulletin of the Geological Society of A merica 55,1165-1216.
Nullo, F., Proserpio, C., and Biasco de Nullo, G., 1981. Estratigrafla del CretAcico superior en el Cerro Indies y alrrededores,
Prov. de Santa Cruz. Actas, VIII Congreso Geol6gico Argentino,
San Luie 3,373-387.
Olsson, A., 1934. Cretaceous of the Amotope Region. Bulletin of
American Paleontology 20 (69), 104 p.

416

C . E . MACELLARI

Pardo, A., and Ztifiiga, F., 1976. Estratigrafla y evolucidn


tectbnica de la regi6n de la Selva del Pertt, Parte If: Mesosoico y
Cenozoico. Memorias, II Congreso Latinoamericano de Geologia,
Caracas 2,588-608.
Pdrez, G., and Salazar, A., 1978. Estratigrafla y facies del Grupo
Guadalupe. Universidad Nacional de Colombia (Bogotd), Geologia
Colombiona 10, 7-85.
Pdrez, V. E., and Bolivar, A. L., 1985. Cuenca Llanos orientales:
Exploracidn petrolera en la subcuenca Apiay-Ariari. In: II Simposio Exploracidn Petrolera en las Cuencas Subandinas, Vol. 1.
Asociacidn Colombiana de Gedlogos y Geofisicos del Petroleo,
BogotA.
Pitman, W. C., 1978. Relationship between sea-level change and
stratigraphic sequences of passive margins. Bulletin of the Geological Society of A merica 89,1389-1403.

Reyes, F. C., 1972. Correlaciones en el Cret~cico de la Cuenca


Andina de Bolivia, Pert~ y Chile. Revista Tdcnica YPFB I (2-3),
101-144.
Reyment, R. A., 1981, Colombia. In: Aspects of Mid-Cretaceous
Regional Geology {Edited by Reyment, R. A., and Bengston, P.),pp.
175-195. Academic Press, N e w York.
Reyment, R. A., and Morner, N. A., 1977. Cretaceous transgressions and regressions exemplified by the South Atlantic.
Palacontological Society of Japan, Special Papers 21,247-261.
Riccardi, A. C., 1971. Estratigrafla en el Oriente de la Bahia de La
Lancha, Lago San Martin, Sta. Cruz, Argentina [Geologia 61].
Revista Museo de La Plata 7,245-318.
Riccardi, A. C., 1977. Berriasian invertebrate fauna from the
Springhill Formation of southern Patagonia. Neucs Jahrbuch fiir
Geologie undPaldontologie 155 (2), 216-252.

Ploszkiewiez, J. V., and Orchuela, I. A., 1982. lnterpretaci6n


geol6gica de anomalias stsmicas en el Yacimiento Loma La Lata,
Provincia del Neuqu~n I Congreso Nacional de Hidrocarburos,
Petroleo y Gas, Exploracidn,Buenos Aires, 283-292.

Riccardi, A. C., 1979. El gdnero Calycoceras Hyatt (Ammonitina,


CretAcico superior) en Patagonia Austral. Obra Centeneario
Museo de La Plata (Paleontologia) 5, 63-72.

Ploszkiewicz, J. V., and Ramos, V. A., 1978. Estratigrafla y tectdnica de ia Sierra de Payaniyeu. Revista Asociaci6n Geoldgica
Argentina 32 (3), 209-226.

Riceardi, A. C., 1984. Las asociaciones de amonitas del Jur#isico,


Crethcico de la Argentina. Actas, IX Congreso Geoldgico Argentino, San Carlos de Bariloche 4,559-595.

PolanIa, H., and Rodriguez, G., 1978. Posibles turbiditas del CretAcico inferior (Miembro SocotA) en el ~trea de Anapoima (Cundinamarca). Universidad Nacional de Colombia (BogotA), Geologia Colombians 10, 87-113.

Riccardi, A. C., 1987. Cretaceous paleogeography of southern


South America. Palaeogeography, Palaeoclimatalogy, and Palaeoecology 59,169-195.

Ramirez, C., and Campos, V. C., 1972. Geologia de la regi6n de la


Grita-San Cristhbal, Estado Tachira. Memorias, IV Congreso
Geoldgico Venezolano, Caracas 2,861-893.

Riccardi, A. C., Aguirre Urreta, M. B., and Medina, F. A., 1987.


Aconeceratidae (Ammonitina)from the Hauterivian-Albian of
southern Patagonia. Palaeontographica (Stuttgart) AI96, 105185.

Ramos, E. D., and Ramos, V. A., 1979. LOs ciclos magm~ticos de la


Repttblica Argentina. Actas, VII Congreso Geoldgico Argentino,
Neuqudn 1,771-796.

Riccardi, A. C., and Rolleri, E. O., 1980. Cordillera Patag0nica


Austral. In: Geologia Regional Argentina, Scgundo Simposio.
Academia Nacional de Ciencias, COrdoba 2, 1173-1306.

Ramos, V. A., 1976. Estratigrafla de los lagos La Plata y Fontana,


Provincia del Chubut, Repttblica Argentina. Actas, I Congreso
Geoldgico Chileno, Santiago de Chile 1, A43-A64.

Richards, R. H., 1968. Cretaceous section in Barco area of


northeastern Colombia. Bulletin of the American Association of
Petroleum Geologists 52, 2324-2336.

Ramos, V. A., 1978. E1 vulcanismo del Cr~!tacico inferior de la


Cordillera Patagbnica de Argentina y Chile. Actas, VII Congreso
Geoldgico Argentino, Neuqudn 1,423-435.

Rivers, R., 1979. Zonas faunisticas del Cretaceo de Lima. Boletin


de la Sociedad Geoldgica del Peril 62,19-24.

Ramos, V. A., 1979. Tectdnica de la regidn del rio y lago Belgrano,


Cordillera Patagonica, Argentina. Actas, II Congreso Geoldgico
Chileno, Arica 1, B1-B32.
Ramos, V. A., 1982. Geologia de la regi6n del Lago Cardiel, Prov.
de Sta. Cruz. Revista Asociocidn Geoldgica Argentina 37 (1), 2349.
Renz, O., 1959. Estratigrafia del Cretaceo en Venezuela Occidental. Bolet~n de Geologia (Caracas) 5 (10), 3-48.
Renz, O., 1960a. Geologia de la parte sureste de la Peninsula de la
Guajira (RepOblica de Colombia). Boletln de Geologia (Caracas),
Publicacidn Especial 3, 317-349.
Renz, O., 1960b. Remarks on the Barquisimeto Trough. Asociocidn Venezolana de Gedlogos Mineros y Petroleros, Boletin Informativo 3 (6), 155-162.
Renz, O., 1977. The lithologic unite of the Cretaceous in western
Venezuela. Memorias, V Congreso Geoldgico Venezolano, Caracas
1, 45-58.
Renz, O., 1981. Venezuela. In: Aspects of Mid-Cretaceous
Regional Geology (Edited by Reyment, R. A., and Bengsten, P.),
pp. 197-220. Academic Press, New York.
Renzoni, G., 1967. Geologia del Macizo de Quetame. Boletln de
Geologia(Bogotd) 5, 75-127.
Renzoni, G., 1985a. Paleoambientes de la Formaci6n Tambor en la
Quebrada Pujamanes. In: Proyecto Cretdcico (Edited by EtayoSerna, F.). Ingeominas (BogotA), Publieacidn Especial 16, II1.1III.18.
Reuzoni, G., 1985b. Paleoambientes en Ins Formaciones Arcabuco
y Cumbre de la Cordillera de los Cobardes. In: Proyevto Cretcicico
(Edited by Etayo-Serna, F.). Ingeominas (BogotA), Publicaci6n
Especial 16, X.1-X-14.

Rivers, R., Petersen, G., and Rivers, M., 1975. Estratigrafla de la


Costa de Lima. In: 1II Congreso Peruano de Geologia. Boletin de
la Sociedad Geol6gica del Perd 45,159-186.
Robles, D., 1982. El desarroHo de la Formacibn Springhi[l en la
Cuenca de Magallanes. In: 1 Congreso Nacional de H idrocarburos,
Petrdleo y Gas, Exploracidn, Bueaos A ires, 293-312.
Robles, D. E., 1984. Los depocentros de la Formacidn Springhill en
el norte del Tierra del Fuego. Actas, IX Congreso Geoldgico Argentino, San Carlos de Bariloche 1,449-457.
Rod, E., 1959. Formacion Capacho en Trujillo septentrional y e n
Lara suroccidental. Boletln de Geologia (Caracas) 5 (10), 49-66.
Rod, E., and Maync, W., 1954. Revision of Lower Cretaceous
stratigraphy of Venezuela. Bulletin of the American Association of
Petroleum Geologists 38,193-283.
Rodriguez, G. A., and Chalco, A., 1975. Cuenca HuaHaga, Resefia:
Geoldgica y posibilidades petroliferas. Boletin de la Sociedad
Geoldgica del Psril 45, 187-212.
Rollins, J. F., 1965. Stratigraphy and Structure of the Guajira
Peninsula, Northwestern Venezuela and Northeastern Colombia.
University of Nebraska, Studies (New Series) 30,102 p.
Russo, A., and Flores, M. A., 1972. Patagonia austral extraandina. In: Geologia Regional Argentina (Edited by Leanza, A.
F.), pp. 707-725. Academia Nacional de Ciencias, Cdrdoba.
Russo, A., Ferello, R., and Chebli, G., 1980. Llanura Chaco Pampeana. In: Geologia Regional Argentina, Segundo Simposio.
Academia Nacional de Ciencias, C6rdoba 1, 139-183.
Rtumomano, 1., and Velarde, H., 1982. Geologia petrolera de la
Cuenca Barinas-Apure. In: Simposio Exploracidn Petrolera en las
Cuencas Subandinas, Vol, 1. Asociacidn Colombiana de Gedlogos
y Geoflsicos del Petroleo, BogotA.

Cretaceous paleogeography and depositional cycles of western South A merica


Salfity, J. A., 1982. Evoluci6n paleogeogrbfica del Grupo Salta
(CretAcico-Eog6nico), Argentina. Actas, V Congreso Latinoamericano de Geologia, Buenos A ires 1, 11-26.
Salfity, J. A., Marquillas, R. A., Gardeweg, M., Ramirez, C., and
Davidson, J., 1985. Correlaciones en el CretAcico superior del
norte de Argentina y Chile. Actas,/V Congreeo Geol6gico Chileno,
Antafagazta 1,654-667.
Sauer, I. W., 1971. Geologie yon Ecuador. Gebruder Borntraeger,
Berlin, 333 p.
Skarmeta, J., and Charrier, R., 1976. Geologia del sector fronterizo de Ays~n entre los 45 y 46 de latitud sur, Chile. Actas, VI
Congreso Geoldgico Argentino, Buenos Aires 1,267-286.
Schlanger, S. O., and Jenkyns, H. C., 1976. Cretaceous oceanic
anoxic events: Causes and consequences. Geologic M~jnbouw 55
(3-4), 179-184.
Scott, K. M., 1966. Sedimentology and dispersal pattern of a
Cretaceous flysch sequence, Patagonian Andes, southern Chile.
Bulletin of the A merican Association of Petroleum Geologists 50,
72-107.
Seminario, S., and Guizado, J. J., 1976. Slntesis bioestratigr~fica
de la regi6n de la selva del Perd. Memorize, H Congreso Latinoamericano de Geologia, Caracas 2,881-899.
Sigal, J., 1969. Quelques acquisitions recents concernant la
chronostratigraphie des formations sedimentaires de l~quateur.
Revista Espafwla de Micropaleontolgia 1 (2), 205-236.
Sigal, J., Grekoff, N., Singh, N. P., Caflbn, A., and Ernst, M., 1970.
Sur l'age et les affinit6s *gondwaniennes" de microfaunes floraminff~res et ostracodes) malgaches, indiennes et chilienes au
semmet du Jurassique et b la base du Cr6tac~. Comptss Rendue de
l'A cad~rnic des Sciences (Paris) 271, 24-27.
Sliter, W. V., 1976. Cretaceous foraminifers from the southwest
Atlantic Ocean, Leg 36, Deep Sea Drilling Project. In: Initial
Reports of the Deep Sea Drilling Project (Edited by Barker, P. F.,
and Dalziel, I. W. D., et a[.). U.S. Government Printing Office
(Washington) 36, 519-573.
Smith, C. H., 1977. Sedimentology of the Late Cretaceous
(Santonian-Maestrichtian) Tres Pasos Formation, Ultima Esperanza District, Southern Chile. Unpublished MSe thesis, University
of Wisconsin (Madison, WI), 126 p.
Seto, F. V., 1979. Facies y ambientes deposicionales cretacicos
area central sur de ia Cuenca Marafi6n. Boletin de la Sociedad
Geoh~gica del Per?z 60, 233-250.
StAinforth, R. M., 1969. The concept of sea floor spreading applied
to Venezuela. Asociaci6n Venezolana de Ge6logos Mincros y
Petroleros(Caracas),Boletinlnformativo 12(8),257-274.
Steinmann, G., 1929. Geologie vonPeru. Heidelberg,448 pp.
Stephan, J. F., 1977. El contacto cadena Caribe-Andes Meride~os
entre Carora y El Tocuyo (EstAdo Lara). Memorias, V Congreso
Geoldgico Venezolano, Caracas 11,789-816.
Stinnesbeck, W., 1986. Zu den faunistischen and Palokologischen
verhaltnissen in der Quiriquina Formation {Maastrichtium)
Zentral-Chiles. Palaeontographica(Stuttgart) 194(4-6),99-237.
Stipanicic, P., and Rodrigo, F., 1970. El diastrofismo Eo y Mesocretbcico en Argentina y Chile con referencias a los movimientos
jurAeicos de la Patagonia. Actas, IV Jornadas Geoldgicas Argentinas(??city?) 2, 337-352.
Sm~rez, M., 1979. A late Mesozoic island arc in the southern
Andes, Chile. Geological Magazine 116, 191-201.
Sutton, F. A., 1946. Geology of the Maracaibo Basin, Venezuela.
Bulletin of the American Association of Petroleum Geologists 30,
1621-1741.
Taborda, B., 1979 (1965). The geology of the De Mares Concession.
In: Geological Field Trips, Colornbla, 1959-1978, pp. 119-159.
Colombian Society of Petroleum Geologists and Geophysicists,
BogotA.
Thomas, C. R., 1949. Geology and petroleum exploration in
Magailanes Province, Chile. Bulletin of the American Association
of Petroleum Geologists 33,1553-1578.

417

Travis, R. B., Gonzblez, G., and Pardo, A. 1976. Hydrocarbon


potential ofcoastsl basins of Peru. In: Circum-Pacific Energy and
Mineral Resources (Edited by Halbouty, M., et al.). American
Association of Petroleum Geologists, Memoir 25, 331-338.
Trump, G. W., and Salvador, A., 1964. Guidebook to the Geology of
Western Tdchira. Asociaci6n Vene~lana de Geologla, Minerla y
Petroleo, Caracas, 25 p.
Tschopp, J. L., 1953. Oil explorations in the Oriente of Ecuador,
1938-1950. Bulletin of the American Association of Petroleum
Geologists 37, 2303-2347.
Uliana, M. A., and Biddle, K.T., in press. Mesozoic-Cenozoic
paleogeographic and geodynamic evolution of southern South
America. H Symposium on South Atlantic Evolution, Rio de
Janeiro (July 1987).
Uliana, M. A., and Dellap~, D. A., 1981. Estratigrafia y evolucibn
paleoambientAl de la sucesibn Maastrichtiano-Eoterciaria del engolfamiento Neuquino (Patagonia septentrional). Aetas, VIII Congreso Geol6gico A rgentino, San Luis 3,673-711.
Uliana, M. A., Dellap~, D. E., and Pando, G. A., 1975. Estratigrafia, distribuci6n y g~nesis de Ins sedimvntitas rayosianas (CretAcico inferior de Ins Provincias de Neuqu~n y Mendoza). Anales,
H Congreso Ibero-A mericano de Geologia Econ6rnica, Buenos A ires
1,151-196.
Uliana, M. A., Dellap~, D. A., and Pando, G. A., 1977. AnAlisis
estratigrAfico y evoluci6n del potencial petrolifero de las Formaciones Mu]ichinco, Chachao y Agrio (Prov. del Neuqu~n y Mendoza). Pstrotecnia (Buenos Aires) 1-2, 31-46; 3, 25-33.
Urien, C. M., and Zambrano, J. J., 1973. The geology of the basins
of the Argentine continental margin and Malvinas Plateau. In:
The Ocean Basins and Margins, Vol. 1: South Atlantic (Edited by
Nairn, A. E. M., and Stehli, F. G.), pp. 135-170. Plenum Press,
New York.
Vail, P. R.,and others 1977. Seismic stratigraphy and global
changes of sea level, Parts 1-11. In: Seismic Stratigraphy - Applications to Hydrocarbon Exploration (Edited by Payton, C. E.).
American Association of Petroleum Geologists, Memoir 26, 49212.
Vail, F. R., 1987. Seismic stratigraphic interpretation procedure.
In: Atlas of Seismic Stratigraphy (Edited by Bally, A. W.). American Association of Petroleum Geologists, Studies in Geology 27, 110.
Vail, P. R., Hardenbol, J., and Todd, R. E., 1984. Jurassic unconformities, chronostratigraphy, and sea level changes from seismic
stratigraphy and biostratigraphy. American Association of Petroleum Geologists, Memoir 36, 129-144.
Valencio, D. A., Giudich, A., Mendia, J. E., and Oliver, G. J., 1976.
Paleomagnetismo y edades K-Ar del Subgrupo Pirgua, Provincia
de SaltA, Argentina. Actas, VI Congreso Geol6gico Argentino,
Buenos Aires 1,527-542.
Vicente, J. C., 1981. Elementos de la estratigrafla Mesozoica sur
Peruana. In: Cuencas Sedimentarias del Jurdsico y Cretdcico de
Amdrica del Sur, Vol. 1 (Edited by Volkeimer, W., and Musacchio,
E. A.), pp. 319-351. Buenos Aires.
Vicente, J. C., Beaudoin, B., Chhvez, A., and Lebn, I., 1982. La
Coenca de Arequipa (sur de PerO) durante el Jur~sico-CretAcico
inferior. Actas, V Congreso Latinoamericano de Geologia, Buenos
Aires 1,121-153.
Vilela, C. R., and Csaky, A., 1968. Las turbiditas en los sedimentos
cretAcicos de la regibn de Lago Argentino (Provincia de Santa
Cruz). Actas, II1 Jornada8 Geoldgicaz Argentinas, Buenos A ires 1,
209-225.
Volkheimer, W., and Salas, A., 1975. Estudio palinolbgieo de la
Formaci6n Huitrin. CretAcico de la Cuenca Neuquina en su localidad tipo. Actas, VI Congreso Geoldgico A rgentino, Buenos Aires 1,
431-453.
Waddell, M. G., 1982. A Depositional Model for the Monserrato
Formation (Upper Cretaceous) of the Neiva Basin, Colombia, South
America. Unpublished MS thesis, University of South Carolina
(Columbia), 83 p.

418

C.E. MACELLARI

Weaver, C. E., 1931. Paleontology of the Jurassic and Cretaceous of


West-Central Argentina. University of Washington (Seattle),
Memoir 1,595 p.
Wichmann, R., 1927. Sobre las facies lacustre Senoniana de los
estratos con dinosaurios y su fauna. Academia Nacional de
Ciencas (COrdoba), Boletin 30 (1-4), 383-406.
Wilson, J. J., 1963. Cretaceous stratigraphy of central Andes of
Peru. Bulletin of the American Association of Petroleum Geologists
47, 1-34.
Wilson, T. J., 1983. Stratigraphic and Structural Evolution of the
Ultima Esperanza Foreland Fold-Thrust Belt, Patagonian Andes,
Southern Chile. Unpublished PhD dissertation, Columbia University (New York), 360 p.
Winn, R. D., Jr., and Dott, R. H., Jr., 1979. Deep-water fanchannel conglomerates of Late Cretaceous age, southern Chile.
Sedimentology 26 (2), 203-228.
Winslow, M. A., 1980. Mesozoic and Cenozoic Tectonics of the Fold
and Thrust Belt in Southernmost South America and Stratigraphic
History of the Cordilleran Margin of the Magallanes Basin.
Unpublished PhD dissertation, Columbia University (New York),
254 p.

Zambrano, E., V~squez, E., Duval, B., Latreille, M., and


Coffinieres, B., 1971. Sintesis paleogeogr~fica y petrolera del
occidente de Venezuela. Memorias, IV Congreso Geoldgico Venezolano,Caracas 1,483-552.
Zambrano, J. J., 1974. Cusncas ssdimentarias en el subsuelo de la
provincia de Buenos Aires y zonas adyacentes. Revista Asociacidn
Geoldgica Argentirm 29 (4), 443-469.
Zambrano, J. J., 1980. Comarca de la Cuenca CretAcica de Colorado. In: Geolog~a Regional Argentina, Segundo Simposio. Academia Nacional de Ciencias, C6rdoba 2,1033-1069.
Zegarra, C. J., 1982. Resultado y futuro de la exploraci6n petrolera
en las Cuencas Subandinas del PerO. In: Simposio Exploraeidn
Petrolera en las Cucncas Subandinas, Vol. 3. Asociacibn Colomhiana de Ge61ogosy Geoflsicos del Petroleo, BogotA.
Zilli, N. J. M., Orchuela, I. A., Dellap~, D., and Otano, R., 1979.
An~lisis de las Formaciones Quintuco y Loma Montosa en el sector
centro-oriental de la Cuenca Neuquina. Actas, VII Congreso Geoldgico Argentino, Neuqu~n 1,609-615.

Você também pode gostar