Está en la página 1de 29

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/236578737

Jurassic-Eocene tectonic evolution of Maracaibo basin, Venezuela

Chapter · January 1995

CITATIONS READS
91 1,344

2 authors, including:

Paul Mann
University of Houston
353 PUBLICATIONS   7,735 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Linking tectonics, regional gravity, structure, depositional response, and sedimentary facies in the Permian Basin View project

Age and provenance of Paleozoic rocks from the Araya-Paria Peninsula, northeastern Venezuela based on new U-Pb zircon ages and geochemical data View project

All content following this page was uploaded by Paul Mann on 19 May 2014.

The user has requested enhancement of the downloaded file.


Jurassic–Eocene Tectonic Evolution of
Maracaibo Basin, Venezuela

Jairo Lugo Paul Mann


Lagoven, S.A. Institute for Geophysics
Caracas, Venezuela University of Texas at Austin
Austin, Texas, U.S.A.

Abstract

T hree main phases occurred in the Jurassic–Eocene geologic history of Lake Maracaibo: (1) Jurassic rifting
related to separation of the North and South American continents; (2) Early–Late Cretaceous passive
margin subsidence following rifting and creation of oceanic crust between North and South America; and (3)
Paleocene–Eocene foreland basin subsidence following the oblique collision of a Pacific-derived Caribbean
plate with the South American passive margin. The distribution of Jurassic red beds and volcanic rocks in deep
wells beneath the Maracaibo basin suggest that they were deposited in north-northeast trending rift basins
separated by horst blocks of Paleozoic metasedimentary rocks. This structural grain was subsequently reacti-
vated during Cenozoic collisional and strike-slip deformation. Lithology and isopach patterns suggest that
Barremian–Santonian shallow marine carbonates in the Maracaibo basin formed in a passive margin setting
above Precambrian and Paleozoic crust thinned during the Jurassic rifting event. Isopach patterns show that
the northwest-trending Mérida arch was a positive feature in Barremian–Santonian time. The tectonic origin of
the Mérida arch and similar arches in Colombia and Ecuador is unclear, but they may be related to reactivation
of underlying Precambrian or Paleozoic structural grains during passive margin subsidence.
Isopach maps show that the Paleocene–Eocene depocenter of the Maracaibo basin was an asymmetric,
elongate trough containing up to 7 km of Paleocene–Eocene clastic rocks. The main Paleocene–Eocene
depocenter is flanked on its north and northeast margin by a partially exposed south-southwest verging fold
and thrust belt. Isopach mapping suggests a northwest-southeast migration of the depocenter from Paleocene
to middle Eocene time. Previous work on sandstone composition, paleocurrent measurements, and seismic
reflections of prograding clinoforms all support the idea that the basin formed as a foreland depression in front
of a west-southwest verging fold and thrust belt along the eastern side of the basin. This model contrasts with a
previous interpretation that deltaic systems in a passive margin setting south and southwest of the basin were
the main source areas. A similar style of younger foreland basin has been previously identified over a distance
of 1000 km in central and eastern Venezuela and Trinidad. Eocene–Recent ages of foreland basin sedimenta-
tion in these areas suggest time-transgressive oblique collision of an exotic Pacific-derived Caribbean plate
along the northern passive margin of continental South America.

Resumen

T res fases principales existieron en la historia geológica del Lago de Maracaibo durante el período
Jurásico-Eoceno: rift durante el Jurásico, relacionado con la separación de Norte y Sur América; subsi-
dencia del margen pasivo durante el Cretácico Inferior a Superior, seguida a la fase de rift y creación de
corteza oceánica entre Norte y Sur América; y subsidencia de la cuenca tipo foreland a continuación de la
colisión oblicua de la placa del Caribe, derivada del Pacífico, contra el margen pasivo de Sur América. La
distribución de capas rojas y rocas volcánicas de edad Jurásica encontradas en pozos profundos de la cuenca
de Maracaibo indica sedimentación en cuencas tipo rift con tendencia nor-noreste separadas por pilares
tectónicos que presentan rocas meta-sedimentarias de edad Paleozoica. Estas antiguas estructuras con
dirección nor-noreste serían subsecuentemente reactivadas durante la colisión y deformación transcurrente del
Cenozoico. La litología y distribución de espesores de rocas de edad Barremiense-Santoniense, principalmente
carbonáticas de mar poco profundo, en la cuenca de Maracaibo indican que estas rocas se formaron en un
margen pasivo por encima de la corteza Precambrica y Paleozoica, adelgazada durante el evento de rift del
Jurásico. Los patrones isópacos muestran que el arco de Mérida fué un rasgo positivo desde el Barremiense

Lugo, J., and P. Mann, 1995, Jurassic–Eocene tectonic evolution of Maracaibo basin, 699
Venezuela, in A. J. Tankard, R. Suárez S., and H. J. Welsink, Petroleum basins of South
America: AAPG Memoir 62, p. 699–725.
700 Lugo and Mann

hasta el Santoniense. El origen tectónico del arco de Mérida y otros arcos similares en Colombia y Ecuador no
está claro pero quizás esten relacionadonados con la reactivación de estructuras precambricas o paleozoicas
pre-existentes, las cuales existían durante la subsidencia del margen pasivo.
Los mapas isópacos, basados en datos sísmicos y de pozos, muestran que el depocentro de la cuenca de
Maracaibo durante el Paleoceno-Eoceno fue una depresión elongada asimétrica, de 50 a 80 km de ancho, con
un espesor de hasta 7 km de rocas clásticas de edad Paleoceno-Eoceno. El principal depocentro esta limitado en
el norte y noreste por pliegues de cabalgamiento parcialmente expuesto con vergencia hacia el suroeste. Los
mapas isópacos de siete unidades sísmicas indican una migración del depocentro del noroeste hacia el sureste
durante el período Paleoceno-Eoceno Medio. Estudios anteriores sobre composición de arenas, mediciones de
paleocorrientes y clinoformes progradantes mostrados en los ejemplos de reflección sísmica soportan la idea
que la cuenca se formó como una depresión tipo foreland en frente de fallas de cabalgamiento con vergencia
hacia el sur suroeste a lo largo del lado este de la cuenca. Este modelo propuesto contrasta con una inter-
pretación previa en la cual los sistemas deltaicos en una cuenca pasiva localizada al sur y sureste constituyeron
las principales fuentes de sedimentos. Una cuenca tipo foreland similar y más joven ha sido previamente identi-
ficada a una distancia de 1000 km en Venezuela central y oriental y en Trinidad. Sedimentación en cuencas tipo
foreland de edad Eoceno a Reciente en estas áreas indican transgresión en el tiempo y colisión oblicua de una
placa Caribe exótica, derivada de la placa del Pacifico, a lo largo del margen pasivo al norte del continente
suramericano.

INTRODUCTION the style of sedimentation in a passive margin


setting and the tectonic control on this sedimenta-
Maracaibo basin covers an area of about 30,000 km2 tion by the development of the Mérida arch.
and is located on the Maracaibo tectonic block, a trian-
• Paleocene–Eocene foreland basin subsidence
gular fault-bounded area in northwestern Venezuela
followed the oblique collision of an exotic Pacific-
(Figure 1). The basin is one of the most prolific hydro-
derived Caribbean plate with the South American
carbon-producing basins in the world; it has produced
passive margin. This event is the main objective of
more than 35 billion bbl of mainly medium-light oil
our paper because most previous workers have
during the last 50 years (Miller et al., 1958; Zambrano et
attributed Paleocene–Eocene clastic sedimentation
al., 1971; Bockmuelen et al., 1983). The production zone is
to a north-facing passive margin. Based on
structurally and stratigraphically complex. Production
geophysical and geologic data presented here, we
comes from hundreds of reservoirs in Tertiary deltaic
suggest that Paleocene–Eocene clastic sedimenta-
and shore zone deposits and fractured Cretaceous
tion was related to south-southwestward over-
carbonate rocks. Extensive exploration for new reservoirs
thrusting of the Caribbean plate over the Creta-
in the Maracaibo basin over the past 50 years has
ceous passive margin of South America. We use
produced tens of thousands of kilometers of high-quality
these results to better constrain tectonic models for
seismic reflection data and thousands of production
Cenozoic eastward migration of the Caribbean
wells. This wealth of data provides us with the opportu-
plate.
nity to better understand the petroleum generating and
storage system in the basin as well as the tectonic
The complex post-Eocene structural and stratigraphic
controls on basin sedimentation.
history of the Maracaibo basin is beyond the scope of this
The objective of this paper is to illustrate three main
paper. The interested reader is referred to Lugo (1991) for
Jurassic–Eocene phases of basin evolution based on a
this phase of basin evolution.
compilation of previous geologic data and an original
compilation of seismic reflection and well data, shown in
Figure 2. These data were kindly made available for use
in this study by Lagoven, S.A., and Maraven, S.A., of TECTONIC SETTING
Caracas, Venezuela. The three phases of basin evolution
are as follows: Plate-Scale Setting
• Jurassic rifting accompanied rifting between the Most workers now accept that the Caribbean region
North and South American continents. We show originated in the eastern Pacific and was transported into
the distribution of rift-related rocks encountered by its present position between the North and South
deep drilling in the basin and discuss the funda- American plates along large-scale strike-slip faults and
mental control that rift-related faults had in the pre- oblique subduction zones of Cenozoic age (e.g., Burke,
Oligocene structural and stratigraphic develop- 1988; Pindell and Barrett, 1990; Montgomery et al., 1992).
ment of the basin. This allochthonous area, referred to here as the
“Caribbean plate,” includes a central Late Cretaceous
• Early–Late Cretaceous passive margin subsidence oceanic plateau in the Colombian and Venezuelan basins
followed rifting and the creation of oceanic crust (Bowland and Rosencrantz, 1988), a Paleozoic(?) conti-
between North and South America. We document nental microplate in northern Central America (Gordon,
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 701

Figure 1—(A) Location map of major crustal tectonic


features of northwestern South America and southern
Central America. Arrows indicate convergence directions
of Caribbean Sea (Caribbean plate) and Pacific Ocean
(Nazca plate) beneath northwestern South America. Earth-
quake epicenters with depths less than 75 km and magni-
tudes >5.0 (Richter scale) are from the ISC catalogue.
Abbreviations: MB, Maracaibo basin; FB, Falcón basin; GP,
Guajira Peninsula; MAB, Magdalena basin; BFZ, Boconó
fault zone; EAFZ, East Andean fault zone of Pennington
(1981); PFZ, Pallantanga fault zone of Winter and Lavenu
(1989); SMBFZ, Santa Marta–Bucaramanga fault zone. Note
the splaying of the elevated areas of the northern Andes in
the region of the Maracaibo basin. (B) Map of crustal types
of northwestern South America and southern Central
America. Southeastward subduction of thicker than Figure 2—(A) Seismic reflection data used in this study. (B)
average oceanic crust of the Caribbean oceanic plateau Well data used in this study. (All data provided by Lagoven
and Panama island arc may account for the width of the and Maraven and used with permission.)
active deformation belt in the northern Andes. Arches of
Cretaceous age that affected passive margin sedimenta-
tion are from Macellari (1988).
702 Lugo and Mann

1990), a fringing Early Cretaceous–Recent island arc, and Maracaibo to the Caribbean Sea. Uplift of the northern
Paleogene–Recent back-arc basins formed as this island Andean ranges in Miocene and later time has caused the
arc migrated northeastward and eastward (Mann et al., Maracaibo basin to become a vast sink for clastic
1991). Elongate belts of marine clastic rocks found at the sediment eroded from the surrounding mountains.
contact between passive margin sequences and the Neogene sedimentary rocks in the eastern part of the
Caribbean assemblage have been interpreted by Pindell lake reach a maximum thickness of 3 km. For this reason,
and Barrett (1990) and Pindell (1991) as foreland basin most Mesozoic and early Tertiary structures and rocks in
deposits that record progressive latest Cretaceous– the Maracaibo basin are deeply buried and accessible
Cenozoic eastward displacement of the Caribbean plate only through indirect geophysical methods and deep
relative to the Americas. Although these authors had wells.
speculated that early Cenozoic deposition in the
Maracaibo basin was in a foreland basin setting, they
were limited in their interpretation of the Maracaibo
basin area of western Venezuela by the availability of DATABASE USED IN THIS STUDY
detailed published studies of the basin stratigraphy and Seismic reflection data used in this study include
structure. about 7000 km of migrated lines acquired by Lagoven
and Maraven during 12 different seismic acquisition
campaigns from 1969 to 1987. These data were repro-
Regional-Scale Setting cessed in-house to suppress noise and recover primary
No simple plate boundary can be drawn between reflections (Figure 2A). The average spacing of seismic
island arc systems along the Pacific margin of Central lines is 2 km in the lake area, although some intensively
and South America and the active right-lateral strike-slip drilled areas of the lake have line spacings less than 1
plate boundary present along the Caribbean coast of km. In the Sierra de Perijá area to the west of the lake,
South America (Figure 1A). Shallow seismicity (<75 km line spacing is 5 km.
in depth) is diffuse, and the topography defines four Seismic data are fair at shallow depths, improve to
major chains of the northern Andes Mountains that are good below the Eocene–Miocene unconformity, but dete-
two to three times as wide as the Andes of southern riorate downward to Paleozoic and Jurassic rocks that
Colombia and Ecuador. The Maracaibo basin is located form the acoustic basement of the basin. Data quality
within this broad band of active deformation in the degradation above basement is mainly related to
center of the Maracaibo tectonic block, a triangular complex, steeply dipping faults and bedding, as well as
lithospheric wedge of the northwestern South America to shallow mud channels that are particularly common
continent (Figure 1B) that is actively “escaping” along the eastern coast of the lake where water depths
northward over thinner crust of the Caribbean Sea along are only a few meters.
the right-lateral Boconó and left-lateral Santa Marta– In this study, 383 wells were used, 81 of which have
Bucaramanga strike-slip fault zones (Mann and Burke, velocity information that includes vertical seismic
1984). Strike-slip offset of the Maracaibo block relative to profiles and sonic logs (Figure 2B). Seismic velocity
the surrounding South America plate is limited to less determinations from wells constrain all depth conver-
than 100 km along the Boconó and Santa Marta–Bucara- sions of seismic reflection lines.
manga fault zones (Salvador, 1986) and appears to have
begun in middle–late Miocene time.
The driving force of the diffuse tectonic activity over
such a wide area may be related to shallow southeast- SUBSURFACE SEISMIC MAP UNITS IN
ward subduction of the Caribbean plate to depths of THE MARACAIBO BASIN
several hundred kilometers beneath this part of north-
western South America (van der Hilst and Mann, 1994). Lugo (1991) combined seismic data with wells and
Such shallow subduction may have produced Laramide- identified four seismic megasequences correlatable to
style uplifts of the Andean chains (Kellogg and Bonini, formations previously established in outcrops of the
1982; Kellogg, 1984) and tectonic escape of the Maracaibo Mérida Andes east of the basin, in outcrops of the Sierra
block in the overriding South America plate (Mann and de Perija west and southwest of the basin (see Gonzalez
Burke, 1984). de Juana et al., 1980, for stratigraphic review), and from
The Maracaibo basin is generally synclinal and subsurface studies of well cores in the lake area itself
contains up to 7 km of unmetamorphosed Jurassic– (Figure 3). These megasequences are further subdivided
Recent sedimentary rocks. It presently forms a topo- into seismic map units along discontinuity surfaces inter-
graphic depression filled by Lake Maracaibo (maximum preted from seismic profiles (Figure 4). Distinctive
water depth 31 m) and is flanked by high (>2 km) seismic characteristics of each seismic map unit are
Andean ranges on three sides. To the east and southeast discussed in more detail by Lugo (1991).
of the basin is the Mérida Andes of Venezuela and to the Age control on seismic map units is provided by pale-
west and southwest is the Sierra de Perija of Venezuela ontologic and palynologic studies of well cuttings tied to
and Colombia. The northern edge of the basin is much seismic lines. Age diagnostic species are summarized in
lower in elevation, with a small strait connecting Lake Figure 5, and examples of poor or absent age controls are
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 703

Figure 3—Names of forma-


tions defined by Gonzalez de
Juana et al. (1980) and used in
this study for the subsurface
and surrounding outcrops of
the Maracaibo basin (see inset
for type areas). Seismic
megasequences defined from
the lake area in this study
using seismic reflection and
well data are shown in far
right column. See Figure 4 for
correlations between seismic
map units defined in this
study and formations.

discussed in the text. Misties between well logs and • Megasequence 2—A 700-m-thick (maximum)
seismic map units made in the first interpretation were section of mostly pelagic Campanian shale mapped
corrected in later iterations that were facilitated by access as one unit, K7. This unit correlates to the Colon
to the results of 12 different and commonly overlapping Formation and grades up into Paleocene neritic
surveys. Lugo (1991) contoured isopach and isochron limestone and shale, mapped as one seismic unit
maps of various seismic map units in the following four designated P1.
megasequences overlying a largely acoustic basement:
• Megasequence 3—A 7000-m-thick (maximum)
• Megasequence 1—A 300-m-thick (maximum) section of predominantly marine and shallow
section of Aptian–Santonian mainly shallow water marine Eocene clastic rocks mapped as seven
platform limestone was subdivided into six seismic Eocene units (E1–E7). Precise ages of Eocene units
map units designated K1–K6 (Figure 4). Corre- are poorly known because of the scarcity of marine
sponding formations and members in the lake area microfauna in most of the shallow marine section
include the Rio Negro, Apon, Lisure, Maraca, La (Figure 5). However, the ages of units P1 and E7
Luna, and Socuy (Figure 3). are well established by late Paleocene and middle
704 Lugo and Mann

Figure 4—Age, lithology, and depositional environments of formations described in the Maracaibo basin area and their
correlation to informal seismic megasequences and map units defined in this study. Formation names are defined by
Gonzalez de Juana et al. (1980) using stratigraphic studies of outcrops and subsurface cores. This paper focused on the
Jurassic–Eocene section; for details of the Oligocene–Recent section, see Lugo (1991).

Figure 5—Faunal and floral assemblages used in this study to calibrate the age of Cretaceous–Tertiary seismic map units.
Age of units E4 and E5 is based on appearance of Monoporites annulatus pollen species about 46.5 Ma. Late middle Eocene
planktonic foraminifera are well dated from unit E7.
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 705

Eocene benthic and planktonic foraminifera, JURASSIC RIFT SEDIMENTATION IN


respectively, and bracket the age of the intervening
and less well dated map units.
MARACAIBO BASIN
• Megasequence 4—A 10,000-m-thick (maximum) Precambrian and Paleozoic Basement
section of mainly Oligocene–Recent clastic rocks
(seismic map units O1 and M1-4) deposited in Tectonic Setting
shallow marine, fluvial, and lacustrine settings in The cratonic basement of Venezuela consists of three
an overall shallowing- and coarsening-upward distinct Precambrian provinces with metamorphic,
section. This section includes the erosional metaigneous, and metasedimentary rocks yielding radio-
products of the Miocene and younger uplift of the metric ages of 3000–1400 Ma (Feo-Codecido et al., 1984;
northern Andes that were trapped in the intermon- Bartok, 1993) (Figure 6A). These Precambrian rocks were
tane Maracaibo basin (Figure 1). rifted in Cambrian time along with rocks in the
Appalachians of North America during the opening of
the Iapetus, or the proto-Atlantic Ocean. Closure of the
Iapetus ocean and its passive margins by suturing of the
Gondwana and Laurentia continental blocks produced
the Appalachian-Ouachita-Marathon orogeny of North
America and a correlative and possibly co-linear Late
Paleozoic orogeny and regional metamorphism in north-
western South America (Pindell, 1985; Bartok, 1993).
Deformed and metamorphosed Paleozoic rocks include
shallow and deep water shale, sandstone, and limestone,
which outcrop in the core of the Andean ranges and are
found in deep wells drilled into basement in the
Maracaibo basin.
Stratigraphy
Deep well data show that the acoustic basement of
Maracaibo basin correlates with upper Paleozoic meta-
morphosed pelagic sedimentary rocks of the Mucu-
chachí Formation and its equivalents found in deep wells
along the northwest-trending Mérida arch and red beds
of the Jurassic La Quinta Formation infilling a north-
northeast-trending rift system (Figure 6B). According to
Shagam (1985), the Mucuchachí Formation is Late
Mississippian–Permian in age; however, Gonzalez de
Juana et al. (1980) conclude that the age of the
Mucuchachí Formation ranges from early to middle
Paleozoic because its stratigraphic relationships with
unmetamorphosed older Ordovician and younger Penn-

Figure 6—(A) Depth to Precambrian–Jurassic basement in


the region of Maracaibo basin (from Feo-Codecido et al.,
1984). Contour interval is 5000 ft. Note 20,000-ft (6100-m)
maximum depth to basement; steepest basement slope is
present along the southeastern Maracaibo basin. The
Mérida arch cross section is shown in Figure 9. (B)
Location of Jurassic rifts and the Merida arch in north-
western South America. Legend: gray, known Jurassic
rifts; black, outcrops of rift red beds and volcanic rocks; X,
inverted rifts that form northern Andes mountain ranges; 1,
Machiques Jurassic rift cropping out in the Sierra de Perija;
2, Uribante Jurassic rift known from deep wells in the
Maracaibo basin of Venezuela; 3, Mérida structural arch
expressed in the map pattern of the Mérida Andes and
known from deep wells and seismic reflections in the
Maracaibo and Barinas basins; 4, Jurassic Barquismeto rift
cropping out in the northwestern Mérida Andes; 5,
Jurassic Mérida rift cropping out in the Mérida Andes.
Cretaceous uplift of the Mérida arch apparently eroded the
central section of the Mérida rift.
706 Lugo and Mann

Stratigraphy
Volcanic and sedimentary rocks are included in the La
Gé Group and in the Tinacoa, Macoita, and La Quinta
formations at different type localities in Colombia and
western Venezuela (Eva et al., 1989; Bartok, 1993) (Figure
3). In the Sierra de Perija rift west of the Maracaibo basin,
calc-alkaline felsic to mafic igneous rocks of the La Gé
Group and Totumo Formation interfinger with red beds
and are cut by dikes of alkaline basalt composition
(Maze, 1984). Red beds are referred to the La Quinta
Formation in the Maracaibo basin area and Mérida
Andes (Schubert et al., 1979). Without marker horizons
and with little paleontologic control, it is difficult to
assess accurately the thickness or lateral correlation of the
red bed sequences among different areas. Nevertheless,
published stratigraphic studies reveal rapid changes
from about 500 m to more than 4500 m (Eva et al., 1989).
Rifts are inferred at the site of these mountain ranges
by the presence of red beds and related volcanic rocks
not found in outcrops or the subsurface of intervening
lowlands (Figure 6B). Sedimentary rocks consist of red
and green shale, sandstone, and conglomerate, all of
which lack marine fauna but contain plant fragments
and freshwater fish and ostracods (Schubert et al., 1979;
Figure 7—Distribution of Jurassic red beds and volcanic
rocks of the La Quinta Formation in the Uribante rift Eva et al., 1989). Sedimentary facies indicate a continental
beneath the Maracaibo basin. Circles mark locations of environment in which alluvial fans alternated with lacus-
deep wells penetrating rift-related rocks or adjacent trine and brackish water deposits (Moticska, 1975;
Paleozoic basement; ticks mark downthrown sides of Schubert et al., 1979). In the Guajira Peninsula along the
faults. Note close control of rift rocks by northeast- to Caribbean margin of Venezuela (Figure 1A), Jurassic
NNW-striking high-angle faults. These faults and the rocks were deposited in a marine environment (Renz,
Mérida arch were apparently formed during Jurassic conti- 1960) and may have formed part of a Jurassic “Hispanic
nental rifting but were later reactivated by Cenozoic corridor” or strait of open marine water that separated
collision and transpressional inversion. North and South America during the early rift phase
(Bartok, 1993). Porphyry sills, locally up to several
hundred meters in thickness, are mainly acidic in compo-
sylvanian–Permian rocks are unclear. Available evidence sition.
suggests that the metamorphic rocks of the Mucuchachí Odreman and Benedetto (1977) and Benedetto and
Formation were tectonically emplaced on Ordovician Odreman (1977) used paleontologic data to assign an
rocks and eroded by Pennsylvanian time (Gonzalez de early Middle Jurassic age to the Tinacoa Formation, a
Juana et al., 1980). Middle Jurassic age to the Macoita Formation, and a
middle Late Jurassic age to the La Quinta Formation.
Jurassic Sedimentary and Volcanic Rocks This age for the La Quinta agrees with a U-Pb radio-
metric age reported for it by Dasch and Banks (1981).
Tectonic Setting
Maracaibo Basin
The Triassic–Jurassic breakup of Pangea, previously
assembled by closure of the Iapetus ocean in late Deep wells in the Maracaibo basin have penetrated
Paleozoic time, began by the formation of a broad belt of red beds of the La Quinta Formation on the downthrown
rifts and rift-related volcanism along the present-day side on the Icotea and Pueblo Viejo fault zones, whereas
margins of the Atlantic Ocean, Gulf of Mexico, and wells to the east of the Icotea fault have penetrated only
northern South America (Pindell, 1985; Eva et al., 1989; Paleozoic rocks of the Mucuchachí Formation (Figure 7).
Bartok, 1993). In northwestern South America, four rifts This subcrop pattern is consistent with the interpretation
have been identified on the basis of field mapping and that the rocks were originally deposited in north-
deep drilling in the Maracaibo basin (Figure 6B). Three of northeast striking rifts separated by horst blocks of
the four rifts have been inverted by late Tertiary uplift of Paleozoic rocks and were subsequently preserved
the northern Andes and correspond to topographic beneath the thick Cretaceous–Cenozoic section of the
uplifts in the Mérida Andes and the Sierra de Perija. A Maracaibo basin (Bot and Perdomo, 1986). The Mérida
fourth rift has been deeply buried below Cretaceous and arch, which may follow an older Precambrian or
Cenozoic rocks of the Maracaibo basin and is known Paleozoic grain, forms a transverse structure with a more
from deep wells (Figure 7). northwest trend than the rifts.
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 707

CRETACEOUS PASSIVE MARGIN water carbonate platform that includes the


overlying Lisure (seismic map unit K3) and Maraca
SEDIMENTATION IN MARACAIBO (seismic map unit K4) formations. The three
BASIN carbonate units are collectively known as the
Cogollo Group. During deposition of the Apon and
Tectonic Setting Lisure formations, paleowater depth varied from 2
to 15 m, whereas during deposition of the Maraca
Pindell (1985), Winker and Buffler (1988), Macellari Formation, water depth fluctuated from 0 to 4 m
(1988), and Eva et al. (1989) have all proposed that (Azpiritxaga, 1991).
carbonate sedimentation in a passive margin setting
followed Middle–Late Jurassic rift-related clastic sedi- • Lisure Formation (seismic map unit K3), an
mentation and igneous activity on circum-Gulf of Mexico Aptian–Albian sandstone and calcarenite unit
and circum-Caribbean rifted margins. In the Maracaibo thinning over the Mérida arch and deposited under
basin, we compiled isopach data and subsidence data to shallow water conditions (Figure 9). Clastic input
test the hypothesis that the mixed carbonate-clastic of this unit was derived from the Mérida arch and
Cretaceous section subsided as a passive margin. South American craton to the southeast of the Mar-
Previous work by Renz (1959), Macellari (1988), Salvador acaibo basin. Southeast of the study area, this form-
(1986), and others has shown the importance of Creta- ation becomes mainly interbedded sandstone and
ceous-age uplift along the Mérida arch in controlling the shale with only a minor proportion of limestone.
distribution of carbonate and clastic units and erosional • Maraca Formation (seismic unit K4), a Ceno-
removal of part of the inverted Mérida Jurassic rift from manian fossiliferous limestone unit thinning over
the tops of the Mérida Andes (Figure 8). Macellari (1988) the Mérida arch (Figures 8B, 9). Gonzalez de Juana
pointed out that the Mérida arch is one of a subparallel et al. (1980) concluded that this distinctive unit is
set of several regional, northwest-trending arches in also found in the Barinas basin to the east of the
northwestern South America that controlled passive Mérida Andes (Figure 1A) and was deposited
margin sedimentation (Figure 1B). The tectonic origin of during a late Aptian transgression.
these arches remains unclear but may relate to reactiva-
tion of older Precambrian or Paleozoic structural grains • La Luna Formation (seismic unit K5), a
during passive margin subsidence. Turonian–Santonian anoxic black shale and
siliceous unit thinning over the Mérida arch
(Figures 8C, 9). This unit was deposited at the time
Cretaceous Megasequences of maximum flooding of the South American
The Cretaceous carbonate-clastic section in the continent during the Late Cretaceous highstand of
Maracaibo basin has been subdivided by Lugo (1991) sea level (Hancock and Kaufman, 1979). The La
into two seismic megasequences. Megasequence 1 Luna Formation is one of the most intensively
consists of five formations and one member, each equiv- studied formations in the Maracaibo basin and
alent to a seismic map unit (Figure 4). Wells were used to other parts of South America because it is known to
better distinguish individual carbonate seismic map be the source rock for most of the hydrocarbons
units that are thin and difficult to resolve on seismic discovered to date. All previous literature agrees
lines. on the anaerobic conditions of deposition (e.g.,
Blaser and White, 1984; Macellari and DeVries,
• Rio Negro Formation (seismic map unit K1), a 1987). However, paleowater depth is controversial.
Barremian basal trangressive clastic deposit found Based on faunal assemblages, Boesi and Goddard
on either side of the Mérida arch (Figure 9). This (1989) have indicated that deposition of the La
formation is characterized by fluvial and alluvial Luna Formation took place at least 1000 m below
fan facies that are thickest in the areas overlying the sea level. However, Mendez (1990) has proposed a
sites of the Uribante, Machiques, and Barquismeto much shallower depth of 50 m that was brought
Jurassic rifts (Renz, 1959; Trump and Salvador, about by upwelling of deeper water onto a
1964) (Figure 6B). In the Maracaibo basin, the Rio shallower platform. Pending further studies, we
Negro Formation is restricted to the area west of assume a minimum 150 m water depth in this
the Icotea fault and gradually thickens to the west- study for the La Luna Formation based on a model
southwest. Farther to the north on the Guajira by Byers (1977) for the minimum depth of anoxic
Peninsula, Renz (1959) identified shallow water environments created along upwelling coasts.
carbonate rocks equivalent in age to the Rio Negro • Socuy Member of the Colon Formation and
Formation. The distribution of the Rio Negro Colon Formation (seismic units K6 and K7),
Formation shows that the Mérida arch was a Campanian–Maastrichtian basal marine micritic
positive feature during the Barremian. limestone and overlying shale (Figure 8D). The
• Apon Formation (seismic map unit K2), an Aptian Socuy Member separates the underlying transgres-
shallow marine carbonate unit found on either side sive megasequence from the overlying regressive
of the Mérida arch (Figures 8A, 9). The Apon marine to shallow marine megasequence of the
Formation forms the basal unit of a stable shallow Colon Formation. In outcrops of the Mérida Andes,
708 Lugo and Mann

Figure 8—Isopachs of Cretaceous carbonate rocks deposited in a passive margin setting. (A) Isopach map of
Apon Formation (seismic map unit K2) based on seismic and well data shown in Figure 2. See Figure 3 for
regional stratigraphic correlation and nomenclature. (B) Isopach map of Maraca Formation (seismic map unit K4).
(C) Isopach map of La Luna Formation (seismic map unit K5). (D) Isopach map of Socuy Member of Colon
Formation (seismic map unit K6).
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 709

Figure 9—Regional southwest-northeast cross section of the Mérida arch based on well data (compiled by Renz, 1959). See
Figure 6A for location. Note onlap of Barremian Rio Negro Formation through Turonian–Santonian La Luna Formation onto
Paleozoic metamorphic rocks and Jurassic rift-related rocks that core the arch. The Socuy Member of the Colon Formation
is not shown, but would correspond to a thin unit at the contact between the La Luna and Colon formations (see Figure 3).

the Socuy Member is only 1–3 m thick and is Precambrian and Paleozoic rocks occupy an approxi-
composed of glauconitic and pyritic pelletal mately 100-km-wide zone in the area where the arch
packstone. Ghosh (1984) has interpreted this unit as orthogonally crosses the topographically uplifted Mérida
a condensed section. The overlying Colon Andes (Salvador, 1986). The Mérida arch continues into
Formation is composed mainly of shale. By the end the basement of the Cenozoic Barinas basin formed along
of deposition of the Colon Formation, the Mérida the southeastern flank of the Mérida Andes. Several
arch was still partially emergent in the south- studies have documented its presence in the subsurface
eastern part of the Maracaibo basin and shed of the Barinas basin (Zambrano et al., 1971; Renz, 1981;
coarse clastic sediments northward (Figure 9). Chigne and Hernández, 1990; Loaiza et al., 1990)
(Figure 1A).
Isopach Maps
Cretaceous Subsidence Rates
Isopach maps of four representative units of mega-
sequence 1 (units K2, K4, K5, K6) are shown in Figure 8 Cretaceous decompacted subsidence on top of
and support the effect of the Mérida arch on sedimenta- basement is plotted for 18 wells distributed over the
tion that was proposed by previous workers based on Maracaibo basin study area (Figure 10). For the period
well and outcrop data (Figure 9). The four units, corre- 115–80 Ma, the subsidence curve follows an initially rapid
sponding to the Apon, Maraca, and La Luna formations phase of subsidence at an average rate of 4 cm/1000 yr
and the Socuy Member of the Colon Formation, all show followed by a long-term exponential decay of subsidence
differential subsidence during Cretaceous time. All units at average rates of 0.75 cm/1000 yr. Renewed rapid subsi-
thin toward the Mérida arch, which maintained an dence of 3 cm/1000 yr characterizes the Late Cretaceous.
average N 30° W trend from Aptian to Campanian time. Late Cretaceous subsidence is followed by a period of
Total thickness changes are about 300 m from the crest to slower subsidence in the Paleocene.
the flanks of the arch. The lack of major Cretaceous In map view, subsidence patterns reflect the presence
unconformities in the Maracaibo basin and surrounding of the Mérida arch (Figure 10, inset). Above the arch,
outcrop areas (Renz, 1959; 1981; Zambrano et al., 1971; subsidence rates are relatively slow at 3 cm/1000 yr.
Bartok et al., 1981; Azpiritxaga, 1991) suggests that the Along the northwestern flank of the arch, subsidence
Mérida arch remained partially submerged during depo- rates increase to 8 cm/1000 yr. Renz (1959), Macellari
sition of seismic units K2–K7 (Aptian–Santonian). (1988), and others have pointed out that Cretaceous
The southeastern extension of the Mérida arch is sections thicken over the sites of former Jurassic rifts
visible in the outcrop pattern of the Mérida Andes where shown on Figure 6B.
710 Lugo and Mann

of the passive margin. The tectonic origin of this arch and


subparallel arches of comparable size in Colombia and
Ecuador (Macellari, 1988) is enigmatic, but may reflect
reactivation of Precambrian–Paleozoic structural grains
during subsidence.

PALEOCENE–EOCENE FORELAND
SEDIMENTATION IN MARACAIBO
BASIN
Passive Margin versus Foreland Model
Previous studies of the Paleocene–Eocene clastic
section in the lake area by van Andel (1958), Zambrano et
al. (1971), and Blaser and White (1984) suggest the
following:

• The stratigraphic section formed a southwest to


northeast gradation from fluvial deposits to deltaic
deposits and deep marine submarine fans.
• The stratigraphic section was mainly derived from
erosion of continental rocks of the South America
craton to the south and southwest.
• The stratigraphic section was deposited on a north-
sloping passive margin (Figure 11A).

Figure 10—Subsidence analysis for Cretaceous–Paleocene


sedimentary rocks in 18 wells in the Maracaibo basin. All
subsidence curves are corrected for compaction effects.
Top shows paleobathymetry based on microfauna summa-
rized in Figure 5. Inset map shows contoured subsidence
rates for the 18 wells. Note the faster rates of subsidence
along the flanks of the Mérida arch.

Summary
The presence of regionally extensive, mixed
carbonate-clastic facies patterns of exponentially
declining subsidence during Barremian–earliest
Campanian time is consistent with a passive margin
setting as proposed on the basis of more regional studies
by Pindell (1985, 1991) and Eva et al. (1989). In general,
sedimentation reflects the Late Cretaceous highstand of
sea level and regional transgression recognized in
eastern Venezuela (Gonzalez de Juana et al., 1980) and
worldwide (Hancock and Kaufman, 1979; Haq et al.,
1987). An increase in Campanian subsidence and regres- Figure 11—(A) Interpretation by previous workers of
sive facies may reflect tectonic accretion of exotic terranes continent-derived Paleocene–Eocene clastic sedimentation
in the Lake Maracaibo area. (B) Proposed interpretation for
along the Pacific margin of Colombia (Restrepo and
north-derived deltaic foreland basin sedimentation in front
Toussaint, 1988) and increased activation of intraplate of an emergent fold and thrust belt. Note relatively minor
features, including the Mérida arch (Figure 1B). component of continent-derived deltaic sedimentation in
Thickness variations suggest that the Mérida arch was comparison to major component of deltaic sedimentation
active from Aptian to Campanian time and was the derived from overthrusting of the Caribbean plate onto the
dominant structural element present along this segment passive margin of South America.
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 711

Figure 12—(A) Isopach map of Paleocene limestone and calcareous sandstone (Guasare Formation, seismic map unit P1).
(B) Isochron map of Eocene fluviodeltaic sedimentary rocks of the lower Misoa Formation (seismic map unit E2). (C)
Isochron map of Eocene fluviodeltaic sedimentary rocks of the lower Misoa Formation (seismic map unit E3). (D) Isochron
map of Eocene prodelta-estuarine sedimentary rocks of the middle Misoa Formation (seismic map unit E4). (E) Isochron map
of Eocene coarsening-upward shallow marine to fluvial sedimentary rocks of the middle–upper Misoa Formation (seismic
map unit E5). (F) Isochron map of Eocene deep marine sedimentary rocks of the Paují Formation (seismic map unit E7). Note
the northwest to southeast decrease in age of maximum foreland basin subsidence. Area to the northeast of major depocen-
ters is an exposed southwest-verging fold and thrust belt inferred to be the main sediment source for seismic map units
E4–E7. All maps are based on the seismic and well data shown in Figure 2.

We propose a contrasting interpretation—that most of Paleocene–Eocene Megasequence


the Paleocene–Eocene deltaic section was derived from
the erosion of emergent highlands northeast of the The Paleocene–Eocene predominantly clastic section
present-day area of Lake Maracaibo and deposited in an or seismic megasequence in the Maracaibo basin has
asymmetric foreland basin formed by southward and been subdivided by Lugo (1991) into seven seismic map
southwestward overthrusting of the Caribbean plate units (Figure 4).
onto this area of the South American passive margin
(Figure 11B). We support this interpretation with isopach Guasare Formation
maps of representative lower Cenozoic seismic map The Guasare Formation (seismic map unit P1) consists
units, subsidence data from wells, and information on of Paleocene limestone and calcareous sandstone (Figure
paleocurrent directions based on geologic studies and 12A). This formation exhibits pronounced lateral facies
seismic lines. variations by grading northwestward into shale,
712 Lugo and Mann

Figure 13—Isopach map for the


entire Eocene interval (seismic
map units E1–E7) of the
Maracaibo basin. Note the
basin asymmetry with thick-
ening toward the east and the
southwest-verging fold and
thrust belt exposed east of the
basin. Highs along the
projected trend of the Mérida
arch are shaded. Thinning east
of the Icotea fault is related to
post-Eocene reactivation and
erosion of Eocene rocks from
the upthrown side of the fault.

northward into shallow marine sedimentary rocks, tolith blocks and rock fragments similar to in situ
westward into coarser grained clastic rocks, and outcrops of the Cretaceous Cogollo Group and La Luna
southward into fluvial and deltaic deposits. On seismic Formation and Paleocene olistolith blocks and rock
lines and in wells, the Guasare Formation appears fragments are common in the Trujillo Formation and
concordant above marine shales of the underlying Colon indicate instability in the upper slope probably caused by
Formation. A hiatus marks the contact between the thrusting in the thrust belt northeast of the Maracaibo
upper Guasare Formation and the overlying Misoa basin (Renz, 1981).
Formation and is probably related to subaerial exposure
marking the end of a regressive cycle begun in Misoa Formation
Campanian time with deposition of shale of the Colon The Misoa Formation (seismic map units E3–E6) is
Formation (Figure 3). characterized by up to 7000 m of deltaic sandstone and
shale found beneath the present-day area of Lake
Trujillo Formation Maracaibo (Zambrano et al., 1971; Van Veen, 1972;
The Trujillo Formation (seismic map units E1 and E2) Barbeito et al., 1985). By middle Eocene time, an
is characterized by turbiditic sandstone and shale found immense fluviodeltaic system represented by the Misoa
in the northeastern part of the Maracaibo basin. East of Formation was present in the lake area. The thickest
Lake Maracaibo in the Mérida Andes, this formation is depocenter of the Misoa Formation occurred in an
about 1800 m thick and composed mainly of turbiditic elongate belt along the northeastern edge of present-day
deposits (Gonzalez de Juana et al., 1980). Cretaceous olis- Lake Maracaibo. In the north-northeastern part of the
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 713

basin, the Misoa Formation conformably overlies the on its flanks. Zambrano et al. (1971) reported no
Paleocene–lower Eocene deposits of the Trujillo Paleocene deposits in the Mérida Andes, Sierra de Perija,
Formation and is conformably overlain by upper middle or Guajira Peninsula.
Eocene shales of the Paují Formation. In the Lake
Maracaibo area and to the west of the lake, the Misoa Trujillo, Misoa, and Paují Formations
Formation unconformably overlies carbonate rocks of Younger seismic units E2–E7 of early–middle Eocene
the Guasare Formation and is unconformably overlain age document a southwest-thinning asymmetric wedge
by Oligocene lacustrine deposits of the Icotea Formation of sediment that migrated southwestward and south-
(Figure 3). To the east of the Icotea fault where no eastward (Figure 12B–F). The thinning of the wedge to
Oligocene deposits occur, the Misoa Formation is overlain the southwest marks the presence of the Mérida arch, a
by transgressive deposits of the La Rosa Formation of feature known from isopach studies to be active
Miocene age. throughout most of the Cretaceous and early Tertiary
A well core of a complete section of the Misoa (Zambrano et al., 1971; Salvador, 1986) (Figure 12).
Formation described by Van Veen (1972) reveals a Seismic lines show that north-northeast trending struc-
variety of facies characteristic of a deltaic complex, tural highs represent Eocene reactivation of underlying
including from base to top, sand-prone distributary north-northeast striking Icotea and Pueblo Viejo high-
channels within a deltaic plain; thin, coarsening-upward angle faults last active during Jurassic rifting (Figure 7).
sandstone, probably representing delta margin sand These complex north-northeast trending structures are
sheets deposited in interdistributary bays during not simple anticlines formed by northwest-southeast
episodic rises of sea level; and cyclic repetitions of marine shortening as suggested by Kellogg and Bonini (1982)
shales (prodelta facies) grading upward abruptly into and Kellogg (1984). Instead, they trend at a high angle to
sand-rich delta front facies. This latter facies exhibits southwest-verging thrusts and anticlines observed in the
abundant cross stratification locally interbedded with subsurface along the northeastern margin of the lake.
trough cross-stratified sets suggestive of deposition in a Their orientation is not controlled by a regional late
channel bar. Tertiary compressional event but rather by the strike of
preexisting and underlying Jurassic rifts (Figure 7).
Pauji Formation
The Pauji Formation (seismic unit E7) is characterized Clastic Sediment Dispersal
by several hundred meters of marine shales found in the
easternmost part of the Maracaibo basin. Sandstone Composition
Kasper and Larue (1986) and Lugo (1991) classified
Isopach Maps the compositions of Eocene sandstones studied by Van
Andel (1958) according to the criteria of Dickinson and
Isopach and isochron maps of seismic units P1, E2, E3, Suczek (1979) and showed a recycled orogen province to
E4, E5, and E7, considered representative of the seven the northwest and west of the lake, a cratonic province to
Paleocene–Eocene seismic map units, are shown in the southeast of the lake, and an intervening transitional
Figures 12A–F. The thickness of all Eocene units (E1–E7) province (Figure 14A, B). Dickinson and Suczek (1979)
are collectively shown in the isopach map of Figure 13. defined the sandstone provinces as follows:
Contacts between units E1–E7 are intraformational
unconformities. • Recycled orogen, for which sand sources are
deformed and uplifted stratal sequences in subduc-
Guasare Formation tion zones, along collision zones, or within foreland
Areas of mainly clastic facies of the Guasare fold and thrust belts.
Formation along the northeastern edge of the basin • Cratonic, for which sand sources are on shields and
exhibit two major depocenters (Figure 12A). Localized platforms or in faulted basement blocks.
clastic Paleocene depocenters contrast with the more • Transitional, for which sand sources are composi-
uniform and thinner sheetlike deposition of Upper tionally transitional between recycled orogen and
Cretaceous carbonate rocks with the underlying passive cratonic sources.
margin section (Figure 8A–D). These depocenters
probably extended more to the northwest prior to their The southeastward shift of the recycled province from
Paleocene–Eocene imbrication in the exposed thrust belt early middle (Figure 14A) to middle Eocene (Figure 14B)
in the Serranía de Trujillo (Stephan, 1985; Mathieu, 1989). time is consistent with the southeastward component of
Folded and thrusted coeval Paleocene sandstone and paleocurrent directions (discussed later), the southeast-
shale crops out in the thrust belt in the Serranía de ward migration of depocenters shown on Paleocene–
Trujillo along with Cretaceous carbonate beds and olis- Eocene isopach maps (Figure 12A–F), and the wedge-like
tolith blocks of the Cogollo Group and La Luna geometry of the Eocene map units (Figure 13). Erosion of
Formation (Renz, 1981). the Mérida arch and underlying Paleozoic Mucuchachi
The more carbonate-rich section in the southern basin Formation may have provided cratonic-type sandstone
exhibits little change in thickness, with 60-m thicknesses along the southern margin of the basin (Figure 7).
above the crest of the Mérida arch and 180-m thicknesses Recycled orogen sources are attributed to the over-
714 Lugo and Mann

Figure 14—(A) Composition of lower middle Eocene sandstone of the Maracaibo basin from Lake Maracaibo using
sandstone petrographic data of van Andel (1958). Letters indicate sampling sites of sandstones similar to “recycled orogen,”
“transitional,” and “craton interior” compositional suites of Dickinson and Suczek (1979). (B) Composition of middle Eocene
sandstone. (C) Early middle Eocene paleocurrents measured by Mathieu (1989) from outcrops of the lower middle Eocene
Misoa Formation northeast of Lake Maracaibo. (D) Arrows show direction of progradation from clinoforms compiled from
various seismic lines. The length of the arrows indicating the clinoforms approximates the length of the seismic section
used to make the interpretation.

thrusting of the arc and oceanic terranes of the Caribbean clinoforms) observed on regional seismic lines as
plate (Figure 1). inclined bedding produced by north to south or
northeast to southwest lateral outbuilding or prograda-
Paleocurrents and Clinoforms tion of Paleocene–Eocene units over the Cretaceous
Mathieu (1989) measured uniform northeast to passive margin. The dispersal pattern for the middle
southwest sole marks on lower–middle Eocene turbidites Eocene (Figure 14C) remains essentially the same as that
of the Trujillo Formation northeast of the lake (Figure for the early–middle Eocene (Figure 14D).
14C). These data are consistent with the idea that the Because the origin of the Eocene clastic wedge in the
Eocene clastic wedge infilled the Maracaibo basin from Maracaibo basin is controversial (Figure 11), we include
the northeast, not the south or southwest (Figure 11). several examples of seismc clinoforms showing apparent
Lugo (1991) interpreted reflection configurations (seismic southeastward and southwestward Eocene prograda-
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 715

Figure 15—Example of clinoforms in seismic map units E1–E6 (Trujillo and Misoa formations) prograding in an approximate
northwest to southeast direction along the northern edge of the Maracaibo foreland basin.

tion. In the Aurare area at the northeastern edge of the vergence directions seen in Figure 16 (to the south-
basin, a northwest-southeast line shows the upper southeast) and Figure 18 (to the west-southwest). Block
Eocene section (seismic map units E1–E6) prograding faults known to be active from isopach maps (Figure 12)
toward the southeast (Figure 15). In another example include the Icotea fault to the west and the Pueblo Viejo
from the eastern margin of the basin, a north-south line fault to the east. These faults were inherited from the
shows seismic map units E3–E6 prograding to the south- Jurassic rift phase (Figure 7) and were subsequently reac-
southeast across an unconformity above map unit E2 tivated during the flexure and compression of the
(Figure 16). A seismic line shown in Figure 17 shows basement of the foreland basin.
clinoforms at different scales within a single line. The
enlargement in the inset of Figure 17 shows smaller scale Basin Subsidence
and more subtle southwest-prograding clinoforms
within unit E5–E6. On a regional scale, units E4–E6 can Lugo (1991) constructed 18 subsidence plots of wells
be seen prograding in a southwestward direction. In a from the Maracaibo basin. Seven of these subsidence
final and more regional example of clinoforms, a 25-km plots, which are corrected for compaction and consid-
line shows units E2–E6 prograding and thinning ered representative of the entire Maracaibo basin, are
westward (Figure 18). Southwest-verging thrust faults shown along with generalized maps of basin depocen-
are interpreted lower in the Cretaceous carbonate section ters in Figure 20. All wells except A show a marked
and are inferred to be part of the leading edge of thrust acceleration in Eocene subsidence followed by an abrupt
detachment beneath this part of the eastern Maracaibo Oligocene rebound. Wells D, E, F, and G along the north-
basin. We therefore correlate these faults to surficial eastern margin of the basin reflect the highest sedimenta-
northwest-striking thrust faults mapped in the area of tion rates of about 40 cm/1000 yr and show a slight
the Serranía de Trujillo by Stephan (1985) and Mathieu younging trend from northwest to southeast. Wells A, B,
(1989). All subsurface thrust faults active during the and C in the Mérida arch area show less pronounced
Eocene are plotted on Figure 19. Note that the strike of subsidence and rebound trends in comparison to the
these thrusts is arcuate to account for the contrast in wells to the northeast. Subsidence patterns are consistent
716 Lugo and Mann

Figure 16—Example of clinoforms in seismic map units E3–E6 (Misoa Formation) prograding across unit E2 (basal Misoa
Formation) in an approximate NNW-SSE direction. Two well logs indicate that unit E4 is mainly shale.

Figure 17—Example of southwest-prograding clinoforms at different scales in units E4–E6 (upper Misoa Formation). Arrows
on inset point out more subtle, smaller scale clinoforms from unit E5. Note sand signature on electric log of units E5–E6.
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 717

Figure 18—Example of WSW progradation of seismic units E2 (Trujillo) and E3–E6 (Misoa). Note the sandy signature of
seismic units E2–E3 and E5–E6 on the electric log. We interpret a west-vergent thrust that detaches near the base of the
Cretaceous carbonate section (Apon Formation, seismic map unit K2). Figure 19 shows map view of all thrusts seen on
seismic lines.

with the interpretation of isochron, isopach, and clastic Maracaibo basin. As the thrust belt moved south, the
dispersal data of an asymmetric basin infilling from the foreland basin enlarged in a south-southeast direction.
north and northeast and wedging out to the southwest. The shape of the foreland basin at any time was trian-
gular because its limits were defined by the craton to the
Proposed Foreland Basin Model south, the sutured area to the west, and the closing
basinal area to the east. Such a triangular pattern is
We propose a coupled thrust belt and foreland basin supported by the shape of the contoured sandstone
model in a zone of oblique convergence to explain the composition data (Figure 14A, B). Flexural formation of
previously described characteristics of Paleocene–Eocene the foreland basin may have become impeded by the
sedimentation in the Maracaibo basin. The proposed presence of the northwest-trending Mérida arch, a
model is illustrated in map view in Figure 21 and in tectonic arch known from isopach, well, and outcrop
cross-sectional view in Figure 22. studies to have been active throughout most of Creta-
ceous time (Figures 8, 9).
Map View The orientation of the northwest-trending fold and
In the early–late Paleocene (Figure 21A), a thrust belt thrust belt along the eastern edge of the Maracaibo basin
at the leading edge of the Caribbean plate moved south relative to a generally eastward moving Caribbean plate
and a Paleocene foredeep developed north of the is problematic. Stephan (1985) and Mathieu (1989)
718 Lugo and Mann

An alternative explanation for the change in the orien-


tation of the thrust front and associated foreland basin is
right-lateral offset along the Boconó fault zone and
internal deformation related to the northward escape of
the Maracaibo block (Mann and Burke, 1984). Stephan
(1985) measured an 80-km apparent right-lateral offset of
the thrust front in the Mérida Andes which he postulated
occurred over the past 5 Ma along the presently active
right-lateral Boconó fault zone. In this model, the fold
and thrust belt to the east of the Maracaibo basin may
have originally had a more east-west orientation that
aligned with the thrust front east of the Boconó fault
(Figure 1A). Offset along the Boconó fault has pushed the
western segment of the thrust belt to the north along
with the rest of the Maracaibo block.

Cross-Sectional View
The sequential evolution of a northeast-southwest
section through the eastern Maracaibo basin is shown for
six stages in Figure 22. The bar along the base of each
diagram schematically shows how the locus of subsi-
dence changed along the length of the profile through
time.
In stage 1 in the middle–late Paleocene, the basin was
a northwest-prograding passive margin. During stage 2
in the late Paleocene, the first thrust faults imbricated a
thrust belt north of the basin. Imbrication produced
slumping of Cretaceous carbonate olistoliths in
Paleocene sandstone and shale matrix of the foreland
basin (Renz, 1981). Early Paleocene foredeep deposits
themselves became imbricated and included as part of
Figure 19—Map showing faults active during the early the thrust belt exposed northeast of Lake Maracaibo
Eocene in the Maracaibo basin based on seismic grid
shown. Southwest-verging thrust faults are interpreted as
(Stephan, 1985; Mathieu, 1989). On the cratonic margin
the upward propagation of thrust detachment at the base south of the foreland basin, fluvial basal sands of seismic
of the Cretaceous carbonate section (see Figure 18). Icotea units E1 and E2 were deposited. By stage 3 in early
fault (to west) and Pueblo Viejo faults are inherited Eocene time, turbiditic sedimentation was occurring in
Jurassic rift faults (see Figure 7) that were reactivated as the foredeep. Fluvial deposition diminished on the
high-angle block faults during this thrusting phase. cratonic southern margin of the foredeep as deltaic and
shoreface sedimentation (seismic units E2 and E3)
heralded higher rates of subsidence along the cratonic
proposed a right-lateral strike-slip offset of a southeast- margin. By stage 4 in early–middle Eocene time, the
verging thrust front in the region of this anomalously thrust belt had impinged on the cratonic margin. Higher
oriented thrust front. This thrust front offset would be subsidence rates along the cratonic margin caused
equivalent to a lateral ramp or “transversale” structure in retrogradation of seismic unit E3 and exclusively
the regional structure of the fold and thrust belt. The shoreline shelf deposition. By stage 5 in the middle
parallelism of the Mérida arch to the belt suggests that Eocene, downlapping clastic sequences of the foredeep
preexisting passive margin salients and reentrants in the prograded across shallower facies of the cratonic margin.
margin may have guided the direction of collision- High rates of subsidence caused transgression of shales
related thrusting and produced the lateral ramp in shallow marine environments. In stage 6 in late
structure. middle Eocene time, subsidence rates diminished as the
We have adapted the strike-slip interpretation in Mérida arch resisted flexure and limited the space
Figure 21C, which shows right-lateral strike-slip faults in available to receive further foredeep sedimentation.
the fold and thrust belt area. The eastward motion of the Seismic unit E6 prograded rapidly cratonward and over-
Caribbean plate and the presence of the Mérida arch may filled the basin. At depth beneath the foreland basin,
have interacted to deflect the foredeep axis from a north- thrusts propagated southward along detachments near
westward trend to a more eastward trend. Continued the base of the Cretaceous passive margin carbonate
eastward motion of the Caribbean plate closed the area section. During this phase the entire thrust belt became
between the Mérida arch and the thrust belt and formed emergent and resulted in the erosion and deposition of
a locus of subsidence farther to the east by early–middle seismic units E6 and E7. Uplift and exposure of the basin
Eocene time (Figure 21C). itself truncated seismic units E5–E7.
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 719

Figure 20—Subsidence history of the Maracaibo


basin based on seven decompacted subsidence
plots. (A) Map showing well locations (circled
letters) and inferred migration of foreland basin
depocenter from Paleocene to late middle Eocene.
This map is based on isopach maps in Figure
12A–F. (B) Subsidence history curves exhibit rapid
subsidence followed by rebound and a decrease in
subsidence toward the craton.
720 Lugo and Mann

DISCUSSION
This study supports the basic interpretation of Eva et
al. (1989) that the Jurassic–Eocene geologic history of
northern South America can be generalized into three
distinct tectonic phases:

• Jurassic rifting accompanied separation of the


North and South American continents.
• Early–Late Cretaceous passive margin subsidence
followed rifting.
• Paleocene–Eocene foreland basin subsidence
followed the oblique collision of a Pacific-derived
Caribbean plate with the South American passive
margin.

This paper has reviewed subsurface data from the


Maracaibo basin related to the first two events which are
fairly well understood on the basis of previous outcrop
and well studies. The distribution of Jurassic red beds
and volcanic rocks in deep wells beneath the Maracaibo
basin suggests that these rocks were deposited in north-
northeast trending rift basins separated by horst blocks
of Paleozoic metasedimentary rocks (Figure 7). This
north-northeast structural grain was subsequently reacti-
vated during Cenozoic collisional and strike-slip defor-
mation events. Lithology and isopach patterns of
Barremian–Santonian mainly shallow marine carbonate
rocks in the Maracaibo basin suggests that these rocks
formed in a passive margin setting above Precambrian
and Paleozoic crust thinned during Jurassic rifting event
(Figure 8). Isopach patterns, outcrop studies, and compi-
lations of well data show that the northwest-trending
Mérida arch was a positive feature from Aptian to
Campanian time (Renz, 1959; Macellari, 1988) (Figure 9).
The tectonic origin of the Mérida arch and similar arches
in Colombia and Ecuador (Figure 1B) is unclear, but is
perhaps related to reactivation of underlying Precam-
brian or Paleozoic structural grains during passive
margin subsidence.
Nearly two-thirds of the total stratigraphic section of
the Maracaibo basin is composed of Eocene sandstone
and shale deposited in fluvial and marine environments.
Previous studies have concluded that these rocks were
derived from the erosion of highlands southwest of Figure 21—Map view model for early Paleocene–early
present-day Lake Maracaibo and formed a southwest to middle Eocene foreland basin sedimentation in the
northeast gradation from fluvial and deltaic deposits to Maracaibo basin based on seismic, well, and geologic data
deep marine submarine fans (Figure 11A). Based on presented in this paper. See text for discussion.
isochron and isopach maps (Figure 12), a variety of sedi-
mentary provenance data (Figure 14), and well subsi-
dence data (Figure 20), we propose that deposition terminal convergent deformation in the lake area is
occurred in a foreland basin setting and that the main constrained as middle Eocene because an angular uncon-
source area for the Paleocene–Eocene clastic rocks was a formity separates folded lower Eocene seismic units
contemporaneous southwest-verging thrust belt that lay E1–E6 from the overlying, tilted, but unfolded upper
to the northeast of the basin (Figure 11B). Eocene seismic unit E7. Thickening trends of units P1
The foreland basin interpretation is consistent with and E1–E7 suggest southeastward migration in the
geologic mapping of southwest-verging thrust faults in thrusting from Paleocene to middle Eocene time (e.g.,
the subsurface (Figure 19) and in surface outcrops Figures 12, 20).
northeast of Lake Maracaibo (Stephan, 1985; Mathieu, Three features observed in the Maracaibo basin are
1989), as well as with northeast-southwest and north- common to better studied foreland basins worldwide.
south paleocurrent data (Figure 14). The age of the First, directions of paleocurrents are basin parallel
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 721

Figure 22—Cross-sectional model for early Paleocene–middle Eocene foreland basin sedimentation in the
Maracaibo basin based on seismic, well, and geologic data presented in this paper. See text for discussion.
722 Lugo and Mann

Figure 23—(A) Location map


and extent of foreland basin
rocks (shaded) and
Jurassic–Cenozoic passive
margins (crosses) surrounding
the Caribbean plate (modified
from Pindell, 1991). Numbered
lines show positions of leading
edge of the Caribbean plate
based on ages of marine
foreland basin deposits: 1, Late
Cretaceous; 2, Paleocene; 3,
Eocene; 4, Oligocene; 5,
Miocene; 6, late Miocene–
Recent. (B) Map of surface and
subsurface exposures of
foreland basin rocks (shaded) in
northern South America.
Numbers indicate ages of
foreland basin deposits as in
part (A). Box indicates Lake
Maracaibo study area shown in
Figure 2.

(Figure 14C, D). Second, a rebound event followed the Along the northern margin of the Caribbean, synoro-
period of maximum Eocene subsidence and was probably genic sedimentation and north-verging folds are late
related to relaxation of plate-bending stresses (Figure 20). Campanian age in Guatemala (Rosenfeld, 1990),
Third, an average subsidence rate of 216 mm/yr in the Paleocene in western Cuba (Bralower et al., 1993), Eocene
Paleocene–Eocene compares well with an average rate of in central Cuba (Pszczolkowski and Flores, 1986), middle
200 mm/yr for foreland basins worldwide. Eocene in Hispaniola, and early Oligocene in Puerto Rico
The Maracaibo basin makes up the western segment (Dolan et al., 1991) (Figure 23). Assuming that the
of a 1500-km-long south-verging thrust belt and flanking Caribbean plate formed a single, rigid entity, it is
foreland basin along northern South America (Figure 23). possible to reconstruct the position of the plate relative to
The age of deformation and foreland basin subsidence the margins of North and South America by connecting
becomes younger eastward from the Maracaibo basin northern and southern areas of coeval folding and
(Pindell, 1991). Figueroa de Sanchez and Hernandez synorogenic sedimentation. It is interesting to note that if
(1990) document late Eocene–Oligocene north to south northern and southern Caribbean foreland subsidence
overthrusting at the Guarumen-1S exploration well 200 was simultaneous during Paleocene–Oligocene time, an
km to the east of Lake Maracaibo (number 4 on Figure outward or radial component of overthrusting is needed
23), while Beck et al. (1990) propose middle–late Eocene to accompany the overall eastward migration of the
thrusting from surface mapping in the same area. We Caribbean plate. A similar style of radial thrusting has
prefer the Oligocene age of thrusting in the area because been observed by Platt et al. (1989) in the western Alps
this younger age allows a more gradual progression and was interpreted by them as the combination of
from Paleocene–Eocene deformation in the Maracaibo forces produced by plate motion and gravitational
basin to middle Miocene deformation in eastern sliding of the mountain belt in the direction of greatest
Venezuela (number 5 in Figure 23) (Erlich and Barrett, surface slope. More detailed studies of circum-Caribbean
1992) and late Miocene–Recent deformation in Trinidad foreland basins and thrust belts are needed to better
and offshore areas (number 6 in Figure 23) (Salvador and constrain the plate reconstruction in the inset of Figure 23
Stainforth, 1965; Speed, 1985; Pindell, 1991; Algar and and the tectonic process of radial thrusting that it implies.
Pindell, 1993). Hydrocarbon occurrences in the Maracaibo basin are
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 723

concentrated along the eastern edge of the basin near the Benedetto, G., and R. O. Odreman, 1977, Nuevas evidencias
thrust front separating the Caribbean plate from the paleontológicas en la Formación La Quinta, su edad y
South American craton (Zambrano et al., 1971; Bock- correlación con las unidades aflorantes en la Sierra de
meulen et al., 1983) (Figure 23). Overthrusting of the Perijá y Cordillera Oriental de Colombia: 5th Venezuelan
Geological Congress, Caracas, Venezuela, p. 87–106.
Caribbean plate onto source rocks such as the Ceno-
Blaser, R., and C. White, 1984, Source rock and carbonization
manian–Santonian La Luna Formation led to their matu- study, Maracaibo basin, Venezuela: AAPG Memoir 35,
ration and migration in Eocene time (James, 1990; p. 229–254.
Pindell, 1991). Pindell (1991) has noted that most circum- Bockmeulen, H., C. Barker, and P. A. Dickey, 1983, Geology
Caribbean hydrocarbon accumulations can be explained and geochemistry of crude oils, Bolivar coastal fields,
by overthrusting of the Caribbean plate onto source Venezuela: AAPG Bulletin, v. 67, p. 242–270.
rocks generated during the Cretaceous passive margin Boesi, T., and D. Goddard, 1989, A new geologic model
phase of Caribbean evolution. The results of this study of related to the distribution of hydrocarbon source rocks in
the Maracaibo basin support this general model and the Falcón basin, northwestern Venezuela, in K. T. Biddle,
suggest that productive trends could be present along ed., Active margin basins: AAPG Memoir 52, p. 35–49.
Bot, P., and J. L. Perdomo, 1986, Evolución tectónica en la
this foreland front in the area of the Guajira Peninsula
cuenca de Maracaibo: Transactions of the 3rd Venezuelan
and in the area east of the Mérida Andes (Figure 1A). Geophysical Congress, p. 484–493.
Bowland, C. L., and E. Rosencrantz, 1988, Upper crustal
structure of the western Colombian basin, Caribbean Sea:
Acknowledgments We thank Lagoven and Maraven for GSA Bulletin, v. 100, p. 534–546.
providing seismic and well data used in this study. Lagoven Bralower, T. J., F. Hutson, P. Mann, M. Iturralde-Vinent, P.
supported the first author as a Ph.D. student in geology for Renne, and W. V. Sliter, 1993, Tectonics of oblique arc-
three and one-half years at the Department of Geological continent collision in western Cuba, 1: stratigraphic
constraints: EOS, Transactions of the American Geophys-
Sciences of the University of Texas at Austin. A. Salvador
ical Union, v. 74, p. 546.
shared his 50 years of experience in Venezuelan geology with us Burke, K., 1988, Tectonic evolution of the Caribbean: Annual
at all stages of this project. Special thanks to W. Galloway, W. Review of Earth and Planetary Sciences, v. 16, p. 210–230.
Muehlberger, L. Brown, I. Azpiritxaga, J. Maguregui, E. Byers, C. W., 1977, Biofacies patterns in euxinic basins: a
Quijada, L. Rodriguez, and G. Marton for valuable discussions. general model, in H. Cook and P. Enos, eds., Deep-water
We appreciate the efforts of Richardson Allen, H. Mullins, and carbonate environments: SEPM Special Publication 25,
an anonymous reviewer for their critical comments of an earlier p. 5–18.
version of this paper and the efforts of A. Tankard, D. Larue, Chigne, N., and L. Hernández, 1990, Main aspects of
and an anonymous reviewer for their reviews of the version petroleum exploration in the Apure area of southwestern
submitted to this volume. UTIG contribution 1040. Venezuela, 1985–1987, in J. Brooks, ed., Classic petroleum
provinces: Geological Society of London Special Publica-
tion, v. 50, p. 55–75.
Dasch, L. E., and P. Banks, 1981, Zircon U-Pb ages from the
REFERENCES CITED Sierra de Perijá, Venezuela (abs.): GSA Abstracts with
Programs, v. 13, p. 436.
Algar, S. T., and J. L. Pindell, 1993, Structure and deformation Dickinson, W. R., and C. Suczek, 1979, Plate tectonics and
history of the Northern Range of Trinidad and adjacent sandstone compositions: AAPG Bulletin, v. 63,
areas: Tectonics, v. 12, p. 814–829. p. 2164–2182.
Azpiritxaga, I., 1991, Carbonate depositional styles controlled Dolan, J., P. Mann, R. de Zoeten, C. Heubeck, J. Shiroma, and
by siliciclastic influx and relative sea-level changes, lower S. Monechi, 1991, Sedimentologic, stratigraphic, and
Cretaceous, central Lake Maracaibo, Venezuela: Master’s tectonic synthesis of Eocene–Miocene sedimentary basins,
thesis, University of Texas at Austin, Austin, Texas, 145 p. Hispaniola and Puerto Rico, in P. Mann, G. Draper, and J.
Barbeito, P. J., R. Pettelli, and A. M. Evans, 1985, Estudio F. Lewis, eds., Geologic and tectonic development of the
estratigáfico del Eoceno en el área de Maracaibo, North America–Caribbean plate boundary in Hispaniola:
Venezuela occidental, basado en interpretaciones paleon- GSA Special Paper 262, p. 217–263.
tológicas y palinológicas: 6th Venezuelan Geological Erlich, R. N., and S. F. Barrett, 1992, Petroleum geology of the
Congress (Caracas, Venezuela), p. 109–139. eastern Venezuela foreland basin, in R. W. Macqueen and
Bartok, P., 1993, Pre-breakup geology of the Gulf of D. A. Leckie, eds., Foreland basins and fold belts: AAPG
Mexico–Caribbean: its relation to Triassic and Jurassic rift Memoir 55, p. 341–362.
systems of the region: Tectonics, v. 12, p. 441–459. Eva, A. N., K. Burke, P. Mann, and G. Wadge, 1989, Four-
Bartok, P., T. J. A. Reijers, and I. Juhasz, 1981, Lower Creta- phase tectonostratigraphic development of the southern
ceous Cogollo Group, Maracaibo basin, Venezuela: Sedi- Caribbean: Marine and Petroleum Geology, v. 6, p. 9–21.
mentology, diagenesis, and petrophysics: AAPG Bulletin, Feo-Codecido, G., F. D. Smith, Jr., N. Aboud, and E. Di
v. 5, p. 1110–1134. Giacomo, 1984, Basement and Paleozoic rocks of the
Beck, C., Y. Ogawa, Y., and J. Dolan, 1990, Eocene paleogeog- Venezuelan Llanos basins, in W. E. Bonini, R. B.
raphy of the southeastern Caribbean: relations between Hargraves, and R. Shagam, eds., The Caribbean–South
sedimentation on the Atlantic abyssal plain at Site 672 and American plate boundary and regional tectonics, GSA
evolution of the South American margin, in J. C. Moore, A. Memoir 162, p. 175–187.
Mascle, et al., eds., Proceedings of the Ocean Drilling Figueroa de Sanchez, L., and L. Hernández, 1990, Exploración
Project, Scientific Results Leg 110, Ocean Drilling Program, geofísica-geológica del área de Guarumen: Transactions,
p. 129–140. 5th Venezuelan Geophysical Congress, p. 219–227.
724 Lugo and Mann

Ghosh, S., 1984, Late Cretaceous condensed sequence, 7th Venezuelan Geological Congress, Caracas, Venezuela,
Venezuelan Andes, in W. E. Bonini, R. B. Hargraves, and p. 852–866.
R. Shagam, eds., The Caribbean–South American plate Miller, J. B., K. L. Edwards, P. P. Wolcott, H. W. Anisgard, R.
boundary and regional tectonics: GSA Memoir 162, Martin, and H. Anderegg, 1958, Habitat of oil in the
p. 317–324. Maracaibo basin, Venezuela: AAPG Bulletin, v. 42,
Gonzalez de Juana, C., J. Iturralde, and X. Picard, 1980, p. 601–640.
Geología de Venezuela y sus Cuencas Petrolíferas: Montgomery, H., E. A. Pessagno, Jr., and I. M. Muñoz, 1992,
Caracas, Foninves, 1031 p. Jurassic (Tithonian) radiolaria from La Désirade (Lesser
Gordon, M. B., 1990, The Chortis block is a continental, pre- Antilles): Preliminary paleontological and tectonic implica-
Mesozoic terrane, in D. K. Larue and G. Draper, eds., tions: Tectonics, v. 11, p. 1426–1432.
Transactions of the 12th Caribbean Geological Conference, Moticska, P., 1975, Sierra de Perijá, Excursión #2: Complejo
St. Croix, U.S. Virgin Islands: Miami Geological Society, volcánico-plutónico de El Totumo-Inciarte: Boletín de
Miami, Florida, p. 505–512. Geología, Special Publication, no. 7, p. 306–311.
Hancock, J. M., and E. G. Kaufman, 1979, The great transgres- Odreman, R. O., and G. Benedetto, 1977, Paleontología y edad
sion of Late Cretaceous: Journal of Geological Society of de la formación Tinacoa, Sierra de Perijá, Estado Zulia,
London, v. 136, p. 175–186. Venezuela: 5th Venezuelan Geological Congress, Caracas,
Haq, B. U., J. Hardenbol, and P. R. Vail, 1987, Chronology of Venezuela, p. 15–32.
fluctuating sea level since the Triassic: Science, v. 235, Pennington, W. D., 1981, Subduction of the eastern Panama
p. 1156–1167. basin and seismotectonics of northwestern South America:
James, K. H., 1990, The Venezuelan hydrocarbon habitat, in J. Journal of Geophysical Research, v. 86, p. 10,753–10,770.
Brooks, ed., Classic petroleum provinces: Geological Pindell, J. L., 1985, Alleghenian reconstruction and the subse-
Society of London Special Publication, v. 50, p. 9–35. quent evolution of the Gulf of Mexico, Bahamas, and
Kasper, D. C., and D. K. Larue, 1986, Paleogeographic and proto-Caribbean Sea: Tectonics, v. 4, p. 1–39.
tectonic implications of quartzose sandstones of Barbados: Pindell, J. L., 1991, Geological rationale for hydrocarbon
Tectonics, v. 5, p. 837–854. exploration in the Caribbean and adjacent regions: Journal
Kellogg, J. N., 1984, Cenozoic tectonic history of the Sierra de of Petroleum Geology, v. 14, p. 237–257.
Perijá, Venezuela-Colombia and adjacent basins, in W. E. Pindell, J. L., and S. F. Barrett, 1990, Geologic evolution of the
Bonini, R. B. Hargraves, and R. Shagam, eds., The Caribbean: a plate-tectonic perspective, in G. Dengo and J.
Caribbean–South American plate boundary and regional E. Case, eds., The Caribbean, Volume H: Decade of North
tectonics: GSA Memoir 162, p. 239–261. American Geology, GSA, p. 405–432.
Kellogg, J. N, and W. E. Bonini, 1982, Subduction of the Platt, J., J. Behrmann, P. Cunningham, J. Dewey, M. Helman,
Caribbean plate basement uplifts in the overriding South M. Parish, M. Shepley, S. Wallis, and P. Weston, 1989,
American plate: Tectonics, v. 1, p. 251–276. Kinematics of the Alpine arc and the motion history of
Loaiza, P., L. Hernández, C. Urbina, and R. Coriat, 1990, Reac- Adria: Nature, v. 337, p. 158–161.
tivación de la exploración en Barinas: 5th Venezuelan Pszczolkowski, A., and R. Flores, 1986, Fases tectónicas del
Geological Congress, Caracas, Venezuela, p. 236–243. Cretácico y del Paleogene en Cuba occidental y central:
Lugo, J. M., 1991, Cretaceous to Neogene tectonic control on Bulletin of the Polish Academy of Sciences (Earth
sedimentation, Maracaibo basin, Venezuela: Ph.D. Disser- Sciences): v. 34, p. 95–111.
tation, University of Texas at Austin, Austin, Texas, 219 p. Renz, O., 1959, Estratigrafía del Cretácico en Venezuela:
Macellari, C. E., 1988, Cretaceous paleogeography and deposi- Boletín de Geología, Caracas, Venezuela, v. 5, p. 3–48.
tional cycles of western South America: Journal of South Renz, O., 1960, Geología de la parte sureste de la Península de
American Earth Sciences, v. 1, p. 373–418. la Guajira (República de Venezuela): 3rd Venezuelan
Macellari, C. E., and DeVries, T. J., 1987, Late Cretaceous Geological Congress Special Publication, no. 3, p. 317–374.
upwelling and anoxic sedimentation in northwestern Renz, O., 1981, Venezuela, in R. A. Reyment and P. Bengston,
South America: Palaeogeography, Palaeoclimatology, eds., Aspects of mid-Cretaceous regional geology: New
Palaeoecology, v. 59, p. 2790–292. York, Academic Press, p. 197–220.
Mann, P., and K. Burke, 1984, Neotectonics of the Caribbean: Restrepo, J. J., and J. F. Toussaint, 1988, Terranes and conti-
Reviews of Geophysics and Space Physics, v. 22, nental accretion in the Colombian Andes: Episodes, v. 11,
p. 309–362. p. 189–193.
Mann, P., G. Draper, and J. F. Lewis, 1991, An overview of the Rosenfeld, J. H., 1990, Sedimentary rocks of the Santa Cruz
geologic and tectonic development of Hispaniola, in P. ophiolite—a proto-Caribbean history: Transactions, 12th
Mann, G. Draper, and J. F. Lewis, eds., Geologic and Caribbean Geological Conference, U.S. Virgin Islands,
tectonic development of the North America–Caribbean p. 513–519.
plate boundary in Hispaniola: GSA Special Paper 262, Salvador, A., and R. M. Stainforth, 1965, Clues in Venezuela to
p. 1–28. the geology of Trinidad, and vice versa: Transactions, 4th
Mathieu, X., 1989, La Serrania de Trujillo-Ziruma aux confins Caribbean Geological Conference, Trinidad, p. 31–40.
du bassin de Maracaibo, de la Sierra du Falcon et de la Salvador, A., 1986, Comment on “Neogene block tectonics of
Chaine Caribe (Venezuela): Ph.D. Dissertation, Université eastern Turkey and northern South America: continental
de Bretagne Occidentale, Brest, 264 p. applications of the finite difference method” by J. F. Dewey
Maze, W. B., 1984, Jurassic La Quinta Formation in the Sierra and J. L. Pindell: Tectonics, v. 5, p. 697–701.
de Perijá, northwestern Venezuela: Geology and tectonic Schubert, C., R. S. Sifontes, V. E. Padrón, J. R. Vélez, and P. A.
environment of red beds and volcanic rocks, in W. E. Loaiza, 1979, Formación La Quinta (Jurásico) Andes
Bonini, R. B. Hargraves, and R. Shagam, eds., The Merideños: Geología de la sección tipo: Acta Científicas
Caribbean–South American plate boundary and regional Venezolana, v. 30, p. 42–55.
tectonics: GSA Memoir 162, p. 263–282. Shagam, R., 1985, Evolución geologíca tectónica del bloque de
Mendez, J., 1990, La Formación La Luna: características de Maracaibo-Santa Marta: 4th Venezuelan Geological
una cuenca anóxica en una plataforma de aguas someras: Congress, Caracas, Venezuela, p. 6521–6549.
Jurassic–Eocene Tectonic Evolution of Maracaibo Basin, Venezuela 725

Speed, R. C., 1985, Cenozoic collision of the Lesser Antilles arc Winter, Th., and A. Lavenu, 1989, Morphological and micro-
and continental South America and the origin of the El tectonic evidence for a major active right-lateral strike-slip
Pilar fault: Tectonics, v. 4, p. 41–69. fault across central Ecuador (South America): Annales
Stephan, J. F., 1985, Andes et Chaine Carïbe sur la transversale Tectonicae, v. 3, p. 123–139.
de Barquisimeto (Venezuela): evolution géodynamique, in Zambrano, E., E. Vásquez, B. Duval, M. Latrielle, and B.
A. Mascle, ed., Geodynamique des Carïbes: Paris, Editions Coffinieres, 1971, Síntesis paleogeográfica y petrolera del
Technip, p. 505–529. occidente de Venezuela: IV Congreso Geológico Vene-
Trump, G. W., and Salvador, A., 1964, Guidebook to the zolano, Memoir 1, p. 481–552.
geology of western Táchira: Venezuelan Association of
Geology, Mining, and Petroleum, Caracas, Venezuela, Authors’ Mailing Addresses
25 p.
van Andel, T. H., 1958, Origin and classification of Creta- Jairo Lugo
ceous, Paleocene, and Eocene sandstones of western Departamento de Geología
Venezuela: AAPG Bulletin, v. 42, p. 734–763. Lagoven, S. A.
van der Hilst, R. D., and P. Mann, 1994, Tectonic implications Apartado 889
of tomographic images of subducted lithosphere beneath Caracas 1010-A
northwestern South America: Geology, v. 22, p. 451–454. Venezuela
Van Veen, F. R., 1972, Ambientes sedimentarios de las forma-
ciones Mirador y Misoa del Eoceno inferior y medio de la
cuenca del Lago de Maracaibo: 4th Venezuelan Geological
Paul Mann
Congress, Caracas, Venezuela, p. 1074–1114. Institute for Geophysics
Winker, C. D., and R. T. Buffler, 1988, Paleogeographic University of Texas at Austin
evolution of early deep-water Gulf of Mexico and margins, 8701 Mopac Blvd.
Jurassic to middle Cretaceous (Comanchean): AAPG Austin, Texas 78759-8345
Bulletin, v. 72, p. 318–346. U.S.A.
View publication stats

También podría gustarte