Sunteți pe pagina 1din 14

Geological Society of America

Memoir 162
1984

Structure and Cenozoic tectonics of the


Falcon Basin, Venezuela, and adjacent areas
Karl W. Muessig
Exxon Minerals Company
P.O. Box 2189
Houston, Texas 77001

ABSTRACT
The Falcon Basin in northwestern Venezuela and adjacent offshore basins devel-
oped within a zone of extensional tectonics during Oligocene and Miocene times.
Extension resulted fromright-lateralmotion along offset, east-west-trending, transcur-
rent faults, including the Oca fault in western Venezuela, the Cuiza fault in northern
Colombia, and the San Sebastin fault along the coastal areas of central Venezuela. On
both local and regional scales, transcurrent and normal faults were active during the
early evolution of the basins. These faults define rhomb-shaped pullapart basins in map
plan. Extension occurred in a northeast direction causing normal faulting along north-
west trends. Basin subsidence was accompanied by crustal thinning and injection of
basaltic magmas.
Evidence of Oligocene magmatic activity is found in the central part of the Falcon
Basin where volcanic rocks and hypabyssal intrusions are exposed. These rocks are
similar to other suites of continental igneous rocks typically associated with rifting
environments. Basaltic rocks of both alkalic and subalkalic affinities are present. Xeno-
liths of the underlying crust and mantle are abundant. Felsic igneous rocks are relatively
rare.
Other structures within the Falcon and Bonaire basins formed during later stages of
basin development. These include folds and reverse faults of northeast trend and conju-
gate sets of small transcurrent faults. These structures were amplified by greater com-
pressional stresses during late Miocene and Pliocene time.
A similar Tertiary tectonic regime is postulated for the larger area of the Bonaire
Crustal Block, a block that includes the Falcon and Bonaire basins. This block is a broad
region of extensional pullapart structures which developed during the Oligocene to
Miocene, right-lateral motion between the Caribbean and South American crustal
plates. Minimum extension within the northern part of the block, along the Venezuelan
and Netherlands Antilles and the Paraguan and Guajira peninsulas, is estimated at 35 to
45 km in an east-west direction. The extension within this nonrigid block should be
considered when determining Tertiary movements between the Caribbean and South
American plates.

217
218 K. W. Muessig
INTRODUCTION
The Cenozoic tectonic boundary between the Caribbean
and South American areas is frequently interpreted as a right-
lateral transform boundary through or adjacent to the Falcon
Basin in northwestern Venezuela (Stainforth, 1969; Molnar and
Sykes, 1969; Malfait and Dinkelman, 1972; Silver and others,
1975). However, the origin of the basin has never been consid-
ered in detail. A transform plate boundary through the Falcon
region might be expressed as a series of right-lateral transcurrent
faults along a line connecting the Oca fault to the San Sebastin
fault (Fig. 1). However, the only documented evidence suggesting
a buried structure is the presence of en echelon folds and block
uplifts attributed to right-lateral wrenching (Vsquez and Dickey,
1972). Elsewhere in northern South America, east-west directed,
right-lateral, transcurrent faulting apparently began about mid-
Eocene (Bell, 1972). This initiation coincides with the cessation Figure 1. Location map of the Falcn Basin, Venezuela, with paleogeo-
of submarine thrusting and flysch deposition in the Siquisique- graphic features and surrounding structures. Broken line ( ) indi-
cates the approximate extent of the basin (Wheeler, 1963; Ferrell and
Barquisimeto area south of the Falcon Basin (Stephan, 1977). others, 1969). Stratigraphic cross section along lines N-S and W-E are
Thus it is possible that the Falcon Basin developed during a shown in Figure 2. Ruled area shows central Falcn Basin study area of
period of transcurrent faulting that postdates compressional de- Figure 3. Bouguer gravity anomaly contours are included from Bonini
formation and thrusting of Eocene and earlier times. Indeed, the and others (1977).
basin may have developed along transcurrent faults.
Case (1974) suggested that many Neogene basins along the Cretaceous black shales, and an underlying granitic basement is
northern South American borderland, including the Falcon suspected (unpublished report of the Texas Petroleum Company;
Basin, may have originated by north- to northwest-directed ex- Ferrell and others, 1969). To the north, the Paraguan high re-
tension. Silver and others (1975) briefly examined the basin in stricted access to the Falcn Basin to the Caribbean Sea. This
relation to its offshore continuation, the Bonaire Basin. They high is a basement massif of igneous and metamorphic rocks
concluded that both developed in a zone of Oligocene to Mio- (MacDonald, 1968; Martin-Bellizzia and Arozena, 1972) ranging
cene rifting between the Venezuelan mainland and the Venezue- in age from Paleozoic through Paleocene (Santamara and Schu-
lan and Netherlands Antilles. bert, 1974). The narrow Urumaco Trough connected the Falcn
Detailed studies of the Falcn Basin geology were under- Basin with the Gulf of Venezuela and separated the Dabajuro and
taken to clarify the history of the basin and to constrain the Paraguan highs. The eastern Falcn Basin opened into the Bo-
Tertiary tectonic evolution of the southern Caribbean borderland. naire Basin (Fig. 1).
This paper summarizes the stratigraphic, structural, geophysical, The earliest Falcn Basin deposits are considered to be
and petrologic data obtained. Based on these data a tectonic upper Eocene (Wheeler, 1963). Deposition continued intermit-
model for the basin is postulated, which reconciles the observed tently through Oligocene and Miocene time. Sedimentary thick-
regional extensional regime with the right-lateral, transcurrent nesses exceeding 9,000 m accumulated and are still preserved in
faulting. The model is extended to adjacent basins, and its impli- the northern and eastern parts of the basin (unpublished report of
cations for Caribbean-South American plate movements are the Texas Petroleum Company; Ferrell and others, 1969). In the
assessed. central basin area, deposition ceased after early Miocene time.
During Oligocene to lower Miocene time, approximately 3.5 km
FALCON BASIN STRATIGRAPHY of sediments were deposited in the central basin at rapid subsi-
dence rates of 500 to 700 m/m.y. (Diaz de Gamero, 1977).
During Oligocene to Miocene time a large part of north- The Oligocene and Miocene stratigraphy is complicated by
western Venezuela was occupied by the marine Falcn Basin many lateral and vertical facies changes (Fig. 2). The stratigraphic
(Fig. 1). It was elongate in an east-west direction and bordered by section is divided into the lower Oligocene El Paraso deltaic
topographic highs to the south, west, and north. The older Paleo- sands and shales, which by middle Oligocene grade upward into
cene to Eoceneflyschbasin, which contains allochthonous ophio- the Pecaya marine shales. Igneous intrusives and extrusives are
lites in the Siquisique region (Bellizzia and others, 1972), was found within the upper El Paraso and lower Pecaya formations
deformed, uplifted and was being eroded at this time. It formed (Fig. 3). The El Paraso Formation is a clastic sequence 500 to
the southern high. To the west, the basin shallowed onto the 1,000 m thick, composed of fine-grained, quartz sandstones and
Maracaibo Platform. The Dabajuro high bordered the basin to siltstones intercalated with mudstones, shales, and thin lignites.
the northwest. Exploration wells on this high bottom in Upper The sandstones are well indurated with siliceous cement and are
Structure and Cenozoic tectonics of the Falcon Basin 219
NORTH SOUTH To the west, shallow-water clastics of the Castillo Forma-
tion replace the Pedregoso and part of the Pecaya formations
GUARABAL
SANn
AGUA CLARA (Wheeler, 1963). To the south, another complicated carbonate
PATIECITOS>
2 LUIS j r
bank is represented by the Churuguara Formation. South of Chu-
ruguara the Oligocene to Miocene strata thin rapidly. The Chu-
ruguara Formation gives way southward to basin margin
conglomerates, sandstones, shales, and coals of the Casupal For-
mation (Wheeler, 1963).
In the Siquisique area south of the Falcon Basin (Fig. 1),
Oligocene to Miocene strata lie in angular unconformity over the
Paleocene and Eocene Matatere Formation (Compania Shell de
Venezuela, 1965; Wheeler, 1963). The Matatere is a thick (3,000
WEST EAST m) sequence of turbiditic shales and intercalated graywacke sand-
z stones (Bellizzia and Rodriguez, 1967). It records deposition in a
20- u
o flysch basin that extended south and west to Barquisimeto and
- 2 Carora and probably to the north under the younger Falcon
Basin. Locally within the flysch, olistostrome (Renz and others,
25- PECAVA 1955) and melange zones are present. Near Siquisique a melange
z occurs with exotic blocks of basalt, spilite, gabbro, peridotite,
ooo
LU (MARINE SHAIE)

chert, limestone, shale, and phyllite (Bellizzia and others, 1972).


These exotic block terranes were generated during the Paleogene
-
EL PARAISO
o 0 10 20
30- I l_J I (DELTAIC SANOS) emplacement of nappes southward into the Matatereflyschbasin
KM
(Stephan, 1977). Paleomagnetic studies indicate that these blocks,
Figure 2. Stratigraphic relationships of the Oligocene and lower Miocene along with many similar allochthonous terranes in northern
formations in the Falcn Basin. Modified from Daz de Gamero (1977). Venezuela, were tectonically rotated prior to or during their ac-
Section locations as indicated on Figure 1. Although quite variable, the cretion and emplacement (Skerlec and Hargraves, 1980).
thickness of the illustrated stratigraphic column is approximately 4,000 By late Eocene both flysch deposition and allochthonous
m.
block emplacement had ceased. Theflyschwas deformed prior to
frequently cross-bedded, ripple-marked, or heavily bioturbated deposition of thefirstupper Eocene Falcon Basin sediments. The
with tracks, trails, and burrows. Bedding thicknesses are variable. younger Falcon Basin was a more restricted basin with a well
The lower boundary of the El Paraso Formation is not exposed defined central axis lying about 45 km north of Siquisique. Paleo-
in the study area, whereas the upper boundary with the Pecaya cene to Eocene events preceding development of the Falcon
Formation is gradational. The El Paraso Formation is an Basin were essentially synchronous with similar compressional
eastward-prograding deltaic complex (Diaz de Gamero, 1977). events to the east described by Bell (1971) and Beck (1978) for
Thus, to the east, sandstone units and coals thin and disappear, central Venezuelan areas south of Caracas.
whereas shales become more abundant.
The Pecaya Formation is composed of a monotonously BASIN STRUCTURE
thick (>2,500 m) section of dark gray, pyritic, calcareous shales.
Bioclastic limestone lenses and small reefal limestone bodies are The Falcon Basin is a nearly east-west-trending structural as
present in the shales. These appear to be fore-reef debris shed well as sedimentary feature (Fig. 1). Throughout its history the
from the partially synochronous San Luis Formation limestones basin averaged deposition rates of 350 m/m.y. and subsidence
to the north (Fig. 3). Studies of benthic foraminiferal assemblages rates of 400 m/m.y. By the end of early Miocene time, deposition
in the shales indicate water depths of 1,000 m or more (Diaz de ceased in the central basin but was continuing in the eastern and
Gamero, 1977). These depths argue against development of the northern parts (Wheeler, 1963). During late Miocene and Plio-
limestone in its present position. Locally anomalous clastic units cene times the basin was uplifted and deformed into an east-
(San Juan de la Vega Member) are also found within the Pecaya northeast-trending anticlinorium (Wheeler, 1963; Zambrano and
Formation. others, 1971). Structural features within the anticlinorium are of
To the north and west of the area shown on Figure 3, the the following types: (1) east-west,right-lateral,transcurrent faults,
upper Pecaya Formation is overlain by, and is in part equivalent (2) east-west normal faults, (3) northwest to north normal faults,
to, the Pedregoso calcareous turbidites and shales. This lower (4) folds with east-northeast axes and east-northeast reverse
Miocene formation was deposited around the southern margin of faults, (5) west-northwest to northwest transcurrent faults, and
the San Luis Formation carbonate reefs. North of the San Luis (6) north to north-northeast transcurrent faults.
reefs, back-reefal facies are represented by the Patiecitos and These six structural types (inset, Fig. 3) are deformational
Guarabal formations. features predicted by Wilcox and others, (1973) for a terrane
220 K. W. Muessig

Figure 3. Geologic map of the central Falcon igneous area, location shown on Figure 1. Igneous
intrusive and extrusive rocks are solid black. The inset is a summary of structural types and trends found
in the Falcon Basin as a whole from published geologic maps by Bellizzia and others, (1976), Bonini and
others (1977), and Case and Holcombe (1980).

undergoing right-lateral wrenching or simple shear. However, haps the most widely known faults in northwestern Venezuela
previous structural syntheses have selectively emphasized differ- and include the Oca, Cuiza, and San Sebastin faults (Fig. 4). The
ent combinations of these structural features in deciphering the Oca fault is the best known of these. Estimates on its right-lateral
evolutionary history of the Falcon Basin. Ferrell and others, displacement range from 15 to 20 km post-Eocene (Feo-
(1969; unpublished report of the Texas Petroleum Company) Codecido, 1972) to 195 km post-Cretaceous to pre-Pleistocene
ascribed Oligocene to Miocene subsidence and uplift to simul- (Vsquez and Dickey, 1972) to 65 km post-Mesozoic (Tschanz
taneous movement on normal and transcurrent faults, and they and others, 1974). The Cuiza fault on the Guajira Peninsula is a
attributed Pliocene through Holocene folding to a north-south similar structure. Alvarez (1967) has estimated 15 to 25 km of
compression. right-lateral movement on this fault.
Vsquez and Dickey (1972) related the Falcon folds and The San Sebastin fault (Fig. 4), located just offshore central
transcurrent faults to the deformation of the incompetent Oligo- Venezuela, is frequently cited as the continuation of both Oca and
cene and Miocene sediments by right-lateral movement along El Pilar fault trends (Rod, 1956; Alberding, 1957; Vsquez and
buried basement structures represented by the Oca and related Dickey, 1972; Silver and others, 1975). Evidence of its existence
faults. Their analysis postulates a rather continuous right-lateral comes from seismic activity, the coastal Venezuelan escarpment,
movement from post-Cretaceous through Pliocene times. and seismic reflection studies. Vierbuchen (1978) estimated less
As a basis for the tectonic model presented later, the struc- than 140 km of post-Miocene, right-lateral displacement for this
tural features listed above are discussed both on a regional scale structure.
and within the area of study. The displacements on other east-west faults are poorly con-
trolled. East-west faults near Coro have frequently been depicted
East-West Transcurrent Faults as transcurrent faults (Smith, 1962; Bellizzia, 1972), but support-
ing evidence of strike-slip displacement is lacking. A fault ex-
East-west trending, right-lateral, transcurrent faults are per- plains the rapid thickening of sediments toward the Gulf of Coro
Structure and Cenozoic tectonics of the Falcon Basin 221

Figure 4. Structural map of northwestern Venezuela and adjacent areas. Shaded areas indicate zones of
relative subsidence during the Oligocene and Miocene. The Bonaire and Baja Guajira basins (unshaded)
were also areas of subsidence. Structures from Bellizzia and others (1976), Bonini and others (1977),
Ferrell and others (1969), Silver and others (1975), and Case and Holcombe (1980).

(unpublished report of the Texas Petroleum Company; Ferrell Northwest to North Normal Faults
and others, 1969), but the movement on the fault could well be
normal. Northwest- to north-trending, normal faults of Oligocene to
Evidence for east-west transcurrent faults in the central basin Miocene age are found throughout the Falcon Basin and northern
area is rather slim. Compaa Shell de Venezuela geologists have offshore areas. In the central Falcon Basin (Fig. 3) normal fault
mapped systems of discontinuous transcurrent faults with this planes dip both east and west, but the former are more abundant.
orientation in western and west central Falcon (unpublished re- Where well exposed in road-cuts, small-scale normal faults of
port of Compaa Shell de Venezuela; Jaeckli and Erdmann, similar trend exhibit the characteristics of growth faults. Changes
1952). in stratigraphie thickness indicate fault activity or growth during
sediment deposition. Changes in sandstone-unit thickness across
East- West Normal Faults faults up to 1 m have been noted in Oligocene rocks. Similar
features are impossible to detect in the monotonous, Miocene
Several east-west trending normal faults are recognized in Pecaya shales. The isolated distribution of the San Juan de La
the Falcon region, and although the data do not require it, motion Vega Member clastics within the Pecaya Formation (Fig. 2) is
on them may also include a strike-slip component. The longest is probably controlled by a northwest-trending, normal fault
the Ancon fault (Fig. 4) located south of and parallel to the bounded graben.
extension of the Oca fault. Downthrow on this fault is to the The Paraguan Peninsula is a horst bordered to the east and
north. A normal fault parallel to the southern coast of Paraguan west by north-trending normal faults (Fig. 4). East-west cross-
has a downthrown southern wall, whereas normal faults in north- sections across the peninsula (unpublished report of Corporacin
ern Paraguan have downthrown northern walls. Although sup- Venezolana del Petrleo-Institut Franais du Ptrole, 1967; Zam-
ported by field evidence, (Martin-Bellizzia and Arozena, 1972; brano and others, 1971) show these as normal faults active during
MacDonald, 1968), these faults are primarily delineated by geo- Oligocene and Miocene sedimentation. Stratigraphie units ra-
physical data (unpublished report of Corporacin Venezolana del pidly increase in thickness across faults. In La Vela Bay east of
Petrleo-Institut Franais du Ptrole, 1967; Zambrano and Paraguan, test wells penetrate thin red-bed units of probable
others, 1971). Vertical east-west faults (Fig. 5) also border the Oligocene age (Vsquez, 1975), which thicken eastward across
Siquisique ophiolite blocks and cut unconformably overlying normal faults into the Bonaire basin (unpublished report of Cor-
Oligocene sediments (Compaa Shell de Venezuela, 1965). poracin Venezolana del Petrleo-Institut Franais du Ptrole,
222 K. W. Muessig

Basin. Section location is indicated on Figure 6. Falcon igneous intrusions are solid black.

1967). Similar normal faults of northwest trend were active dur- central basin area are slightly asymmetric with steeper northern
ing sedimentation in the Puerto Cumarebo area in northeastern limbs, axial planes dipping steeply south, and limb dips of 65 to
Falcon (Payne, 1951). 45.
Structures between the islands of the Netherland and Vene-
zuelan Antilles are northwest-trending normal faults, but may West-Northwest to Northwest Transcurrent Faults
have some strike-slip components. These faults form the borders
of basins that accumulated about 2 km of Neogene sediments Small transcurrent faults of west-northwest to northwest
(Silver and others, 1975; Case, 1974; Feo-Codecido, 1971). The trend are found throughout the basin (Fig. 4) and are particularly
Bonaire Basin, which contains from 2 to 5 km of Tertiary sedi- abundant in the Puerto Cumarebo and Urumaco Trough areas.
ments (Galavis and Louder, 1971; Case, 1974; Silver and others, As noted previously, transcurrent faulting along northwest trends
1975), also contains north- to northwest-trending normal faults, has overprinted older normal faults in the Urumaco Trough.
as does La Orchila Basin (Los Roques Canyon of Peter, 1972). In the central basin area, transcurrent faults of northwest to
The widespread evidence for normal faulting contempo- west-northwest trend offset and terminate folds, reverse faults,
raneous with sedimentation suggests the existence of an exten- and topographic ridges of resistant sandstone. Fault traces
sional tectonic regime during Oligocene and Miocene time. through resistant sandstones are frequently followed by streams.
Although these north to northwest faults are rather prominent, The majority of these faults exhibit right-lateral displacement. In
their significance has largely been ignored in tectonic models of many cases a significant component of dip-slip as well as strike-
the Falcon Basin. This neglect is partly a result of recent over- slip motion is required to produce observed map patterns.
printing by transcurrent motion on the same faults. In this regard,
basement structures in the Urumaco Trough area indicate normal North to North-Northeast Transcurrent Faults
faulting (Bellizzia and others, 1976), but surface exposures of
younger formations show evidence of transcurrent displacement. Compared to other structural features in the basin, north to
Younger transcurrent faulting followed the zones of weakness of north-northeast transcurrent faults are rare. Many have both
earlier normal faults. normal and strike-slip movement. Concentrations of these faults
are found south of Coro, and in the southwest part of the Falcon
Folds and Reverse Faults Basin where fold trends are more northeasterly. These faults and
the west-northwest to northwest transcurrent faults make up a
Folds are the most conspicuous structural feature of the conjugate pair and are hereafter called conjugate strike-slip faults.
Falcon anticlinorium. Their axes trend N80E in the central and In summary, the structural trends in the Falcon system sug-
eastern basin areas and N50E in the western part. Fold trends in gest that a primary direction of right-lateral shear was directed
the offshore and Bonaire Basin areas to the east have easterly axes slightly south of east (Fig. 3). Early deformation in the Falcon
(Silver and others, 1975). Asymmetric and overturned folds are area was confined to east-west transcurrent faults and northwest
well developed south of Coro (Fig. 5) where Pliocene formations normal faults. These structures were active during the initiation of
are overturned to the north and folds are broken by southerly the basin. They controlled the rates and patterns of basin sedi-
dipping thrust faults. Reverse faults of similar trend are found mentation and subsidence.
throughout the basin and within the study area. Anticlines in the Folds and reverse faults and the conjugate strike-slip faults
Structure and Cenozoic tectonics of the Falcon Basin 223
69 50' 69 30'

69 50' 69 30'

Figure 6. Map of magnetic anomalies and Bouguer gravity anomalies in the central Falcon area.
Magnetic anomalies are contoured in 25-gamma intervals (dashed lines); the +25-gamma closures are
shaded. Bouguer gravity anomalies (solid lines) are contoured in 5-mgal increments supplemented by
1-mgal contours where data permits. Locations of major igneous exposures (*) are indicated.

are later deformational features imposed during late Miocene to nini, 1978) exhibits several noteworthy features. For most of the
Pliocene time. southern Caribbean borderland, the zero Bouguer anomaly con-
tour approximates the Caribbean shoreline. The Falcon Basin is
GRAVITY A N D M A G N E T I C A N O M A L I E S the largest regional inland deflection of the zero anomaly in
Venezuela, anomalies reaching +50 mgal (Fig. 1). The central
The regional gravity field of the Falcon and Bonaire basins basin area is characterized by positive Bouguer anomalies up to
and adjacent area (Bowin, 1976; Bonini and others, 1977; Bo- 30 mgal (Fig. 6).
224 K. W. Muessig
In the Bonaire Basin, where Oligocene and Miocene sedi- TABLE 1. REPRESENTATIVE WHOLE-ROCK ANALYSES

mentary thicknesses are similar to the Falcon Basin, Silver and


others (1975) used gravity to model a dense, thinned continental Alkali olivine Trachybasalt Subalkaline
crust with 6 km of sediments and a crust-mantle boundary ele- basalt basalt

vated to a depth of about 18 km. In the Falcon Basin a similar


model would satisfy the observed gravity. In both regions the
Si0 2 45.45 44.52 50.37

A1 ? 0,
positive Bouguer anomalies indicate a density excess in the basins
15.69 15.64 15.18

when compared to adjacent continental areas. Injection of dense 9.57 10.99 8.44

basaltic material into the basin crust and sediments occurred in MgO 5.17 5.33 9.20

the central Falcon Basin. Similar intrusive activity or general CaO 6.29 6.31 8.36

crustal thinning may account for the positive anomalies elsewhere Na 2 0 5.14 4.39 2.99

in the region. K2O 1.79 2.59 0.88

The regional magnetic field in the central Falcon Basin is Ti0 2 2.31 2.40 1.72

rather featureless. Locally, however, thefieldis influenced by the MnO 0.17 0.23 0.14

presence of igneous intrusions and by the basin structure. Anom- LOI 8.25 7.08 4.56

alies up to 225 gammas are directly associated with igneous TOTAL 99.83 99.48 101.84

intrusions. Larger regions of regional northwest-trending mag- Trace elements, ppm


netic anomalies up to 25 gammas (Fig. 6) are also loosely asso-
ciated with outcropping igneous rocks. These elongate trends Cr 100 40

suggest that the exposed shallow-level sills,flows,and plugs may


Ni 36 45
Cu 25 24
be fed by northwest-trending dikes at depth. Thus the present Zr 270 300

northeast trend of igneous outcrops is an artifact of the younger Y


Rb
30 30
33
fold trends. The northwest magnetic anomalies may also reflect
22.8
Ba 600 650
the normal faults of similar trend discussed previously. Sr
87 S r /86
1027
0.7036
1005
Sr

FALCON IGNEOUS ROCK PETROLOGY Note: Alkali olivine basalt from Cerro Atravesado, trachy-
basalt from La Guarabitas, and subalkaline basalt from Cerro

Within the central Falcon Basin, basaltic igneous bodies


Manaure. For analytical uncertainties see Muessig (1979).

crop out in a northeast-trending 50-km by 10-km belt (unpub-


lished report of Compania Shell de Venezuela; Brueren, 1949; netite, and kaersutite, are set in a groundmass of the above phases
Coronel, 1970). Outcrop patterns, structural orientations, and the with the addition of late stage biotite and analcite. These rocks
occurrence of border-zones with chilled igneous margins, contact are nepheline normative, enriched in alkalis (Table 1) and plot in
metamorphism, local sedimentary rock inclusions, and breccias the field of "within plate" basalts as defined by Pearce and Cann
were documented by Muessig (1979). These bodies intrude the (1973) on a Ti, Zr, and Y triangular diagram. Approximately
Oligocene sedimentary rocks as shallow-level sills and plugs or 40% of the exposed igneous rocks have these alkaline affinities.
occur as extrusive flows. Potassium-argon isotopic analyses of a The transitional to subalkaline types are extensively altered
trachybasaltic sill yield an age of 22.9 0.9 m.y. (Muessig, 1978). diabase intrusions with relict augite phenocrysts, plagioclase, il-
This age is 5 m.y. younger than reported paleontologic ages of menite, and pseudomorphs of magnetite and olivine. Compared
surrounding sedimentary rocks. A span of igneous activity from to the alkaline basalts, these rocks are lower in iron, titanium, and
28 to 23 m.y. can be inferred from stratigraphic constraints. The alkalis but higher in silica. Alteration has obscured their nature by
Falcon igneous rocks are younger than and distinctive in mor- introducing silica and leaching alkalis.
phology, composition, and origin from both the Siquisique ophio- Assorted xenoliths and amphibole megacrysts are abundant
lite blocks (35 km south) and the Paraguana igneous-metamor- in alkali basalt lavas near vents at Atravesado, La Azulita, El Sol,
phic massif (100 km north). and Paraguachoa (Fig. 3). Calc-silicate gneiss is the most com-
The Falcon igneous rocks are similar in petrology to other mon xenolith type (95%), followed by spinel peridotite (4%),
continental igneous rock suites associated with extensional tec- gabbro (1%), and rare schist and granite xenoliths. Most of the
tonic environments. Basaltic rocks compose about 98% of the xenoliths represent a sampling of the underlying crust and mantle
igneous suite, whereas felsic rocks are of only minor abundance. through which the basaltic magmas passed. The megacrysts and
Chemical and mineralogic analyses reveal two parental rock some of the gabbros may be samples of deeper level magma
types: (1) an alkaline type and (2) a less alkaline type transitional chambers that were the source of the basalts.
to subalkaline (Muessig, 1979).
Rocks of thefirsttype are alkali olivine basalts and diabases, TECTONIC MODEL
hawaiites, trachybasaltic diabases, and trachytes. Phenocrysts of
olivine and titanaugite, together with sporadic plagioclase, mag- The development of a diffuse pullapart basin between paral-
Structure and Cenozoic tectonics of the Falcon Basin 225
(a) (b)

PRE-OLIGOCENE\
\ BONAIRE \
FALCON \ BASIN \

LOS M O N J E S
ARUBA
CUiZa
CURACAO
BONAIRE
LAS A V E S
LOS ROQUES
LA O R C H I L A
PARAGUANA

Figure 7. The evolution of the Falcon and Bonaire pullapart basins. Options (a) and (b) represent two
configurations of the area prior to extension in the late Eocene. To evolve to the present configuration
(c), option (a) requires a greater amount of Oligocene to Miocene extension and right-lateral transcur-
rent displacement. In option (b) a pre-Oligocene Bonaire basin has a thinner, more mafic crust and
requires less extension to evolve to (c).

lei but offset right-lateral transcurrent faults is postulated for the the Salton Trough, for example, 6 km of Pliocene to Holocene
Falcon region. This extensional tectonic model accounts for (1) a sediments were deposited in a pullapart structure while adjacent
complex Oligocene-Miocene stratigraphy and rapid basin subsi- areas underwent rapid erosion. A rhomb-shaped trough subsided
dence history, (2) synsedimentary east-west transcurrent and east- at 600 m/m.y. due toright-lateralmovement along offset faults in
west and northwest normal faulting, (3) positive gravity the San Andreas system. Basaltic and felsic igneous activity is
anomalies indicative of a relatively thin or dense crust, and concentrated along the normal fault trends that connect the offset
(4) alkaline basaltic volcanism and intrusion along northwest transcurrent faults (Hill, 1977).
trends. The development of pullapart structures in the Falcon Basin
A pullapart basin is a zone of crustal extension and subsi- reflects the Tertiary tectonic evolution of a larger crustal block
dence that develops at a bend in a transcurrent fault system if the termed the Bonaire Block by Silver and others (1975). This block
motion at the bend is divergent (Crowell, 1974). Subsidence was also subject to right-lateral shearing between the Caribbean
occurs on normal faults that connect parallel sections of the trans- and South American plates (Fig. 7).
current fault system. Significant subsidence may also occur along
the transcurrent faults. Due to the distribution of stresses within Falcon Basin Pullapart
this zone of divergence, the bordering normal and transcurrent
faults define a rhomb-shaped basin in map view. In cross-section In the Falcon pullapart system, extension was dispersed over
the structure of the basin is grabenlike. If the pullapart basin is a wide zone resulting in zones of relative stability and subsidence
intruded extensively by mafic magma and develops a new simatic (Fig. 4). The Paraguana, Dabajuro, Guajira, and southern main-
crust, then it is a rhombochasm as defined by Carey (1958). land (Siquisique) areas remained as highs shedding sediments into
Within a large system of transcurrent faults, pullapart struc- grabenlike, pullapart zones. The Falcon Basin, Urumaco Trough,
tures may be rather complicated. A classic example has been La Vela Bay, and Bonaire Basin were major zones of subsidence.
documented along the San Andreas fault system by Elders and On a small scale, rhombic grabens are well defined in the
others (1972), Wilcox and others (1973), and Crowell (1974). In Falcon region. A rhombic pullapart basin with distinct structural
226 K. W. Muessig
boundaries and thick sediment accumulations is located north of anomalies across the basin indicates at least 5 km of crustal
the Ancon fault (Fig. 4) in the western Falcon Basin. It devel- thinning (Silver and others, 1975). Lacking definitive information
oped where movement on the Oca fault was transferred on the nature of the basement underlying the Bonaire Basin, the
southward to the Ancon fault. The Urumaco Trough is also a following hypotheses are feasible: (1) the basement and the over-
distinct pullapart structure connecting the Cuiza and Oca fault lying sediments of the basin have been injected with large
trends. amounts of mafic material, or (2) the crust was originally thinner
The central Falcn Basin is a more diffuse zone of subsi- or more mafic prior to basin development. In the second option a
dence. A northern boundary fault at the southern margin of the preexisting Bonaire Basin (Fig. 7b) would be extended about 150
San Luis limestone platform (Fig. 3) is postulated due to the km during the Oligocene to Miocene pullapart event. The first
relative stability and the shallow depositional environment of the option requires about 300 km of east-west extension to produce
limestone reefs, in contrast to the deep water, rapidly subsiding the Bonaire Basin and would imply a major east-west transcur-
environment of the synchronous Pecaya shales. The southern rent fault between the basin and the island chain (Fig. 7a). The
border fault follows east-west scarps in the Siquisique area, which transcurrent fault allows extension and subsidence in the basin to
actively shed coarse Miocene sediments northward. exceed that along the island chain.
The eastern Falcn Basin and La Vela Bay also exhibit Additional evidence comes from the Eocene or older Soebi
widespread subsidence, which increases in degree to the north Blanco conglomerates of Bonaire, which contain gneiss, schist,
and east. The southern border is along fault trends south of the and marble cobbles indicative of a mainland source (Pijpers,
Cerro Misin uplift (Fig. 4). The uplift represents a horst that 1933; Beets and others, 1977). Silver and others (1975) inferred
remained high during subsidence of the surrounding area since that the Bonaire Basin formed later than deposition of this con-
early Oligocene time (Hunter, 1972). Northwesterly normal faults glomerate. Removing the Oligocene to Miocene extension be-
with downthrown eastern blocks in the Gulf of Triste indicate tween the Netherlands Antilles and Paraguan, as proposed by
that the gulf developed as a pullapart basin to the north of the San the pullapart model, would place Bonaire and/or Curaao adja-
Sebastin transcurrent fault. cent to Paraguan (Fig. 7). Thus, Jurassic basement rocks on
In addition to pullapart zones of divergence, irregular trans- Paraguan could be the source rocks for the conglomerates.
current fault systems inevitably develop zones of convergence Contrary to the suggestion of Silver and others (1975), the
with associated compressional deformation (Wilcox and others, Bonaire Basin may already have existed as a basin (Fig. 7b)
1973; Crowell, 1974). The southern border of the Falcn pulla- during Eocene conglomerate deposition as long as the island
part developed zones of convergence particularly in the western chain was continuous from Paraguan to Bonaire. Extension and
basin. Convergence caused distortion of fold axes in this area to breakup of the island chain, however, must have been
N20E orientations and development of north- to northeast- post-Eocene.
trending transcurrent faults.
Bonaire Block Pullapart
Bonaire Basin Pullapart
The Bonaire Block, bounded by the Curaao Ridge on the
Stratigraphie and structural characteristics of the Falcn north and the Oca-San Sebastin faults on the south, is a small
Basin appear to extend into the Bonaire Basin (Silver and others, crustal block between the Caribbean and South American plates
1975). Northwest-trending normal and east-west-trending nor- (Fig. 8). Silver and others (1975) concluded that nonrigid defor-
mal and transcurrent faults controlled sediment distribution and mation of this block occurred during its clockwise rotation be-
subsidence in the Bonaire Basin. The San Sebastin transcurrent tween the major crustal plates.
fault borders the southern margin of the basin (Fig. 4). East-west However, the nature of many of the Neogene basins within
faults south of Aruba and Curaao and along the northern basin this block suggests that a pullapart model may explain their ori-
margin were interpreted by Silver and others (1975) from seismic gins. Case (1974) concluded that the northwest- and east-west-
data as normal faults, but transcurrent motion on these faults is trending Colombian basins (Baja Guajira, Chichibacoa; Fig. 4),
also possible. In conjunction the northwest and east-west fault in addition to the Venezuelan basins between the offshore islands,
trends outline the rhombic shape of the Bonaire Basin pullapart originated from north to northwest extension within the Bonaire
structure. Block. The model presented here redefines the Bonaire Block as a
The structures adjacent to the Netherlands and Venezuelan pullapart zone that experienced east to northeast extension be-
Antilles are also consistent with a pullapart origin. The sedimen- tween the transcurrent faults that border the Caribbean and South
tary basin between Aruba and Curaao, for example, exhibits the American plates (Fig. 8). Extension occurred mainly during
typical rhombic form with surrounding normal faults (Fig. 4). Oligocene and Miocene times.
Bouguer gravity anomalies for most of the Bonaire Basin are Geologic and seismic evidence suggest that the Oca, al-
positive (Bonini, 1978; Bonini and others, 1977) despite sedimen- though active, and Cuiza transcurrent faults are not presently
tary thicknesses of 2 to 6 km. As discussed previously this indi- zones of major strike-slip displacement (Molnar and Sykes,
cates a thin or relatively dense crust. Modeling of free-air 1969). Instead, Caribbean-South American plate motion from
Structure and Cenozoic tectonics of the Falcon Basin 227
TABLE 2. EXTENSION ESTIMATES WITHIN THE NORTHERN
PART OF THE BONAIRE BLOCK

Basin Sediment Total East-west


thickness subsidence extension
BONAIRE (km) (km) (km)

PULLAPART Chichibacoa 4. 0 4.,2

\ 1.5 1.
.7
1. 1.
los Monjes-Aruba
BLOCK Aruba-Paraguan 5 .7
Curacao-Aruba 2. 0 3.,1
Bonaire-Curacao 2. 0 3.,5
1. 2.,7
0.
SAN SEBASTIAN FAULT Las Aves-Bonaire 0
5 2..5
0. 1.
Los Roques-Las Aves
8 ,8
1.
La Orchila-Los Roques
SOUTH AMERICAN PLATE La Orchila Basin 2 5..1

Figure 8. The Bonaire Block pullapart structure and its relationship to TOTAL 26.,3 35.1

the Caribbean and South American plates. The block is identical to that Urumaco Trough 8. 3
0
8..5

outlined by Silver and others (1975). Baja Guajira 3. 3.,0


TOTAL (Minus Chichibacoa*) 33,.6 44.8

Note: Subsidence equals sediment thickness plus the depth of


water in the basin. A basin name with a hyphen (e.g., Curacoa-
late Miocene to Holocene was transferred from the El Pilar-San Aruba) indicates the basin between the two areas. Sediment

Sebastin, right-lateral fault systems to a broad zone of compres-


thickness data compiled from Edgar and others (1971), Zambrano
and others (1971), Silver and others (1975), Peter (1972),
sional deformation and transcurrent movement within the Bo- Case (1974), and corporacin Venezolana del Petrleo-Institut

naire Block, and along the Bocon fault and the Curaao Ridge
Francais du Ptrole (unpublished report, 1967).
Alternate estimate of extension not including the Chichibacoa

(Silver and others, 1975; Jordan, 1975; Pennington, 1981). These Basin is explained in the text.

events are responsible for the overprinting of earlier Oligocene


and Miocene northwest-trending normal faults in the Falcon area consider crustal stretching. (4) Estimates from basement depths
by transcurrent motion, folding, and thrusting. do not account for extension by magma injection. (5) Normal
faults may shallow at depth allowing greater extension. (6) Esti-
Extension within the Bonaire Block mates of sediment thickness are probably minimum estimates.
An estimate on the minimum amount of extension within TECTONIC SUMMARY
the Bonaire Block pullapart has been calculated assuming that
Eocene "basement" sediments (high velocity) attained their pres- Prior to initiation of the Falcon Basin, Paleocene to Eocene
ent depths by subsidence along normal faults. A similar type of rocks at the southern border of the basin recorded the culmina-
analysis was presented by lilies (1970) to estimate extension in tion of a compressional orogen involving flysch deposition,
the Rhine Graben. Planar normal faults with dips of 60 are thrusting, ophiolite obduction, tectonic rotations, and olisto-
assumed and are consistent with seismic profiles (Silver and oth- strome development (Skerlec and Hargraves, 1980; Beets and
ers, 1975). The calculation is made for the northern part of the others, 1984). The Antillean islands formed a contiguous unit
pullapart along the Antillean chain primarily because depth to connected to the Paraguan Peninsula and the Faltn area (Fig. 7
"basement" is controlled by a variety of geophysical data. The and 9A).
Falcon and Bonaire basins were avoided because of the asymmet- A major change in the deformation style occurred during the
ric, half-graben faulting and later erosional effects. The basins Eocene in northern Venezuela. The change is marked by the
included are the La Orchila Basin, the basins between all the cessation of southward thrusting and flysch deposition in both
Antillean Islands, the basin between Aruba and Paraguana, and western and central Venezuela by the end of middle Eocene time
the Chichibacoa Basin (Fig. 7). Estimated post-Eocene, east-west (Bell, 1972; Maresch, 1974; Stephan, 1977; Beck, 1978). East-
extension from the Guajira Peninsula to the La Orchila Basin is west, right-lateral, transcurrent faulting began in northern South
about 35 km (Table 2). The value is increased to 45 km by using America by late Eocene time. Andesitic volcanism and subduc-
the Urumaco Trough instead of the Chichibacoa Basin and by tion in the Lesser Antilles also started during the Eocene (Bunce
adding an extension estimate for the Baja Guajira Basin. and others, 1971; Arculus, 1976). This implies that the Caribbean
By this argument, at least 35 to 45 km of extension occurred plate had begun to move eastward with respect to the South
within the Bonaire Block during its Oligocene and Miocene de- American plate regardless of its pre-Eocene displacement (Bell,
formation between the Caribbean and South American plates. 1972; Beets and others, 1984).
This is a minimum extension for the following reasons: (1) Major The Falcon Basin originated by extension during late Eo-
faults along the northern border of the Bonaire Basin may allow cene to Oligocene time within a pullapart zone caused by trans-
greater amounts of east-west extension to the south. (2) Extension current motion between the Caribbean and South American
within stable blocks (Paraguana, Guajira) was not considered. plates (Fig. 9B). Indeed, the Faltn Basin preserves a record of
(3) Estimates are based on a brittle fracture model that does not the sedimentation and deformation that occurred within this zone
228 K. W. Muessig
(g) EARLY EOCENE South American plates (Fig. 9B). Minimum extension within the
northern part of the block was at least 35 to 45 km in an east-
west direction. The extension within this nonrigid block must be
O* y M considered when estimating Caribbean-South American plate
NORTH Y <* SOUTH movements.
Pullapart structures also developed elsewhere along the Car-
OCEANIC CRUST TRANSITIONAL
CRUST CONTINENTAL ribean-South American plate boundary at this time. The La Or-
CRUST
chila Basin and the Cariaco Trench are excellent examples.
Similar structural basins formed along the northern margin of the
Caribbean plate; however, here the motion was in a left-lateral
sense with respect to the North American plate (Burke and oth-
(D LATE OLIGOCENE ers, 1980).
I BONAIRE BLOCK
:
, <F> In the Bonaire Block, folds, reverse faults, and conjugate sets
of secondary transcurrent faults formed later during the develop-
ment of the pullapart and were amplified by greater compres-
sional stresses during the late Miocene to Holocene tectonic
evolution (Fig. 9C). Shallow underthrusting of the block by the
Caribbean plate formed the Curaao Ridge accretionary prism
(Silver and others, 1975; Jordan, 1975; Ladd and Watkins,
1979). The Falcon anticlinorium was formed and uplifted by the
addition of compressional stress across this generally transform-
type plate boundary zone between the Caribbean and South
America.
EAST WEST TRANSCURRENT AND NORMAL FAULTS

ACKNOWLEDGMENTS
This research was supported by the Princeton University
PRESENT FALCON Department of Geological and Geophysical Sciences and the
ANTICLINORIUM Venezuela Ministerio de Energia y Minas. I am indebted to the
many geologists from these organizations who provided field as-
sistance, suggestions, criticism, and above allstimulating discus-
sions. Among these I am particularly grateful to A. Bellizzia, J.
Rios, R. Hargraves, W. Bonini, J. Suppe, R. Vierbuchen, and M.
Diaz de Gamero. I thank T. Harding and M. Schuepbach of
Exxon Production Research Company for discussing their ideas
on wrench faulting and pullapart basins. I also thank H. A. Jordi
of Shell Internationale Petroleum and the Venezuela Ministerio
de Energia y Minas for providing copies of unpublished geologic
0 50 100
studies of the Falcon region.
km
Figure 9. North-south, schematic cross sections illustrating the develop- REFERENCES CITED
ment of the Falcon Basin and the Bonaire Block (discussed in the text)
from late Eocene to present. Extension in the basin occurs in an east-west Alberding, H., 1957, Application of principles of wrench-fault tectonics of Moody
direction perpendicular to the section. Scale is approximate with no and Hill to northern South America: Geological Society of America Bulletin,
vertical exaggeration. On transcurrent faults, A indicates the block mov- v. 68, p. 785-790.
ing away from the reader and T is toward the reader. Alvarez, W., 1967, Geology of the Simurua and Carpintero areas, Guajira Penin-
sula, Colombia [Ph.D. thesis]: Princeton University, 168 p.
Arculus, R. J., 1976, Geology and geochemistry of the alkali basalt-andesite
of transcurrent faulting. Right-lateral motion along offset, trans- association of Grenada, Lesser Antilles island arc: Geological Society of
current faults caused extension in a northeast direction by normal America Bulletin, v. 87, p. 612-624.
faulting along northwesterly trends, subsidence, crustal thinning, Beck, C. M., 1978, Polyphase Tertiary tectonics of the interior range in the central
and injection of basaltic alkaline magmas (Fig. 9B). part of the western Caribbean chain, Guarico State, northern Venezuela:
On a iarger scale the Bonaire Block of Silver and others Beets,Geologie Mijnbouw, v. 57, p. 99-104.
D. J., MacGillavry, H. J., and Klaver, G., 1977, Geology of the Cretaceous
(1975) originated as a diffuse zone of pullapart structures during and early Tertiary of Bonaire, in Guide to the field excursions on Cura$o,
the Oligocene right-lateral motion between the Caribbean and Bonaire, and Aruba, Netherlands Antilles: Caribbean Geological Confer-
Structure and Cenozoic tectonics of the Falcon Basin 229
ence, VIII, Curaao Netherlands Antilles, p. 18-28. Feo-Codecido, G., 1971, Geologa y recursos naturales de la Pennsula de Para-
Beets, D. J., Maresch, W. V., Klaver, G. Th., Mottana, A., Bocchio, R., and guan, in Symposium on Investigations and Resources of the Caribbean Sea
Beunk, F. F., 1984, Magmatic rock series and high-pressure metamorphism and Adjacent Regions: UNESCO, Paris, France, p. 231-240.
as constraints on the tectonic history of the southern Caribbean, in Bonini, Feo-Codecido, G., 1972, Breves ideas sobre la estructura de la Falla de Oca,
W. E., and others, eds., The Caribbean-South American plate boundary and Venezuala: Transactions, Caribbean Geological Conference, VI, Margarita,
regional tectonics: Geological Society of America Memoir 162 (this Venezuela, 1971, p. 184-190.
volume). Galavis, J. A., and Louder, L. W., 1971, Preliminary studies on geomorphology,
Bell, J. S., 1971, Tectonic evolution of the central part of the Venezuelan coast geology and geophysics on the continental shelf and slope of northern South
ranges, in Donnelly, T. W., ed., Caribbean geophysical, tectonic, and petro- America: Eighth World Petroleum Congress, Caracas, Venezuela, Proceed-
logic studies: Geological Society of America Memoir 130, p. 107-118. ings, v. 2, p. 107-120.
Bell, J. S., 1972, Global tectonics in the southern Caribbean area, in Shagam, R., Hill, D. P., 1977, A model for earthquake swarms: Journal of Geophysical
and others, eds., Studies in earth and space sciences: Geological Society of Research, v. 82, p. 1347-1352.
America Memoir 132, p. 369-386. Hunter, V. F., 1972, A middle Eocene flysch from east Falcn, Venezuela:
Bellizzia, A., 1972, Is the entire Caribbean Mountain belt of northern Venezuela Transactions, Caribbean Geological Conference, VI, Margarita, Venezuela,
allochthonous? in Shagam, R., and others, eds., Studies in earth and space 1971, p. 128-130.
sciences: Geological Society of America Memoir 132, p. 363-368. lilies, J. H., 1970, Graben tectonics in relation to crust mantle interaction, in
Bellizzia, A., and Rodrquez, D., 1967, Guia de la excursion a la regin de lilies, J. H., and Muller, S., eds., Graben problems: Stuttgart, Schweizerbart,
Duaca-Barquisimeto-Bobare: Venezuela Boletn Geologa, v. 8, p. 289-309. p. 4-27.
Bellizzia, A., Rodrguez, D., and Graterol, M., 1972, Ofiolitas de Siquisique y Rio Jordan, T., 1975, The present-day motions of the Caribbean plate: Journal of
Tocuyo y sus relaciones con la fallas de Oca: Transactions, Caribbean Geo- Geophysical Research, v. 80, p. 4433-4439.
logical Conference, VI, Margarita, Venezuela, 1971, p. 182. Ladd, J. W., and Watkins, J. S., 1979, Tectonic development of trench-arc
Bellizzia, A., Pimentel, N., and Bajo, R., compilers, 1976, Mapa geolgico estruc- complexes on the northern and southern margins of the Venezuela Basin, in
tural de Venezuela: Venezuela Ministerio de Minas e Hydrocarbures, escala Watkins, J. S., Montadert, L., and Dickerson, P. W., eds., Geological and
1:500,000. geophysical investigations of continental margins: American Association of
Bonini, W. E., 1978, Anomalous crust in the Eastern Venezuela Basin and the Petroleum Geologists Memoir 29, p. 363-371.
Bouguer gravity anomaly field of Northern Venezuela and the Caribbean MacDonald, W. D., 1968, Estratigrafa, estructura y metamorfismo de las rocas
borderland: Geologie Mijnbouw, v. 57, p. 117-122. del Jursico Superior, Pennsula de Parguan, Venezuela: Venezuela Boletn
Bonini, W. E., Pimstein de Gaeta, C., and Graterol, V., compilers, 1977, Mapa de Geologa, v. 9, p. 441-458.
anomalas gravimtricas de Bouguer de la parte norte de Venezuela y areas Malfait, B. T., and Dinkelman, M. G., 1972, Circum-Caribbean tectonic and
vecinas: Venezuela Ministerio de Energa y Minas, escala 1:1,000,000. igneous activity and the evolution of the Caribbean plate: Geological So-
Bowin, C., 1976, The Caribbean: Gravity field and plate tectonics: Geological ciety of America Bulletin, v. 83, p. 251-272.
Society of America Special Paper 169, 79 p. Maresch, W. V., 1974, Plate tectonics origin of the Caribbean Mountain system of
Bunce, E. T., Phillips, J. D., Chase, R. L., and Bowin, C. O., 1971, The Lesser northern South America: Discussion and proposal: Geological Society of
Antilles and the eastern margin of the Caribbean Sea, in Maxwell, A. E., ed., America Bulletin, v. 85, p. 669-682.
The sea, Volume 4, Part II: New York, John Wiley & Sons, p. 359-385. Martn-Bellizzia, C., and Arozena, J.M.I., 1972, Complejo ultramfico zonado de
Burke, K., Grippi, J., and Sengor, A.M.C., 1980, Neogene structures in Jamaica Tausabana-El Rodeo, gabro zonado de Siraba-Capuana y complejo
and tectonic style of the northern Caribbean plate boundary zone: Journal of subvolcnico estratificado de Santa Ana, Paraguan, Estado Falcn:
Geology, v. 88, p. 375-386. Transactions, Caribbean Geological Conference, VI, Margarita, Venezuela,
Carey, S. W., 1958, A tectonic approach to continental drift: University of Tas- 1971, p. 337-356.
mania Department of Geology Symposium, p. 177-355. Molnar, P., and Sykes, L. R., 1969, Tectonics of Caribbean and Middle America
Case, J. E., 1974, Major basins along the continental margin of northern South regions from focal mechanisms and seismicity: Geological Society of Amer-
America, in Burk, C., and Drake, C., eds., The geology of continental mar- ica Bulletin, v. 80, p. 1639-1684.
gins: New York, Springer-Verlag, p. 733-741. Muessg, K. W., 1978, The central Falcn igneous suite, Venezuela: Alkaline
Case, J. E., and Holcombe, T. L., 1980, Geologic-tectonic map of the Caribbean basaltic intrusions of Oligocene-Miocene age: Geologie Mijnbouw, v. 57,
region: U.S. Geological Survey Miscellaneous Investigations Map 1-1100, p. 261-266.
scale 1:2,500,000. 1979, The central Falcn igneous rocks, northwestern Venezuela: Their
Compaa Shell de Venezuela, 1965, Igneous rocks of the Siquisique region, State origin, petrology and tectonic significance [Ph.D. thesis]: Princeton Univer-
of Lara: Asociacin Venezolana de Geologa Minera y Petrleo Boletn sity, 252 p.
Informativo, v. 8, p. 286-305. Payne, A. L., 1951, Cumarebo oilfield,Falcn, Venezuela: American Association
Coronel, G., 1970, Igneous rocks of central Falcn: Associacin Venezolana de of Petroleum Geologists Bulletin, v. 35, p. 1850-1878.
Geologa Minera y Petrleo Boletn Informativo, v. 13, p. 155-162. Pearce, J. A., and Cann, J. R., 1973, Tectonic setting of basic volcanic rocks
Crowell, J. C., 1974, Sedimentation along the San Andreas Fault, California, in determined using trace element analyses: Earth and Planetary Science Let-
Dott, R. H., and Shaver, R. H., eds., Modern and ancient gosynclinal ters, v. 19, p. 290-300.
sedimentation: Society of Economic Paleontologists and Mineralogists Spe- Pennington, W. D., 1981, Subduction of the eastern Panama Basin and
cial Publication 19, p. 292-303. seismotectonics of northwestern South America: Journal of Geophysical
Daz de Gamero, M. L., 1977, Estratigrafa y micropaleontologa del Oligocene y Research, v. 86, p. 10753-10770.
Miocene inferior del centro de la Cuenca de Falcn, Venezuela: Escuela de Peter, G., 1972, Geologic structure offshore north-central Venezuela: Transac-
Geologa y Minas, Universidad Central de Venezuela, Caracas, GEOS no. tions, Caribbean Geological Conference, VI, Margarita, Venezuela, 1971,
22, p. 3-60. p. 283-294.
Edgar, N. T., Ewing, J. I., and Hennion, J., 1971, Seismic refraction and Pijpers, P. J., 1933, Geology and paleontology of Bonaire (Dutch West Indies)
reflection in Caribbean Sea: American Association of Petroleum Geologists [Ph.D. thesis]: University of Utrecht, Netherlands, 103 p.
Bulletin, v. 55, p. 833-870. Renz, O., Lakeman, R., and Van der Muelen, E., 1955, Submarine sliding in
Elders, W. A., Rex, R. W., Meidav, T., Robinson, P. T., and Biehler, S., 1972, western Venezuela: American Association of Petroleum Geologists Bulletin,
Crustal spreading in southern California: Science, v. 178, p. 15-24. v. 39, p. 2053-2067.
230 K. W. Muessig
Rod, E., 1956, Strike-slip faults of northern Venezuela: American Association of Geologic evolution of the Sierra Nevada de Santa Marta, northeastern
Petroleum Geologists Bulletin, v. 40, p. 457-476. Colombia: Geological Society of America Bulletin, v. 85, p. 273-284.
Santamara, F. J., and Schubert, C., 1974, Geochemistry and geochronology of Vsquez, E., 1975, Results of exploration in La Vela Bay: Ninth World Petro-
the southern Caribbean-northern Venezuela plate boundary: Geological So- leum Congress, Tokyo, Japan, Proceedings, v. 3, p. 195-197.
ciety of America Bulletin, v. 85, p. 1085-1098. Vsquez, E., and Dickey, P., 1972, Major faulting in northwestern Venezuela and
Silver, E. A., Case, J. E., and MacGillavry, H. J., 1975, Geophysical study of the its relation to global tectonics: Transactions, Caribbean Geological Confer-
Venezuelan borderland: Geological Society of America Bulletin, v. 86, ence, VI, Margarita, Venezuela, 1971, p. 191-202.
p. 213-226. Vierbuchen, R. C., Jr., 1978, The tectonics of northeastern Venezuela and the
Skerlec, G. M., and Hargraves, R. B., 1980, Tectonic significance of paleomag- southeastern Caribbean Sea [Ph.D. thesis]: Princeton University, 175 p.
netic data from northern Venezuela: Journal of Geophysical Research, v. 85, Wheeler, C. B., 1963, Oligocene and lower Miocene stratigraphy of western and
p. 5303-5315. northeastern Falcn Basin, Venezuela: American Association of Petroleum
Smith, F. D., Jr., 1962, Mapa geolgico-tectnico del Norte de Venezuela: Primer Geologists Bulletin, v. 47, p. 35-68.
Congreso Venezolana Petroleo, escalo 1:1,000,000. Wilcox, R. E., Harding, T. P., and Seely, D. R., 1973, Basic wrench tectonics:
Stainforth, R. M., 1969, The concept of seafloor-spreading applied to Venezuela: American Association of Petroleum Geologists Bulletin, v. 57, p. 74-96.
Asociacin Venezolana de Geologa, Minera y Petrleo Boletn Informa- Zambrano, E., Vsquez, E., Duval, B., Latreille, M., and Coffinieres, B., 1971,
tivo, v. 12, p. 257-274. Sntesis paleogeogrfica y petrolera del occidente de Venezuela: Memoria,
Stephan, J. F., 1977, Una Interpretacin de los complejos con bloques asociados a Congreso Geolgico Venezolano, IV, Caracas, 1969, Venezuela Ministerio
los flysch Paleocene-Eoceno de la cadena Caribe Venezolana: el emplaza- de Minas e Hidrocarburos, Boletn de Geologa, Publicacin Especial no. 5,
miento submarino de la napa de Lara: Caribbean Geological Conference, v. 1, p. 481-552.
VIII, Curaao Netherlands Antilles, Abstracts, p. 199-200.
Tschanz, C. M Marvin, R. F., Cruz, B., Mehnert, H. H., and Cebula, G. T., 1974, MANUSCRIPT ACCEPTED BY THE SOCIETY SEPTEMBER 1,1983

Printed in U.S.A.

S-ar putea să vă placă și