Download as pdf or txt
Download as pdf or txt
You are on page 1of 46

2

Altamira, Armando, and Kevin Burke, 2015, The Ribbon Continent of South America
in Ecuador, Colombia, and Venezuela, in C. Bartolini and P. Mann, eds.,
Petroleum geology and potential of the Colombian Caribbean Margin:
AAPG Memoir 108, p. 39–84.

The Ribbon Continent of South America


in Ecuador, Colombia, and Venezuela
Armando Altamira
Instituto Mexicano del Petróleo, Eje Central Lázaro Cárdenas Norte #152 Col. San Bartolo Atepehuacan,
Delegación Gustavo A. Madero, Distrito Federal, C.P 07730, México (e-mail: aareyan@imp.mx)

Kevin Burke
University of Houston, Department of Earth and Atmospheric Sciences, 312 Science and Research Bldg.
1, Rm. 312, Houston, Texas 77204, U.S.A. (e-mail: geos63@central.uh.edu)

Abstract
Isotopic ages in volcanic arc igneous and subduction complex rocks in Venezuela and on the
­offshore island of Aruba are consistent with the finding that the ages of arc igneous a­ ctivity
and high pressure–low temperature metamorphism in both of those areas are restricted to
times ­between ca 150 Ma and ca 70 Ma. That age range, and the restriction of fossil ages in the
­subduction complexes to between mid-Jurassic (ca 170 Ma) and late Cretaceous (ca 70 Ma) times,
reveals a match to the ages of volcanic arc rocks that were involved in collision with the ­Andean
Margin of Ecuador more than 2000 km (1242.7 mi) away. The similarity of ages can be explained
if the rocks in both areas are those of the Great Arc of the Caribbean, which has been considered
to have collided with the west coast of South America during the late C ­ retaceous. S
­ ynthesizing
results from Ecuador and Colombia shows that in those areas the Great Arc was ­involved in
­collisions, first with the Caribbean–Colombian Oceanic Plateau (CCOP) and then with the
­Andean Margin of South America. By 70 Ma a ca 200-km(124.2-mi)-wide Ribbon C ­ ontinent
­consisting of fragments of both the Great Arc of the Caribbean and the CCOP was traveling to
the north in a transpressive plate boundary zone (PBZ) along the Colombian coast.
By 65 Ma, as the CCOP began to enter the Atlantic Ocean and a newly formed Caribbean
plate (CARIB) separated from the Farallon plate, parts of the Ribbon Continent began to be
carried in a southern CARIB transform PBZ eastward along the north coast of South America.
We characterize three W–E-trending belts in that part of the Ribbon Continent: (1) a ­Northern
Belt consisting largely of Great Arc of the Caribbean intrusions and subduction complex rocks;
(2) a Central Belt, very well known in Venezuela, consisting of Great Arc of the ­Caribbean
subduction complex rocks; and (3) a fold-and-thrust belt in the Serrania del Interior and Lara
nappes of Venezuela. A receiver function (seismic) study has shown where rocks of the Great
Arc of the Caribbean abut the South American continent along an E–W-trending line in Ven-
ezuela. We find that rocks of both the subduction complex of the Great Arc and rocks of the
Serrania del Interior have been thrust across that boundary in secondary thrusts as the Ribbon
­Continent has propagated to the east in the South Caribbean transform PBZ.

Copyright ©2015 by The American Association of Petroleum Geologists.


DOI:10.1306/13531931M108846

39

13880_ch02_ptg01_039-084.indd 39 10/27/15 10:03 AM


40  ALTAMIRA and Burke

The structure of the north coast of South America is being radically altered by the northward
movement of the Maracaibo Block as it escapes from deformation related to the collision of the
Panama Arc with Colombia. Restoration of movements within that block during the past 15 My
has been essential to reconstruct the structure and history of the Ribbon Continent on the north
coast of South America.

Introduction: Ribbon Continents Arc of the Caribbean (Burke, 1988) collided with the
west coast of South America in Ecuador. The cessation
The idea that continents have been assembled from of South America arc system’s igneous and high pres-
volcanic arc systems goes back to soon after the estab- sure–low temperature (HP–LT) metamorphic activ-
lishment of the plate tectonic theory (e.g., Green and ity, dated within that southern part of the Great Arc
Ringwood, 1968; McKenzie and Weiss, 1975). For many at ca 70 Ma, can be recognized in many of the parts
years interpretations of the evolution of c­ ontinents of the Ribbon Continent that have subsequently been
applied Wilson cycle models that pictured their redistributed, first along the Colombian west coast of
­assembly from relatively large numbers of island arcs South America and, since ca 65 Ma, along the north-
and the further addition of island arcs to the assembled ern coast of the continent. The full significance of
continents as well as the episodic breakup and a­ ssembly the cutoff in high-temperature igneous and meta-
of continents and even, occasionally, the assembly morphic activity at 70 Ma in Venezuela and islands
of supercontinents comparable to Pangea that con- off the north coast of South America would not have
tained most of the then continental mass of the planet. been so clear without the findings of workers from
Advances in the past 25 years have come in parsimony. Eidgenössische Technische Hochschule Zurich
First, it is clear that pre-Pangean continents, such as (ETH) (e.g., Luzieux et al., 2006; Vallejo et al., 2006)
Rodinia, did not contain all, or even nearly all, of the that first established the time of initial assembly of the
then existing continental mass and were not supercon- Ribbon ­Continent in Ecuador to have been ca 70 Ma.
tinents. Rodinia, for example, cannot be shown to have
been larger in area than Eurasia and Africa are in sum
today (ca 90 3 106 km2), so that Pangea remains the The Ribbon Continent on the West Coast of
only supercontinent (> 150 3 106 km2 in area) for which South America
there is geological evidence (Burke, 2007). Second, it
is unnecessary to postulate the involvement of large Assembly of the Ribbon Continent in Ecuador during
numbers of ancient arcs in the post-Archean assembly Late Cretaceous Time
of continents (Burke, 2007). A small number of large arc
systems that have become highly deformed during the Major components of the Ribbon Continent of the
assembly process better explain the data. For example, ­C ordillera of northwestern South America were
the Altaid assembly of ca 5 × 106 km2 of Asia between erupted, deposited, deformed, assembled, accreted to
ca 700 Ma and ca 250 Ma is best explained by processes the Andean Margin of South America, and began to be
dominantly involving only two arc systems (e.g., redistributed northward along the coast of the conti-
­Sengor and Natalin, 1996). Perhaps because of renewed nent by strike-slip movements during a ca 20 My inter-
interest in the Lomonosov ridge, which is a type of rib- val between ca 90 Ma (late Turonian times) and ca 70
bon continent, large and complex arc systems of Altaid Ma (mid-Campanian times). Substantial fragments of
type are now being recognized as ribbon continents this material continue to occupy the Andean forearc of
in such mountain belts as the western Cordilleras of Ecuador from the Gulf of Guayaquil to the ­Colombian
North and South America (Centeno-Garcia et al., 2008; border (Figure 1). Improved understanding of the his-
Johnston, 2008; Hildebrand, 2009, 2013; Hildebrand and tory of this remarkable temporally concentrated epi-
Whalen, 2014). sode of continental evolution has come from recent
Here we identify and briefly describe the history field, sedimentological, petrological, geochemical,
of a ribbon continent, ca 0.5 M km2 in area, that has paleomagnetic, and isotopic age studies (e.g., Luzieux
been added to the northwestern Cordillera of South et al., 2006; Vallejo et al., 2006; Villagomez and Spikings,
America within the past 75 My. Isotopic ages (Table 2013). Here we mainly use these results in summarizing
1 and Altamira et al., in press) demonstrate the unity the geology of the faulted blocks that were assembled
of the Ribbon Continent and show that its formation in Ecuador to contribute to the newly formed Ribbon
dates from a time when the southern part of the Great Continent. Outcrops in Ecuador preserve blocks of the

13880_ch02_ptg01_039-084.indd 40 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  41

Figure 1. Ribbon Continent in Ecuador and


­ olombia. The Ribbon Continent on the west coast
C
of South America consists of fragments of (1) the
Caribbean–Colombian Oceanic Plateau (CCOP)
and (2) the Great Arc of the Caribbean. In Ecuador,
where the record of the sequential collision of the
Great Arc is best preserved, first with the CCOP
and then with the then active Andean Arc of
South America, some Great Arc blocks overlie the
CCOP and others are caught up in the Pallatanga
collisional suture and fault zone of the Cordil-
lera. Farther north in Colombia, the intensity of
transpressional deformation in the 200-km
(124.2-mi)-wide Ribbon Continent plate boundary
zone (PBZ), which separates the main body of the
CCOP from the continent of South America, has
narrowed the outcrop of the Great Arc to < 100
km (< 62.1 mi) in the Quebradagrande and Arquia
subduction complexes, which are part of the ca
600-km (372.8-mi)-long Romeral suture and fault
zone. Polymetamorphic Complex fragments within
those zones at (1) and (2) are parts of the base-
ment of the then active Andean volcanic arc that
have been incorporated by strike-slip movement
into the Romeral suture and fault zone. Blue schist
and eclogite fragments (*) are the southernmost
of many occurrences of rocks in those facies that
are sporadically distributed in Great Arc subduction
complex outcrops. The Antioquia batholith of Juras-
sic and Cretaceous age within the active Andean
Arc of the late Cretaceous is among the larger
objects on the Earth that indicates a sense of strain
accumulation. Geology at the end of Cretaceous
times is shown on present geography (interpreta-
tion based mainly on work by Luzieux et al., 2006;
Nivia et al., 2006). 100 km (62.1 mi)

Ribbon Continent at or close to the site of their original The Caribbean–Colombian Oceanic Plateau
assembly. The particular importance of the Ecuadorian in Ecuador
blocks is therefore that they have traveled only a short The Caribbean–Colombian Oceanic Plateau (CCOP) is
distance with respect to South America during the time the large igneous province (LIP) that formed when a
since the Great Arc of the Caribbean collided with the mantle plume reached the surface and established the
continent. By contrast, only far-traveled fragments of parent body of the Galapagos Islands hot spot (Figure 2;
the Ribbon Continent are exposed in Colombia and Kerr et al., 1997). Figures 2, 5, 6, and 7 are sketches, but
Venezuela and in the offshore islands of the southern like situations for similar Cretaceous times have also
Caribbean that are distributed along the northern coast been depicted on more formal plate reconstructions
of South America as far to the east as Tobago. Blocks of (e.g., Pindell and Kennan, 2001; ­Pindell et al., 2005). We
the Ribbon Continent in those regions have been car- did not attempt such reconstructions as (1) they tend to
ried in strike-slip motion, mainly on transform faults, obscure uncertainties about regional geology because
first along the western and then along the northern they must close vector triangles, (2) they contain too
coast of South America for distances up to 3000 km many formal uncertainties including multi-plate cir-
(1864.1 mi) (Figure 1). We first summarize the charac- cuits with propagating errors about which there is no
ter of the faulted blocks in Ecuador and explain how consensus, and (3) they contain assumptions about hot
they attained their present distribution as a result of the spot population movements in various reference frames
sequence of events that culminated in the departure of that are also unresolved. Unfortunately, plate rotations
the fragments that now occur as fault-bounded blocks involving the CARIB become robust only for times after
far away to the north and east. the spreading center in the Cayman trough formed at

13880_ch02_ptg01_039-084.indd 41 10/27/15 10:03 AM


42  ALTAMIRA and Burke

Figure 3. Sketch map showing that the Galapagos Islands


today vertically overlie the edge of the Pacific large low
shear-wave velocity province (LLSVP) (Garnero et al., 2007)
at the core–mantle boundary. Hot spot volcanoes with
that property have been shown to have risen from the
core–mantle boundary (Torsvik et al., 2006). South America
is shown in its approximate position with respect to the
LLSVP and the Caribbean–Colombian Oceanic Plateau,
which was the parent large igneous province of ­today’s
Galapagos archipelago at ca 90 Ma, the time of initial
­eruption of the LIP.
Figure 2. Sketch map and cross section (based on Luzieux
et al., 2006, figure 9, and Rogers et al., 2007, figure 9)
showing the relations of the Great Arc of the Caribbean and their parent LIPs by tracks. Only the youngest part
the Caribbean–Colombian Oceanic Plateau (CCOP) to the of the track is preserved in the case of the Galapagos
Americas before the arc collided with the plateau and the hot spot because older parts of the track have been
polarity of subduction reversed. 10 km (6.2 mi) subducted.
Fragments of the CCOP form parts of eight fault-
bounded blocks of the Andean forearc in the coastal
ca 50 Ma (Leroy et al., 2000). From that time on, the lowlands and the western Cordillera of Ecuador. In
CARIB has moved close to due east with respect to both the coastal lowlands, the CCOP forms the basements
North and South America. of the Pinon, San Lorenzo, Pedernales–Esmeraldas,
Other names have been used for the CCOP, but and Santa Elena Blocks (Figure 1). In the western
we use the name that emphasizes its most extensive ­Cordillera, the CCOP is represented by the basements
areas of preservation in the Caribbean and Colombia. of the Rio Desgracia, Naranjal, and Pallatanga Blocks.
The Galapagos archipelago lies above the 1% slow Basement is not exposed in the westernmost Cordil-
shear-wave velocity contour that occupies the edge of leran block, the Macuchi Block.
the Pacific large low shear-wave velocity ­p rovince The age of eruption of the CCOP has been deter-
(LLSVP) (Garnero et al., 2007) on the core–mantle mined from basement gabbros of the Pallatanga
boundary (core–mantle boundary., Figure 3). That Block that yielded a U/Pb zircon age of 87.1 1 0.83
observation is significant because Torsvik et al. (2006) Ma ­(Luzieux et al., 2006). A hornblende 40Ar/39Ar pla-
and Torsvik and Burke (2014) demonstrated that teau age of 88.8 1 1.6 Ma from a gabbro within the
LIPs and hot spots derived from deep-seated mantle basement of the Pinon Block (Pinon Formation) is
plumes overlie those 1% slow contours at the edges not significantly different from the U/Pb zircon age
of LLSVPs on the CMB. Hot spot volcanoes and LIPs (Luzieux et al., 2006). Both ages are within the nar-
derived from the CMB in those loci make up a dis- row range of ages for the CCOP that have recently
tinct population among non-plate-margin igneous been published from widely distributed localities
rocks. The conclusion that the Galapagos Islands (e.g., Sinton et al., 1998; Hoernle et al., 2002; Vallejo
and their parent LIP (CCOP) are directly linked to et al., 2006). At least partly pelagic sedimentary rocks
the CMB is a robust one (Burke and Cannon, 2014). (black shales, calcareous lutites, and radiolarites) of
Hot spots with a CMB origin, such as the Tristan hot the Calentura F ­ ormation, which overlies the CCOP
spot in the South Atlantic, are commonly linked to in the Pinon Block, yielded fossils of Coniacian age

13880_ch02_ptg01_039-084.indd 42 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  43

(89–86 Ma) (Luzieux et al., 2006), confirming a finding


of Velasco-Sanchez and Mendoza (2003). They relate
to sedimentary rocks intercalated among volcaniclas-
tic sandstones in the upper parts of the Calentura
­F ormation. We consider that those sandstones are
derived not from the CCOP but from the Great Arc of
the Caribbean and postdate the collision of the Great
Arc with the CCOP.

Great Arc of the Caribbean in Ecuador


The Great Arc of the Caribbean, parts of which out-
crop in Mexico, Nicaragua, the Greater Antilles, and
the Lesser Antilles, on the sea bed of the Aves Swell, on
islands offshore of Venezuela (from Aruba to Tobago),
and on the continent of South America in Venezuela,
Colombia, and Ecuador, is the older of two island arcs
that presently occupy parts of the margin of the CCOP
in the Caribbean (Burke, 1988). Parts of the Great Arc
are now in contact with the CCOP on the western
flank of the Saba Bank and the Aves Swell (Figure 4).
On the northern and southern sides of the Caribbean,
the Great Arc is also in contact with the CCOP in the
North and South Caribbean plate boundary zones
(PBZs), but those contacts are strike-slip dominated
and the original relationship of the Great Arc to the
CCOP is not so readily discernable. Volcanic arc-type
igneous rocks in the Great Arc have yielded isotopic
ages of >100 Ma (e.g., in Tobago) that are older than
the eruption age of the CCOP (ca 90 Ma). For that rea-
son the relationship between the arc and the contigu- Figure 4. Sketch map and E–W crustal cross section
ous CCOP has been interpreted to have been the result (at 15° N) from Donnelly (1994) showing the Aves Swell
of collision (Burke, 1988). An alternative idea, that in contact with the Caribbean–Colombian Oceanic Plateau
the Great Arc never collided with the CCOP, requires (CCOP). The Aves Swell was the active part of what is
the contiguity of the Great Arc with the CCOP west now the Lesser Antillean segment of the Great Arc of the
of the Aves Swell and west of the Saba Bank to be a ­Caribbean until ca 60 Ma. The relationship shown ­requires
the Great Arc, which is older than the CCOP, to have
coincidence (Pindell et al., 2005; Hildebrand, 2013
­collided with the CCOP after that LIP erupted at ca 90 Ma.
for summaries of published reasons for adopting the Evidence in Ecuador places the timing of that collision
alternative idea). In that case the eruption of the CCOP at ca 86 Ma. 500 km (310.7 mi)
plume head fortuitously juxtaposed the plateau and
the arc. Island arc volcanic rocks on top of the CCOP
in Ecuador show that the Great Arc did indeed collide flows, columnar basalts, hyaloclastites, and pillow
with the CCOP and provide the best currently availa- lavas intercalated with sedimentary rocks of the San
ble information about the timing of that collision at ca Lorenzo Formation overlie the CCOP Pinon Forma-
86–80 Ma (Luzieux et al., 2006; Van Melle et al., 2008). tion basement in the San Lorenzo Block. Because San
The Great Arc is represented in the western Cordil- Lorenzo Formation volcanic rocks show subduction-
lera of Ecuador by volcanic and volcaniclastic sedi- related geochemical affinities (Goossens and Rose,
mentary rocks derived from an arc built on the ocean 1973; Reynaud et al., 1999) and contain no continen-
floor, which form much of the Macuchi, Rio Desgra- tal detritus, we interpret the San Lorenzo Formation
cia, Naranjal, and Pallatanga Blocks (Luzieux et al., rocks to represent the Great Arc of the Caribbean and
2006) (Figure 1). The relationship of the Great Arc to to record the collision of the Great Arc with the CCOP
the CCOP in the Ecuadorian forearc is also displayed as having taken place within the Pacific Ocean at, or
in the Pinon, San Lorenzo, Santa Elena, and Peder- close to, Santonian time (ca 86–84 Ma).
nales–Esmeraldas Blocks of the coastal lowlands. For The Pinon, San Lorenzo, Santa Elena, and
example, volcanic arc igneous rocks including basaltic ­Pedernales–Esmeraldas Blocks of the coastal lowlands

13880_ch02_ptg01_039-084.indd 43 10/27/15 12:38 PM


44  ALTAMIRA and Burke

Figure 5. Sketch map and cross section (based on Luzieux Figure 6. Sketch map and cross section (based on Luzieux
et al., 2006, figure 9, and Rogers et al., 2007, figure 9) et al., 2006, figure 9, and Rogers et al., 2007, figure 9)
showing the collision of the Great Arc of the Caribbean with showing how subduction polarity was reversed imme-
the Caribbean–Colombian Oceanic Plateau (CCOP) at diately after the collision between the Great Arc and the
ca 86 Ma. Subduction of the Farallon plate to the northeast ­Caribbean–Colombian Oceanic Plateau (CCOP)
­immediately ceased. South limit of the ­Chortis Block (see Figure 5). Ocean floor generated at Atlantic Ocean
(gray and dashed areas) formed the suture with the Great spreading ­centers immediately began to be subducted.
Arc of the Caribbean after 75–70 Ma ­collision (Rogers et al.,
2007) (see Figure 7a). 10 km (6.2 mi)

of Ecuador are all in strike-slip contact with fragments at ca 75 Ma (Jaillard et al., 2004; Luzieux et al., 2006,
of the CCOP that together with Great Arc fragments figure 3). Paleomagnetic data (Luzieux et al., 2006,
occupy a transform PBZ that separated the main figure 8) show that the collision of the Great Arc with
CCOP body from the late Cretaceous Andean Margin the Andean Margin of South America can also be
of South America (Figure 1). Great Arc lithounits in the dated by the occurrence of clockwise rotations of dis-
coastal lowland CCOP blocks, such as the San ­Lorenzo crete fragments of the Pinon and San Lorenzo Blocks
Formation, may represent a part of the Great Arc that between 75 Ma and 70 Ma (late Campanian to early
collided with the southern margin of the CCOP. Maastrichtian times). From roughly 75 Ma, the Great
On the basis of the evidence on the San Lorenzo Arc and the associated CCOP moved ENE toward the
Block, the Santonian appears to have been the time gap between the ­A mericas and the Atlantic Ocean
of collision and immediate reversal of the subduction beyond. With the establishment of the Panama Arc
polarity of the Great Arc (Burke 1988) (Figures 5 and on the southwestern flank of the CCOP (presumably
6). After that reversal of polarity, the Great Arc, trave- ca 70 Ma), the Caribbean became isolated from the
ling to the ENE and carrying the CCOP behind it, con- Farallon plate and entered the Atlantic Ocean as a
verged obliquely with the Andean Margin of South newly independent CARIB.
America. Collision of the arc with the already-existing
Andean Arc of Ecuador while traveling to the ENE is The Pallatanga Suture and Fault Zone
dated by the earliest occurrence of continent-derived The Great Arc and the attached CCOP were sutured
detritus in the Pallatanga Formation within the to the Andean Margin of South America as a result
Pallatanga Block (Figure 1) during late Campanian of collision by ca 70 Ma (Burke, 1988; Luzieux et al.,

13880_ch02_ptg01_039-084.indd 44 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  45

2006; Vallejo et al., 2006). The structure of that suture CCOP (Figure 5). That track joined the CCOP to the
zone is relatively obscure because of the following then position of the Galapagos hot spot (Figures 2 and
reasons: 5). Much of the track has now been subducted.
1. The entire 100- to 200-km (62.2- to 124.2-mi)- 3. Collision of the CCOP with the Great Arc of
wide sutured body, made up of the CCOP and Great the Caribbean in late Coniacian or Santonian times
Arc rocks that had just been assembled to initiate the (ca 86–84 Ma) (Burke, 1988) is recorded in the San
establishment of the Ribbon Continent, immediately ­L orenzo Block of coastal Ecuador and in several
became involved in a ca 200-km (124.2-mi)-wide, gen- other blocks in the Ecuadorian forearc. After the colli-
erally transpressional, transform PBZ and began to sion the polarity of the Great Arc’s subduction direc-
move northward. For that reason, workers in Ecuador tion immediately reversed, and the combined CCOP
refer to the structure as the ­Pallatanga fault. We refer and Great Arc began to move in a generally north-
to it as “the Pallatanga suture and fault zone” (PSFZ east direction with respect to North America (Burke,
in Figure 1) to emphasize that it has the characters of 1988). Farther south, on the coast of Peru, no strong
both types of structures. evidence of collision of the Great Arc, or any other
2. Today’s active Andean volcanic arc in Ecuador buoyant object with the Andean Margin of South
overlies the Pallatanga fault. The active arc is crested America during the ­C retaceous, has been recog-
by the Inter-Andean depression, which is an arc- nized (Hildebrand and Whalen, 2014, for another
crestal rift having similar gravitational collapse origin view). That observation is consistent with the idea
to an arc-crestal rift from Peru described by ­Dalmayrac that the Great Arc and CCOP at the time of their col-
and Molnar (1981). lision with the Andean Margin of South America
3. An active strike-slip fault occupies the site of extended no farther to the south than the Gulf of
the Pallatanga fault on the eastern side of the Inter- Guayaquil (Burke, 1988) (­Figure 1).
Andean depression in much of Ecuadorian Andes. 4. The northeastward motion (with respect to North
That fault extends north-northeastward from the Gulf America) of the combined CCOP and Great Arc during
of Guayaquil to the southernmost point of the triangu- the late Cretaceous (ca 84–70 Ma) had different conse-
lar Maracaibo Block at Bucaramanga, Colombia. quences in different parts of the Americas ­(Figure 7).
On the southwestern margin of North America in
History of the Assembly of the Ribbon Continent Mexico, the collision of the Great Arc began in the
in Ecuador west as early as 75 Ma (e.g., Centeno-Garcia et al.,
The descriptions of the geology of the CCOP, the 2008; Ratschbacher et al., 2009) and progressed east-
Great Arc of the Caribbean, and the Pallatanga suture ward reaching the southern margin of the Yucatan by
and fault zone in the foregoing sections enable us to ca 65 Ma. The parts of the Great Arc now preserved in
present a history of the sequence of events involved the Greater and Lesser Antilles moved toward a space
in the assembly of the Ribbon Continent in Ecuador: between the Americas that was increasing in width in
1. The CCOP erupted at ca 90 Ma (late Turonian its north to south dimension between ca 84 Ma and
time) (Luzieux et al., 2006), perhaps in the present ca 68 Ma. The southeastern and southern parts of the
latitude and longitude of the Galapagos (0° N 90° W) Great Arc, which had collided with the Andean coast
(Vallejo et al., 2006). The plateau was centered on the of South America in Ecuador and Colombia at ca 70 Ma
present site of the Galapagos (Burke and Cannon, (Luzieux et al., 2006), began immediately after that
2014) (Figure 3). It may have had an approximately cir- collision, to move to the north-northeast in a transpres-
cular shape and have been 1000 to 2000 km (621.4 to sional transform PBZ (Figure 7B and C; Table 2).
1242.8 mi) in diameter (Figure 3). The E–W trend of the
­Pinon- and Santa Elena–collided volcanic arc blocks on
the north side of the Gulf of Guayaquil (Luzieux et al., Transport of Fragments of the CCOP and the
2006) (Figure 1) perhaps indicates the shape of the south- Great Arc Northward from Ecuador
ern border of the CCOP at the time of that collision.
2. Deposition of pelagic sediments of the Calentura A ca 150-km (93.2-mi)-wide block of the CCOP lies
Formation on top of the CCOP during Coniacian times beneath the coastal lowlands of the Ecuadorian forearc
(ca 89–86 Ma) (Luzieux et al., 2006) marks the inter- ­(Figure  1). The southern and western coasts of the
val when the CCOP was an isolated oceanic plateau ­lowlands contain fragments of the Great Arc in the Pinon,
within the Farallon plate and was moving along with Santa Elena, San Lorenzo, and Pedernales–Esmeraldas
that plate. As the plate moved, a hot spot track began Blocks that were juxtaposed with the 150-km (93.2-mi)-
to form on the Farallon plate to the southwest of the wide CCOP Block during the late Cretaceous (75–65 Ma)

13880_ch02_ptg01_039-084.indd 45 10/27/15 10:03 AM


46  ALTAMIRA and Burke

in major, mainly strike-slip, faults, including the Puerto


Cayo fault and the Canande fault (Luzieux et al., 2006)
(Figures 1 and 7B and C). The collision of the Great Arc
with the CCOP has been characterized by ­Luzieux et al.
(2006) and Vallejo et al. (2006) as involving the CCOP, not
only in the Pallatanga Formation but also in the San Juan
unit. This unit is in faulted contact with the Pallatanga
Block and also contains fault juxtaposed fragments of the
Great Arc; e.g., at Peltetec (1.5° S, 78.5° W), rocks of island
arc affinity have yielded early Cretaceous ages. A gra-
nitic block in the Pujili mélange of the Pallatanga Block
yielded a U/Pb age of 85 1 1.4 (2 sigma) Ma (Vallejo et
al., 2006). That block has the geochemical character of an
island arc built on ocean floor (Vallejo et al., 2006). From
its age, the Pujili granitic block seems likely to have been
generated by subduction of the CCOP after collision with
the Great Arc and before collision of the combined CCOP
and Great Arc with the Andean Margin of South Amer-
ica (see Figure 7 b). Vallejo et al. (2006) pointed out the
compositional and age resemblance of the Pujili mélange
block to the Aruba batholith that White et al. (1999) inter-
preted to have formed in the Great Arc and to have been
related to the subduction of the CCOP. Figure 7 (based on
figure 9 of Luzieux et al., 2006) shows that assembly of
Figure 7. (A) Slab rollback pulling the Great Arc of the Ribbon ­Continent may have extended into Colombia,
the ­Caribbean into the Atlantic Ocean and related but before discussion of what happened in Colombia, we
­transpressional collision of part of the Great Arc and the first outline the characteristics of the rocks of the newly
­Caribbean–Colombian Oceanic Plateau (CCOP) with assembled Ribbon Continent that were carried away to
Northwest South America in Ecuador. (B) Map detail in the north from Ecuador after the collision of the com-
Ecuador and (C) cross section showing structure within bined CCOP and Great Arc with the then active Andean
the ca 200-km (124.2-mi)-wide plate ­boundary zone in Margin of South America. Identification of rocks with
which the newly assembled ­Ribbon Continent, consisting those specific ages and those specific rock types provides
of blocks of the Great Arc and of the CCOP, ­began to move the basis for recognizing the Great Arc and CCOP frag-
to the ­northeast. Rotations in the San ­Lorenzo and Pinon
ments within fault-bounded blocks in western Colombia
Blocks were determined ­paleomagnetically ([A] is based on
and on and off the north coast of South America as far
figure 9 of Rogers et al., 2007 and [B] is based on figure 9 of
Luzieux et al., 2006). to the east as 61° W. Blocks as far away as Tobago can be
recognized to be parts of the Ribbon Continent by means
of those identifications.

Table 2.  Events at ca 65 Ma illustrating the change of


motion of the Caribbean plate from northeast to east as it Rocks and Structures of the Ribbon Continent That
entered the Atlantic. Can Be Expected to Outcrop in Fragments Carried
North from Ecuador along the Western and Northern
65 Ma: Ending of collisional igneous activity of the Great
Arc in Yucatan. Coasts of South America
65 My: Time needed for the fragment carried farthest to the
east along the north coast of SOAM (Tobago) to travel at ca Fragments of the CCOP
20 km/My. Characteristics of LIPs (Ernst and Buchan, 2003)
68 Ma: Change of motion, from separation to convergence, include the following: (1) huge volumes of igneous,
of NOAM with respect to SOAM as estimated using mainly basaltic, rock; (2) eruption over a short inter-
Atlantic Ocean floor magnetic anomalies. val, commonly not much more than 1 My; (3) associ-
65 Ma to 55 Ma: Formation of Yucatan and Grenada Basins ated giant dike swarms; and (4) distinctive rare-earth
in response to shear stresses developed as the Caribbean geochemistry showing that they are not volcanic arc
plate squeezed its way past the Americas and into the rocks or mid-oceanic ridge basalts (MORBs). In rela-
Atlantic Ocean. tively small fragments, such as those that may occur

13880_ch02_ptg01_039-084.indd 46 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  47

in subduction complexes, on the west coast of South compositions. Ages of basalts of MORB composition
America in Colombia, and on and offshore of the north can be expected to be greater than 90 Ma, reflecting
coast of South America, criteria (1) and (2) cannot pre-CCOP eruption ages. Volcanic arc rocks, mainly
be applied. We suggest that criterion (2), the narrow basalts, andesites, and rhyolites as fragments within
range of isotopic ages of a LIP, which for the CCOP is volcaniclastic rocks, as well as intrusions and flows,
close to 89 Ma, is likely to be the most useful distinc- can also be expected to occur in accretionary prism
tive criterion. Not all igneous rocks with isotopic ages units, especially as re-sedimented blocks. Faunas and
close to 89 Ma on the north coast of South America are floras in accretionary prism rocks can be expected to
necessarily fragments of the CCOP, but basalts of that range throughout the entire pre-collisional history
age should be considered as likely candidates. of the Great Arc, that is, from at least as old as mid-­
Jurassic (ca 170 Ma, or earlier) to late Campanian
Fragments of the Great Arc (ca 70 Ma) in age.
Various island arc environments can be expected
to be represented among a variety of transported Fragments of Triassic and Older Intrusives
­fragments. Great Arc plutonic rocks from beneath arc The collision of the Great Arc with the Andean ­Margin
volcanoes will show two diagnostic features: (1) they of South America in Ecuador at 70–75 Ma involved
will yield igneous (i.e., isotopic systems with high clo- the juxtaposition of rocks of two arc systems in the
sure temperatures) ages no younger than ca 70 Ma, ­Pallatanga suture and fault zone (Figure 1). The late
which was when the Great Arc collided with the Cretaceous Andean Margin, now exposed in the east-
Andes in E ­ cuador and, at that time, ceased to erupt ern Cordillera of Ecuador, contained Paleozoic to early
igneous rocks, and (2) they will show isotopic and Cretaceous metasedimentary and intrusive rocks,
trace element compositions characteristic of an island including Triassic migmatites from which Litherland
arc constructed on ocean floor during late Jurassic to et al. (1994) reported U/Pb zircon ages of 227 1 2 Ma.
late Cretaceous times. Vallejo et al. (2006) obtained 40Ar/39Ar ages on micas
Forearc sedimentary rocks, accretionary prism from those migmatites of ca 68.5 Ma, a time when the
rocks, and HP/LT metamorphosed rocks of the sub- Ribbon Continent was traveling to the north in a trans-
duction complex are all likely to be represented form PBZ. The Triassic-aged migmatites crop out close
among transported fragments. HP/LT metamorphic to the Pallatanga suture and fault zone (see Vallejo et
rocks of the Great Arc have not been reported from al., 2006) (Figure 1). We suggest that Triassic plutonic
outcrop in Ecuador where we consider that they lie rocks as well as similar rocks of Paleozoic and Prote-
at depth. Transport by strike-slip movement has long rozoic ages occur as tectonically isolated fragments
been recognized as an effective way of moving rocks within the Great Arc subduction complex and are a
formed at blue schist and eclogite formation depths to minor but significant component of the Ribbon Conti-
the surface, especially where thrusts are also involved nent in Ecuador, Colombia, and Venezuela.
(Karig, 1980). Blue schists that are inferred to be at
depth in Ecuador are likely therefore to outcrop in the
far-traveled blocks of Colombia and Venezuela. Their The Romeral Suture and Fault Zone of Colombia:
HP/LT metamorphic ages in minerals with high clo- A Northern, Along-strike, Extension of the Pallatanga
sure temperatures, like those of the igneous ages of Suture and Fault Zone of Ecuador
Great Arc rocks, can be expected to be no younger
than ca 70 Ma. That would have been the last time Almost 100,000 km2 (62137 mi2) of the CCOP occu-
when those rocks were at blue schist– and eclogite- pies the western Cordillera of Colombia, and a com-
forming depths. parable area, extending at least to the offshore island
Accretionary prism material in transported blocks of Gorgona, underlies lower ground farther to the
in western Colombia and on the north coast of South west and closer to the coast (Nivia et al., 2006).
America is likely to be preserved in green schist and ­Structural ­complexity has been mapped in generally
prehnite–pumpellyite facies more commonly than N–S-­trending faults parallel to the regional strike of
in blue schist facies. Rocks of the accretionary prism the Cordillera and in cross faults striking between NE
can also be expected to include graywackes and and SE (Figure 1) (Nivia et al., 2006; Correa-Martinez,
shales as well as slivers of deep-water cherts and 2007). As in Ecuador, CCOP outcrops in Colombia
carbonate rocks, although carbonate rocks are likely expose mafic (basaltic) and ultramafic rocks as well
to be rare. Igneous rocks in accretionary prism frag- as rocks that have been identified as “plateau vol-
ments are likely to include serpentinized harzburgites canic rocks” (Nivia et al., 2006) and overlying marine
as well as ocean-floor basalts of CCOP and MORB sediments.

13880_ch02_ptg01_039-084.indd 47 10/27/15 10:03 AM


48  ALTAMIRA and Burke

The Romeral system of Maya and Gonzalez (1996), They correspond to the structure in Ecuador, along
a ca 100-km (62.2-mi)-wide and ca 700-km (435-mi)- strike to the south, that we have called the Pallatanga
long geologically complicated and heavily faulted belt suture and fault zone (Figure 1).
to which Burke (1988) had applied the name Romeral
suture, lies east of the CCOP outcrop and separates it The Quebradagrande Complex
from the outcrop of the Cretaceous Andean Margin of A description of the Quebradagrande Complex by
Colombia in which an Andean Arc had been constructed Nivia et al. (2006) is consistent with the idea that the
on a basement called “The Polymetamorphic Complex complex is mainly composed of relatively shallow
of the Central Cordillera” (Restrepo and Toussaint, level subduction complex accretionary prism rocks
1982). That complex consists of rocks similar to those of the Great Arc of the Caribbean. The “Quebradag-
underlying the Cretaceous Andean Arc of Ecuador. For rande rocks consist of imbricated slices of strongly
consistency with the nomenclature that we use in Ecua- deformed dynamometamorphic rocks with crenula-
dor we here use the term “Romeral suture and fault tion cleavage and Andean mylonitic foliation that
zone” for that narrow complex area. Maya and Gon- bears NNE and dips 50° to 70° to the east” (Nivia
zalez (1996) and Nivia et al. (2006) distinguished two et al., 2006). Mylonites include slices up to 1 km
complexes within the structure here called the Romeral (0.6 mi) thick formed from clay-rich carbonaceous
suture and fault zone. Those two complexes, the Arquia mudstones intercalated with thin beds of limestones
Complex and the Quebradagrande Complex , have and cherts. The complex includes both metasedimen-
been mapped as two parallel ca 500-km (310.7-mi)-long, tary and metavolcanic rocks, the protoliths of the latter
north-trending, fault-bounded units. Those complexes, being basaltic to andesitic lavas, and pyroclastic rocks
each of which average ca 25 km (15.6 mi) wide but which affected by metamorphism of zeolite, prehnite–pumpel-
vary in width along strike, separate the western Cordil- lyite, and green schist facies. Imbricated slices of gab-
lera of Colombia from the central Cordillera. Nivia et bro and ultramafic rocks are closely associated with the
al. (2006) named bounding faults of the Romeral suture Quebradagrande Complex and often show the same
and fault zone (1) the Cauca-Almaguer fault, on the west degree of deformation. The rocks of the complex are
against the CCOP, and (2) the Jeronimo fault on the east further described as being sufficiently deformed that
against the Cretaceous Andean volcanic arc that overlies discrete sedimentary sequences cannot be identified.
the basement (“the Polymetamorphic Complex”). The Siliciclastic sedimentary rocks that do crop out within
Silvia–Pijao fault separates the Arquia Complex in the the complex range from breccias and conglomerates
west from the Quebradgrande Complex. to coarse sandstones with clasts of cobbles and peb-
If the late Cretaceous tectonic history that was estab- bles of both volcanic rocks and cherts. Underwater vol-
lished by Luzieux et al. (2006) and Vallejo et al. (2006) caniclastic sedimentation involving mass movements
for Ecuador can be extended into Colombia, it is neces- appears to have been involved in the formation of the
sary that the Arquia and Quebradagrande Complexes, complex. Fossils within the metasedimentary rocks of
which separate outcrops of the CCOP to the west from the complex include ammonites, gastropods, bivalves,
the late Cretaceous Andean Margin of northern South radiolarians, and residues of plants. The faunas have
America to the east, must contain parts of the Great been suggested by Gonzalez (1980) to have lived in
Arc of the Caribbean. Those parts of the Great Arc epineritic to brackish water conditions, although their
would have collided with the CCOP at ca 84 Ma (late present occurrence is in rocks interpreted to be parts of
Santonian time) and, after a reversal of subduction a turbiditic sequence, and they were possibly depos-
polarity, collided with the Andean Margin of South ited in a deep trench. Ages of Quebradagrande fossils
America at 75–70 Ma. On that interpretation the rocks range from Valanginian to Albian (140–100 Ma, Nivia
of the Arquia and Quebradagrande Complexes (1) et al., 2006). In summary, the rocks, structures, faunas,
mark the site of suturing of the Great Arc and CCOP and metamorphic grade of the Quebradagrande Com-
to South America, as well as (2) occupy the fault zone plex appear compatible with the idea that the complex
along which the sutured block began to be displaced mainly consists of fragments of the accretionary prism
to the north after ca 75 Ma, when rocks of the collided or subduction complex of the Great Arc. They satisfy
Great Arc and the CCOP began to travel together to the the criteria defined in an earlier section.
NNE. The Silvia–Pijao fault separates the Arquia Com-
plex from the Quebradagrande ­Complex; to the west, The Arquia Complex: Green Schist and
the Cauca-Almager fault places these two complexes Amphibolite Facies Rocks, Blueschists, Eclogites,
against the CCOP, and on their east the San Jeronimo and Ophiolitic Slivers within the Romeral Suture
fault separates them from the Cretaceous Andean arc and Fault Zone
basement. These system of faults and units together The Arquia Complex contains metamorphosed mafic
form the Romeral suture and fault zone of this chapter. and ultramafic rocks as well as metasedimentary rocks

13880_ch02_ptg01_039-084.indd 48 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  49

in green schist and amphibolite facies (Nivia et al., inadequate recognition of the importance of major
2006; Correa-Martinez, 2007). Part of the complex strike-slip movement within the Romeral suture and
consists of a ca 200-km (124.2-mi)-long discontinuous fault zone during the late Cretaceous interval while
belt of metamorphic rocks in blue schist and eclogite the Ribbon Continent was moving to the north in a
facies cropping out ca 5–10 km (3.1 to 6.2 mi) east of the transpressional PBZ with respect to South America
Cauca-Almaguer fault (Figure 1). Feininger (1980) and (Figure 1). In that interpretation all the occurrences
McCourt and Feininger (1984) suggested that rocks of of old rocks within the two complexes of the zone are
the Arquia ­Complex represent parts of a late Jurassic considered to result from tectonic juxtaposition related
to late ­Cretaceous subduction complex and mélange to northward movement of blocks of the Ribbon
zone with tectonic inclusions of Paleozoic metamor- ­Continent by strike-slip motion during late Cretaceous
phic rocks (Orrego et al., 1980). We suggest that those times (between ca 75 Ma and ca 65 Ma).
tectonic inclusions came from the Polymetamor-
phic Complex of the central Cordilleran basement of
Colombia and were mainly incorporated into the com- Northern Limit of Exposure of North-trending Blocks
plex by strike-slip movements between 75 Ma and 65
of the Ribbon Continent in Western Colombia and the
Ma. Arquia Complex lithounits have been suggested
Change to the E–W Trend Seen in the Blocks of the
by several authors to be formed of lower Cretaceous
rocks (e.g., Toussaint and Restrepo, 1989; Gonzalez North Coast of South America
and Nunez, 1991; Restrepo et al., 1991; ­Gonzalez, 1993,
The northernmost outcrop of the CCOP in northwest-
2001). That, taken with the nature of the contained lith-
ern Colombia is at Cerro Matoso (7° 54' N, 75° 33' W)
ologies, is also compatible with the idea that the Arquia
and that of an ophiolitic sliver in the Romeral suture
Complex represents part of the accretionary prism or
and fault zone is about 100 km (62.2 mi) to the south-
subduction complex of the Great Arc of the Caribbean.
east of Cerro Matoso at Ure (7° 50' N, 75° 14' W) in the
Nevertheless there remains a problem because oth-
Rio Cauca Valley (Correa-Martinez, 2007). These two
ers have suggested that the Arquia Complex is of Pale-
occurrences, which are separated by a poorly localized
ozoic age (e.g., McCourt and Aspden, 1983; McCourt
northern extension of the Cauca-Almaguer fault, are
et al., 1984; Aspden et al., 1987). The occurrence of
the northernmost exposures of the Ribbon Continent
Triassic granitoid plutons “intruding schists west of
in western Colombia (Figure 8). The westernmost out-
the Quebradagrande Complex” (Nivia et al., 2006) as
crops of the ca 1500-km (932.1 mi)-long string of blocks
shown on Nivia et al. (2006, figure 1 therein) is not con-
of the Ribbon Continent on the north coast of South
sistent with a Cretaceous age for the complex. N ­ either
America lie nearly 500 km (310.7 mi) to the northeast
is the occurrence of Paleozoic rocks west of the Que-
of those two localities in the Guajira Peninsula of the
bradagrande Complex (e.g., Mosquera, 1978; Calle
north coast of Colombia. Between La Guajira and
et al., 1980; Mejia et al., 1983a,b; Gonzalez, 2001). For
Cerro Matoso and Ure the Ribbon Continent is not
all those reasons Nivia et al. (2006, p. 432) recognized
only unexposed but may have been largely squeezed
“strong arguments” favoring a Neoproterozoic age for
out as the CCOP and the Lesser Antillean part of the
the Arquia Complex. The same authors have also sug-
Great Arc began to enter the Atlantic Ocean (Figure 8).
gested, mainly on the basis of major and trace element
The area has also been modified since ca 15 Ma by
geochemistry, that the Quebradagrande Complex rep-
the collision of the Panama Arc with the west coast of
resents a “Lower Cretaceous ensialic marginal basin
Colombia (Burke, 1988).
formed within the central Cordillera of the Colombian
Andes” (Nivia et al., 2006).
A paradoxical situation thus exists: on the basis of
(1) location between the CCOP and the Cretaceous The Ribbon Continent on the North Coast
Andean Margin of South America, (2) along-strike of South America
continuity with the structure in Ecuador, and (3) the
nature of contained rock types and structures, the Introduction
Arquia and Quebradagrande complexes of Colombia Bucher (1952) pointed out that the geology of the north
both appear to be mainly parts of subduction com- coast of Venezuela indicates that it is occupied by a
plexes of the Great Arc of the Caribbean. However, major right-lateral strike-slip fault system. That system
very different alternative interpretations currently was recognized to be a transform fault system when,
suggest that the Arquia Complex is of Proterozoic one year after he had introduced plate tectonic theory,
age and that the Quebradagrande Complex marks Wilson (1966) showed that the CARIB had entered the
the site of an “ensialic marginal basin” (Nivia et al., Atlantic Ocean from the Pacific Ocean. Wilson’s model
2006). We attribute those alternative explanations to remained the standard model of the structure of the

13880_ch02_ptg01_039-084.indd 49 10/27/15 10:03 AM


50  ALTAMIRA and Burke
13880_ch02_ptg01_039-084.indd 50 10/27/15 10:03 AM
The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  51

Figure 8. The Ribbon Continent of the northwestern South American Cordillera restored to its condition before the collision
of Panama with the west coast of Colombia. The development of the Maracaibo Block, beginning about 15 Ma at the time
of the collision of Panama with Colombia, has greatly modified the structure of the Ribbon Continent on the north coast of
South America. The Santa Marta, Burro Negro, and Bocono strike-slip faults (dotted), which had not begun to move at the
time shown, have been dominant in the evolution of the Maracaibo Block. The La Victoria fault has also been reactivated
within the Maracaibo Block during the past 15 My. On the west coast the major effect of the Panama collision on the Ribbon
Continent has been shortening in a west to east direction. The structure in the northwest corner of South America for an
along-strike distance of ca 400 km (248.5 mi) was destroyed in the Panama collision. The Pallatanga (PA) and Romeral (RO)
faults are major faults on the west coast, which together with the Cuisa (C), Oca (O), San Sebastien (SS), and El Pilar (EP) faults
of the north coast make up a ca 3000-km (1864.1 mi)-long strike-slip system (the PAROCOSSEP system). 50 km (31.1 mi)

CARIB for ca 20 years, with the southern border of the against the oblique convergence model. The oblique
CARIB plate envisaged as being occupied by a com- convergence model predicts Cenozoic volcanic arc
plex E–W-trending, right-­lateral, transform PBZ (e.g., ages to become progressively younger from west to
Vierbuchen, 1984). A minor modification of Wilson’s east along the Venezuelan Margin, while the trans-
model involved the idea that the Great Arc of the Car- form PBZ model predicts (1) only Cretaceous igne-
ibbean had collided with the west coast of South Amer- ous ages along that margin and (2) no progression
ica during late Cretaceous times and that fragments in age of volcanic arc igneous and HP/LT metamor-
of the Great Arc had subsequently been carried north phic rocks from west to east. Previously published
along the west coast before being redistributed within results and Table 1 include only two Cenozoic ages
the E–W-trending transform PBZ on the southern bor- for volcanic arc igneous rocks (for La Blanquilla and
der of the CARIB during Cenozoic times (Burke, 1988, Los Testigos) out of the more than 300 age determina-
figure 6 therein). tions. Those two Cenozoic ages, located on the north-
Speed (1985) introduced a radically different idea ern rim of the CARIB–South America transform PBZ,
when he suggested that a volcanic arc system had col- are related to Aves Swell and Lesser Antilles volcanic
lided obliquely and progressively from west to east arc igneous rocks displaced on E–W-trending faults
with the north margin of South America. Speed’s idea that are visible on bathymetric and gravity field maps
was radical because it called for a convergent plate and are not related to regional oblique convergence
boundary between the CARIB and South America, (­Figure 1). No other evidence of Cenozoic volcanic arc
rather than Wilson’s (1966) transform boundary. Plate igneous, or HP/LT metamorphic activity has emerged
rotations for the time since 49 Ma, when a spread- in recent research, let alone any evidence of progres-
ing center formed in the Cayman trough (Leroy et sion of volcanic arc igneous ages from west to east
al., 2000), effectively rule out the possibility of the along the northern margin of South America in Ven-
existence of such a convergent plate boundary. How- ezuela and its neighboring offshore islands (­ Figure 9
ever, for the part of the Cenozoic before the ­Cayman and Table 1).
spreading center formed (i.e., from ca 65 Ma to 50 Ma), Information from the BOLIVAR project has contrib-
there exists the remote possibility of a convergent uted greatly to help establish the history of the struc-
CARIB against the South American plate boundary, ture of the Ribbon Continent of South America as its
although geological evidence of such a boundary has leading edge migrated eastward along the north coast
not been discerned. In contrast, convergence in the of South America. That has been true, for example, in
northeastern PBZ of CARIB against North America helping to improve the understanding of the Marac-
between ca 65 Ma and 34 Ma is demonstrated by the aibo Block, which is a giant triangular wedge some
occurrence of volcanic arc intrusions from Cuba to the 0.25 3 106 km 2 in area (ca 1% of the area of South
Virgin Islands (Altamira-Areyan, 2009). Several later America) that has been driven to the north, cutting its
versions of Speed’s model involving oblique conver- way across the southern CARIB PBZ, during the past
gence have been published (e.g., Pindell et al., 1988 15 My (Figure 9) (Burke, 1988). About half the area
2005; Ave-Lallemant and Sisson, 2005; Beardsley and of the Ribbon Continent on the north coast of South
Ave-­Lallemant, 2007). America now lies within the Maracaibo Block; for that
The large-scale BOLIVAR collaborative geophysi- reason, we here summarize the structure of the block
cal, geochemical, and geological research program and describe how we have removed its effects, so that
(Ave-Lallemant and Sisson, 2005) afforded an unprec- we can better reconstruct the evolution of the Ribbon
edented opportunity to test the transform PBZ model Continent before 15 Ma.

13880_ch02_ptg01_039-084.indd 51 10/27/15 10:03 AM


52  ALTAMIRA and Burke
13880_ch02_ptg01_039-084.indd 52 10/27/15 10:03 AM
The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  53

Figure 9. Northern margin of South America as it is today showing the Maracaibo Block, the evolution of which has d ­ estroyed
the simple structure of the Ribbon Continent. Isotopic ages (Altamira-Areyan, 2009) are new 40Ar/39Ar determinations on
­amphiboles of the Great Arc of the Caribbean intrusions and subduction complex rocks (in Aruba and on the mainland of
South America). Inset (modified from Sisson et al., 2006) shows the location of samples in the central part of Villa de Cura
Blueschist Belt. All 40Ar/39Ar ages are older than ca 70 Ma, which was the time when the Great Arc collided with South
America in Ecuador. An age of ca 41 Ma from Los Testigos is attributed to strike-slip movement that carried that island along
the northern edge of the south Caribbean plate (CARIB) transform plate boundary zone from a source in the Lesser Antilles
near Grenada, from which island we have obtained an 40Ar/39Ar age of 38 6 1 Ma in ground mass from a basalt. 8 km (5 mi)

Structure and Evolution of the Maracaibo Block Maracaibo Block wedge to the north (Trenkamp et al.,
2002; Colmenares and Zoback, 2003).
Introduction The Maracaibo Block is traditionally treated as
The Maracaibo Block (Figure 8) formed at ca 15 Ma, bounded by the Bocono fault on its east, the Santa
and for that reason its formation has been regarded as Marta fault on its west, and by an unnamed bound-
one of several responses to the collision of the Panama ary, extending to a depth of ca 200 km (124.2 mi), at
Volcanic Arc with the west coast of Colombia, which its base ­(Figure 8). The underlying structure has been
has been interpreted to have begun at about that suggested based on an interpretation of tomographic
time (Wadge and Burke, 1983; Droxler et al., 1998). results to have extended below the base of the litho-
Newer work indicates much older ages for a colli- sphere and to involve the subduction of more than 100
sion (e.g., Farris et al., 2011; Montes et al., 2012; Barat km (62.2 mi) of the CARIB into the convecting mantle
et al., 2014), and there is no reason to doubt those under South America (Van der Hilst and Mann, 1994).
older ages but a problem is that large scale strike slip We, however, regard the tomographic evidence as more
faults cut the rocks of the currently tectonically active consistent with the idea that the Maracaibo Block is an
isthmus. Some of those faults underwater reach to intra-lithospheric structure and consider the subducted
the sea bottom and are active. It is not known how slabs discerned tomographically in a complex volume
recently older rocks among the faulted blocks of the beneath the block to be related, not to the subduction of
Panama isthmus came to be juxtaposed. Paradoxically the CARIB into the convecting mantle (for which we see
worldwide the best dates for arc-continent collisions no evidence), but to the subduction of Cocos plate lith-
are known to come from the neighborhood of the col- osphere from the Pacific Ocean side of South America
lision zone and not from the collision zone itself. The (Gutscher et al., 2000; Colmenares and Zoback, 2003).
Zagros–Arabia and the Timor–Australia collisions are Strike-slip offsets on the Bocono and Santa Marta
two familiar examples (Burke, 1996; Rutherford et al. faults are each about 100 km (62.2 mi), so the block as
2001). The Maracaibo Block is the largest neighboring a whole has been considered to have moved about ca
structure to the Panama isthmus that is actively being 70 km (43.5 mi) to the north. The northern margin of
deformed. We consider that the Maracaibo Block holds the block crops out in a thrust front on the ocean floor
the key to the timing of the collision between Panama at a water depth of ca 4 km (2.5 mi). In that thrust the
and Colombia, which was at ca 15 Ma. The results of Maracaibo Block abuts the main body of the CCOP,
that collision included (1) the spectacular subsequent which presently occupies much of the ocean floor of the
deformation of the Panama Arc into the shape of an S CARIB. The boundary zone on either side of that thrust
lying on its back; (2) the formation of the Atrato suture is marked by thrusts and folds to which the name South
in western Colombia between the N–S-trending south- Caribbean deformed belt (or fold belt) has long been
ern end of the collided Panama arc in the Sierra Baudo applied (Silver et al., 1975). The Maracaibo Block is
and the outcrop of CCOP fragments in northwestern about 750 km (466.1 mi) wide at 12.5° N but only 400
Colombia; and (3) the progressive separation of the km (248.6 mi) wide farther south at ca 9° N (Figure 9).
waters of the Caribbean Sea from those of the Pacific The greater width results mainly from the transfer of
Ocean between ca 7 Ma and ca 3 Ma, culminating as material by strike-slip faulting from the southern part
part of the Panama arc gradually became the Panama to the northern part of the Maracaibo Block as the block
isthmus (Coates et al., 2004). The immediate indicator has moved to the north. Extension within the block has
of continuing movement of the Maracaibo Block has also played a part. Some of that extension has been in
been considered to be shortening, seen in GPS results, two pull-aparts that we identify from the observation of
across the Atrato suture, which separates the Sierra thin crust in receiver function analyses (Niu et al., 2007).
Baudo from the CCOP fragments of western Colom- There is a lot of internal deformation within the
bia, and the consequent escape of the great triangular Maracaibo Block involving normal faults, strike-slip

13880_ch02_ptg01_039-084.indd 53 10/27/15 10:03 AM


13880_ch02_ptg01_039-084.indd 54

Table 1. Northern Ribbon Continent of South America (CARIB–SOAM PBZ) isotopic ages.

54  ALTAMIRA and Burke


Los Monjes and Aruba Islands
Sample Location Rock type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
LM-145 Los Monjes Orthoamphibolite Whole rock 12.346 −70.868 K/Ar 114 12 Santamaria and Schubert
(1974)
LM-145 Los Monjes Orthoamphibolite Amphibole 12.346 −70.868 K/Ar 116 13 Santamaria and Schubert
(1974)
40
ARU04 Aruba Diabasa Hornblende 12.525 −69.932 Ar/39Ar 91.2 6.3 Altamira-Areyan (2009)
40
ARU02 Aruba Tonalite Hornblende 12.455 −69.891 Ar/39Ar 87.9 4.7 Altamira-Areyan (2009)
  Aruba batholith Tonalite Zircon     U/Pb 89 1 Wright and Wyld (2004)
  Aruba batholith Tonalite Apatite     FT 55 6 Wright and Wyld (2004)
AR-123 Aruba Quartz diorite Biotite     K/Ar 67 4 Santamaria and Schubert
(1974)
40
ARU96-131 Aruba Tonalite Biotite     Ar/39Ar (1s) 82.84 0.36 White et al. (1999)
40 39
ARU96-131 Aruba Tonalite Biotite     Ar/ Ar (1s) 83.09 0.4 White et al. (1999)
40 39
ARU96-158 Aruba Diorite Biotite     Ar/ Ar (1s) 84.99 0.35 White et al. (1999)
40
ARU96-123 Aruba Mafic dyke Hornblende     Ar/39Ar (1s) 77.46 2.09 White et al. (1999)
40 39
ARU96-152 Aruba Hornblende–gabbro Hornblende     Ar/ Ar (1s) 79.92 2.76 White et al. (1999)
40 39
BK77-165 Aruba Diorite Hornblende     Ar/ Ar (1s) 82.2 0.51 White et al. (1999)
Biotite (2) Aruba Granodiorite Biotite     K/Ar 71 4 Priem et al. (1966)
Biotite Aruba Granodiorite Biotite     Rb/Sr 74 4 Priem et al. (1966)
Biotite (1) Aruba Granodiorite Biotite     K/Ar 75 4 Priem et al. (1966)
  Aruba Granodiorite       Rb/Sr 85.1 0.5 Priem et al. (1978)
ANT 122 Aruba (Macuarima) Quartz–biotite Biotite 12.496 −69.976 K/Ar 70.7 2.3 Priem et al. (1986)
tonalite
ANT 120 Aruba (Hooiberg) Quartz–hornblende– Biotite 12.515 −69.990 K/Ar 73 2.9 Priem et al. (1986)
biotite tonalite
ARU 1 Aruba, C. Compa Quartz–hornblende– Biotite 12.511 −69.965 K/Ar 74.2 3 Priem et al. (1986)
biotite tonalite
ANT 102 Aruba (Ceru Quartz–hornblende– Biotite 12.600 −70.029 K/Ar 87.6 3.5 Priem et al. (1986)
Muskita) biotite tonalite
ANT 108 Aruba (Ceru Quartz–biotite– Biotite 12.598 −70.025 K/Ar 88 3.5 Priem et al. (1986)
Grandi) tonalite
ANT 136 Aruba (West of Quartz–hornblende Hornblende 12.465 −69.892 K/Ar 83.5 3.3 Priem et al. (1986)
Rincon) tonalite
ANT 113 Aruba (Matividiri) Quartz–hornblende Hornblende 12.552 −69.978 K/Ar 83.7 3.4 Priem et al. (1986)
gabbro
ANT 137 Aruba, C. Blanco Quartz–hornblende Hornblende 12.456 −69.907 K/Ar 85 3.4 Priem et al. (1986)
tonalite
10/27/15 10:03 AM

ANT 114 Aruba (Matividiri) Quartz–hornblende Hornblende 12.552 −69.978 K/Ar 88 3.5 Priem et al. (1986)
gabbro
13880_ch02_ptg01_039-084.indd 55

ANT 119 Aruba (Hooiberg) Hooibergite Hornblende 12.515 −69.990 K/Ar 89.2 3.6 Priem et al. (1986)
ANT 102 Aruba (Ceru Quartz–hornblende– Hornblende 12.600 −70.029 K/Ar 89.5 3.6 Priem et al. (1986)
Muskita) biotite tonalite
ANT 115 Aruba (Matividiri) Quartz–hornblende Hornblende 12.552 −69.978 K/Ar 129 5.2 Priem et al. (1986)
gabbro
ANT 126 Aruba (Dos Playas) Diabasa Whole rock 12.504 −69.919 K/Ar 62 1.9 Priem et al. (1986)
ANT 161 Aruba (Andicouri) Semi-lamprophyric Whole rock 12.540 −69.959 K/Ar 66.2 2.7 Priem et al. (1986)
dikes
ANT 162 Aruba (Andicouri) Semi-lamprophyric Whole rock 12.540 −69.959 K/Ar 68.3 2.7 Priem et al. (1986)
dikes
ANT 159 Aruba (Altovista) Semi-lamprophyric Whole rock 12.575 −70.013 K/Ar 70.1 2.8 Priem et al. (1986)
dikes
ANT 160 Aruba (Altovista) Semi-lamprophyric Whole rock 12.575 −70.013 K/Ar 70.1 2.8 Priem et al. (1986)
dikes
ANT 158 Aruba (Altovista) Semi-lamprophyric Whole rock 12.575 −70.013 K/Ar 71.6 2.9 Priem et al. (1986)
dikes
ANT 114 Aruba (Matividiri) Quartz–hornblende Whole rock 12.552 −69.978 K/Ar 75.9 3 Priem et al. (1986)
gabbro
ANT 123 Aruba (Dos Playas) Diabasa Whole rock 12.504 −69.919 K/Ar 78.1 2.3 Priem et al. (1986)
ANT 113 Aruba (Matividiri) Quartz–hornblende Whole rock 12.552 −69.978 K/Ar 81.3 3.3 Priem et al. (1986)
gabbro

The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  55


ANT 115 Aruba (Matividiri) Quartz–hornblende Whole rock 12.552 −69.978 K/Ar 83 3.3 Priem et al. (1986)
gabbro

Curacao Island
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
  NW Curacao Sills and dikes Whole rock     K/Ar 72 7 Beets et al. (1972)
CU-126 NW Curacao Lava Trachyandesite Hornblende     K/Ar 76 6 Santamaria and Schubert
Formation (1974)
CU-125 NW Curacao Lava Quartz Whole rock     K/Ar 74 5 Santamaria and Schubert
Formation Trachyandesite (1974)
CU-126 NW Curacao Lava Trachyandesite Whole rock     K/Ar 84 6 Santamaria and Schubert
Formation (1974)
CU-127b Curacao (Lava Dolerite Whole rock     K/Ar 118 10 Santamaria and Schubert
Formation) (1974)
CU-127a Curacao (Lava Dolerite Whole rock     K/Ar 126 12 Santamaria and Schubert
Formation) (1974)
40
79KV-9 Curacao (Koraal Diabase Whole rock     Ar/39Ar (1s) 75.8 2 Sinton et al. (1998)
Tabak sill)
40
BK79-262 Curacao (Lava Plagioclase– Whole rock     Ar/39Ar (1s) 88 1.2 Sinton et al. (1998)
Formation— clinopyroxene
upper part) dolerite
40
79BE-73 Curacao (Lava Olivine tholeiitic Whole rock     Ar/39Ar (1s) 89.5 1 Sinton et al. (1998)
Formation—lower pillow basalt
10/27/15 10:03 AM

part)
  Curacao Basalts Whole rock     Re–Os 85.6 8.1 Wadge and MacDonald (1985)
13880_ch02_ptg01_039-084.indd 56

56  ALTAMIRA and Burke


Bonaire Island
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
ANT 32 Bonaire Hornblende tuff Hornblende 12.263 −68.334 K/Ar 73.1 2.2 Priem et al. (1979)
ANT 34 Bonaire Hornblende tuff Hornblende 12.263 −68.334 K/Ar 86.4 2.6 Priem et al. (1979)
ANT 33 Bonaire Hornblende tuff Hornblende 12.263 −68.334 K/Ar 89.4 2.7 Priem et al. (1979)
ANT 35 Bonaire Hornblende tuff Hornblende 12.263 −68.334 K/Ar 100 3.0 Priem et al. (1979)
ANT 31, 47, Bonaire (Wecuwa Dacitic lava Whole rock 12.227 −68.402 Rb/Sr 54   Priem et al. (1979)
48, 49, and 50 porphyries)
ANT 47 Bonaire (Wecuwa Dacitic lava Whole rock 12.227 −68.402 K/Ar 57 1.7 Priem et al. (1979)
porphyries)
ANT 52 Bonaire (Poeneitic K-rich spilite Whole rock 12.263 −68.334 K/Ar 58.3 1.7 Priem et al. (1979)
lava)
ANT 30 Bonaire (Wecuwa Dacitic lava Whole rock 12.227 −68.402 K/Ar 60.5 1.8 Priem et al. (1979)
porphyries)
ANT 74 Bonaire (Sumpina Volcanic breccias Whole rock 12.288 −68.381 K/Ar 62 1.9 Priem et al. (1979)
porphyries)
ANT 49 Bonaire (Wecuwa Dacitic lava Whole rock 12.227 −68.402 K/Ar 64.3 1.9 Priem et al. (1979)
porphyries)
ANT 140 Bonaire Diabase Whole rock 12.241 −68.307 K/Ar 68.5 2.1 Priem et al. (1979)
ANT 51 Bonaire (Poeneitic K-rich spilite Whole rock 12.263 −68.334 K/Ar 70.9 2.1 Priem et al. (1979)
lava)
ANT 68, 69, Bonaire (Sumpina Volcanic breccias Whole rock 12.288 −68.381 Rb/Sr 72   Priem et al. (1979)
70, 71, 72, 73, porphyries)
and 74
ANT 53 Bonaire (Poeneitic K-rich spilite Whole rock 12.263 −68.334 K/Ar 75.5 2.3 Priem et al. (1979)
lava)
ANT 66 Bonaire (Wash. Dacitic rocks from Whole rock 12.257 −68.369 K/Ar 75.9 2.3 Priem et al. (1979)
porphyries) a sill
ANT 71 Bonaire (Sumpina Volcanic breccias Whole rock 12.288 −68.381 K/Ar 76.2 2.3 Priem et al. (1979)
porphyries)
ANT 60 Bonaire (Wash. Dacitic rocks from Whole rock 12.257 −68.369 K/Ar 77.6 2.3 Priem et al. (1979)
porphyries) a sill
ANT 68 Bonaire (Sumpina Volcanic breccias Whole rock 12.288 −68.381 K/Ar 79.4 2.4 Priem et al. (1979)
porphyries)
ANT 63 Bonaire (Wash. Dacitic rocks from Whole rock 12.257 −68.369 K/Ar 80.3 2.4 Priem et al. (1979)
porphyries) a sill
40
BON 94-09 Bonaire Wash Rhyodacite sill K-feldspar     Ar/39Ar (1s) 96 4.0 Thompson et al. (2004)
Formation (N.
Complex)
10/27/15 10:03 AM
13880_ch02_ptg01_039-084.indd 57

Gran Roque Island


Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
GR-129 Gran Roque Pegmatite       K/Ar 66 5 Santamaria and Schubert
(1974)
GR-131 Gran Roque Quartz diorite Biotite     K/Ar 65 3.6 Santamaria and Schubert
(1974)
GR-132 Gran Roque Quartz diorite Hornblende     K/Ar 66 6 Santamaria and Schubert
(1974)
GR-128 Gran Roque Diabase Amphibole     K/Ar 127 15.0 Santamaria and Schubert
(1974)
GR-130 Gran Roque Lamprophyre Amphibole     K/Ar 130 14.0 Santamaria and Schubert
(1974)

Aves Ridge
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
  Aves Ridge Granodiorite Biotite 13°25.12' −63°43.25' K/Ar 65   Santamaria and Schubert
(1974)

The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  57


La Blanquilla Island
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
LB-140 La Blanquilla Trondhjemite Biotite     K/Ar 64 3.5 Santamaria and Schubert
(1974)
LB-141 La Blanquilla Trondhjemite Biotite     K/Ar 62 3.5 Santamaria and Schubert
(1974)
LB-142 La Blanquilla Pegmatite Feldspar     K/Ar 64 3.4 Santamaria and Schubert
(1974)
LB-142 La Blanquilla Pegmatite Biotite     K/Ar 62 3.4 Santamaria and Schubert
(1974)
LB-143 La Blanquilla Trondhjemite Biotite     K/Ar 64 3.5 Santamaria and Schubert
(1974)
LBQ-5 La Blanquilla Granodiorite Zircon     U/Pb 59   Levander et al. (2006)

Los Hermanos Island


Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
LH-133 Los Hermanos Hornblende–gneiss Hornblende       71 6.0 Santamaria and Schubert
(1974)
LH-134 Los Hermanos Hornblende–gneiss Hornblende       70 5.4 Santamaria and Schubert
(1974)
LH-135 Los Hermanos Hornblende–gneiss Hornblende       67 5.1 Santamaria and Schubert
10/27/15 10:03 AM

(1974)
13880_ch02_ptg01_039-084.indd 58

58  ALTAMIRA and Burke


Los Frailes Island
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
LF-144 Los Frailes Diabase Whole rock     K/Ar 66 5.1 Santamaria and Schubert
(1974)

Los Testigos Island

Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
LT-138 Los Testigos (Main Pegmatitic Amphibole     K/Ar 44 5.4 Santamaria and Schubert
Island) metadiabase (1974)
LT-139 Los Testigos (Main Pegmatitic Amphibole     K/Ar 44 5.5 Santamaria and Schubert
Island) metadiabase (1974)
LT-137 Los Testigos (Main Pegmatitic Amphibole     K/Ar 47 6.1 Santamaria and Schubert
Island) metadiabase (1974)
LT-136 Los Testigos (Main Pegmatitic Feldspar     K/Ar 44 4.5 Santamaria and Schubert
Island) metadiabase (1974)
40
LTG5 Los Testigos Diorite Hornblende 11.367 −63.123 Ar/39Ar 41.3 1.8 Altamira-Areyan (2009)

Trinidad
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
TR13 Trinidad (Sans Basalt Whole rock     K/Ar 87 4.4 Wadge and MacDonald
Souci Formation) (1985)

Margarita Island
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
40 39
IM86-85 Margarita (La Biotite-garnet schist Biotite 11.088 −64.008 Ar/ Ar 24.9 0.9 Sisson et al. (2005)
Rinconada Unit)
40
IM86-85 Margarita (La Muscovite schist White mica 11.088 −64.008 Ar/39Ar 68.3 0.2 Sisson et al. (2005)
Rinconada Unit)
40
VM86-12 Margarita (La Metacong Biotite 11.067 −64.003 Ar/39Ar 44.9 0.7 Sisson et al. (2005)
Rinconada Unit)
40
VM86-12 Margarita (La Metacong White mica 11.067 −64.003 Ar/39Ar 62.3 0.2 Sisson et al. (2005)
Rinconada Unit)
40
VM83-21 Margarita (Juan Garnet–amphibolite Amphibole 11.156 −63.897 Ar/39Ar 52.8 0.2 Sisson et al. (2005)
Griego)
40
IM86-27 Margarita (Los Muscovite schist White mica 10.997 −63.856 Ar/39Ar 53.5 0.5 Sisson et al. (2005)
Robles Unit)
40
IM86-93 Margarita (La Amphibolite Amphibole 11.063 −64.004 Ar/39Ar 62.3 1.4 Sisson et al. (2005)
10/27/15 10:03 AM

Rinconada Unit)
40
IM86-229 Margarita (Los Qtz–Fs schist White mica 11.006 −63.875 Ar/39Ar 86 0.3 Sisson et al. (2005)
Robles Unit)
IM86-229 Margarita (Los Qtz–Fs schist White mica 11.006 −63.875 Ar/ Ar 86 0.3 Sisson et al. (2005)
Robles Unit)
13880_ch02_ptg01_039-084.indd 59

40
IM84-14 Margarita (Juan Eclogite Amphibole 11.029 −63.870 Ar/39Ar 92.4 0.5 Sisson et al. (2005)
Griego)
40
Stage 9 Margarita Basaltic to Amphibole     Ar/39Ar 52 - 47   Stockert et al. (1995)
andesitic dikes
Stage 7 Margarita (Juan Granitic augen Whole rock     Rb/Sr 50   Stockert et al. (1995)
Griego Group) gneiss
Stage 7 Margarita Trondjhemite Zircon     FT 53–50   Stockert et al. (1995)
(Matasiete
trondhjemite)
Stage 6 Margarita Greenschist White mica     K/Ar 50–55   Stockert et al. (1995)
Stage 4 Margarita (Juan HP schist Phengite     K-Ar & Ar 90–80   Stockert et al. (1995)
Griego)
Stage 4 Margarita (El Granite Zircon     U/Pb 86   Stockert et al. (1995)
Salado granite)
Stage 2 Margarita Trondjhemite and Zircon     U/Pb 114–105   Stockert et al. (1995)
(Matasiete orthogneiss
trondhjemite)
Stage 1 Margarita (Juan Granitic augen Zircon     U/Pb 315 35 Stockert et al. (1995)
Griego Group) gneiss
MAR 35B Margarita Granitic Phengite     K/Ar 57.1 2.9 Chevalier et al. (1988)
orthogneisses

The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  59


MAR 35A Margarita Granitic Phengite     K/Ar 62.5 3.1 Chevalier et al. (1988)
orthogneisses
MAR 199 Margarita Mylonitic sole Barroisite     K/Ar 79.3 4 Chevalier et al. (1988)
MAR 36 Margarita Eclogites, Paragonite     K/Ar 84.6 4.2 Chevalier et al. (1988)
amphibolites
W 809 Margarita (La Gneiss Chloritized     K/Ar 37 19 Loubet et al. (1985)
Rinconada Group) amphibole
XII-J Margarita (La Gneiss Muscovite     K/Ar 40.1 2.5 Loubet et al. (1985)
Rinconada Group)
XII-J Margarita (La Gneiss Clinozoisite     K/Ar 43 31 Loubet et al. (1985)
Rinconada Group)
W 675 Margarita (La Gneiss Clinozoisite     K/Ar 44.5 3.5 Loubet et al. (1985)
Rinconada Group)
W 675 Margarita (La Gneiss Chloritized     K/Ar 45.5 7.5 Loubet et al. (1985)
Rinconada Group) amphibole
XII-J Margarita (La Gneiss Amphibole     K/Ar 47 4 Loubet et al. (1985)
Rinconada Group)
W 675 Margarita (La Gneiss Muscovite     K/Ar 50.5 15 Loubet et al. (1985)
Rinconada Group)
W 167 Margarita (La Gneiss Muscovite     K/Ar 51 1.5 Loubet et al. (1985)
Rinconada Group)
W 1252 Margarita   Muscovite     K/Ar 54 2 Loubet et al. (1985)
W 167 Margarita (La Gneiss Chlorite     K/Ar 56.5 26 Loubet et al. (1985)
Rinconada Group)
10/27/15 10:03 AM

(continued)
13880_ch02_ptg01_039-084.indd 60

60  ALTAMIRA and Burke


Margarita Island (continued)
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
W 809 Margarita (La Gneiss Clinozoisite     K/Ar 57.5 15 Loubet et al. (1985)
Rinconada Group)
W 158 Margarita (La Gneiss Muscovite     K/Ar 59.5 2 Loubet et al. (1985)
Rinconada Group)
SJ-BI Margarita   Muscovite     K/Ar 67.5 2 Loubet et al. (1985)
W 167 Margarita (La Gneiss Clinozoisite     K/Ar 98 30 Loubet et al. (1985)
Rinconada Group)
IM-122 Margarita Pegmatite Feldspar     K/Ar 32 2 Santamaria and Schubert
(Western) (1974)
IM-121 Margarita Trondjhemite Hornblende     K/Ar 70 6 Santamaria and Schubert
(Western) (1974)
IM-120 Margarita Sodic granite Amphibole     K/Ar 72 6 Santamaria and Schubert
(Matasiete granite) (1974)

Cordillera de la Costa Belt (I)


Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
PP-113 Paraguana Andesite Whole rock     K/Ar 63 4 Santamaria and Schubert
Peninsula (1974)
PP-116 Paraguana Gabbro Whole rock     K/Ar 118 10 Santamaria and Schubert
Peninsula (1974)
PP-114 Paraguana Gabbro Whole rock     K/Ar 120 11 Santamaria and Schubert
Peninsula (1974)
PP-115 Paraguana Dolerite Whole rock     K/Ar 129 14 Santamaria and Schubert
Peninsula (1974)
  Falcon Basaltic bodies       Stratigraphy 28 - 23   Muessig (1984)
  Falcon Basaltic bodies       K/Ar 22.9 0.9 Muessig (1984)
23-M-2X Offshore Paraguana Gneiss 11.671 −69.522 114   Feo-Codecido et al. (1984)
Peninsula
  Gulf of Venezuela Granite Whole rock     K/Ar 138 6.9 Kohn et al. (1984)
  Gulf of La Vela Quartz–sericite Whole rock     K/Ar 83.5   Loubet et al. (1985)
phyllite
  Gulf of La Vela Metagabbro Feldspar     K/Ar. 114   Morgan (1969)
  Gulf of Venezuela Granite Whole rock     K/Ar 304   Feo-Codecido et al. (1984)
91VSn15 Highway to Augen gneiss Apatite 10.362 −68.092 FT 19.8 1.2 Sisson et al. (2005)
Valencia
10/27/15 10:03 AM
13880_ch02_ptg01_039-084.indd 61

91VSn15 Highway to Augen gneiss Zircon 10.362 −68.092 FT 16.3 1.6 Sisson et al. (2005)
Valencia
91VSn19 Highway to Metagranite Apatite 10.327 −68.092 FT 19.8 1.2 Sisson et al. (2005)
Valencia
91VSn19 Highway to Metagranite Zircon 10.327 −68.092 FT 15 1.4 Sisson et al. (2005)
Valencia
40
92VSn19 Highway to Cata Actinolite schist   10.356 −67.616 Ar/39Ar 37 3 Sisson et al. (2005)
40 39
94VSn13 Highway to Schist White mica 10.324 −68.091 Ar/ Ar 28 0.2 Sisson et al. (2005)
Valencia
40
94VSn15 Highway to Schist White mica 10.326 −68.090 Ar/39Ar 34.2 0.2 Sisson et al. (2005)
Valencia
40
94VSn16 Highway to Schist White mica 10.349 −68.109 Ar/39Ar 42.3 0.3 Sisson et al. (2005)
Valencia
40
94VSn21 Pantanemo Schist White mica 10.471 −67.929 Ar/39Ar 42.6 0.3 Sisson et al. (2005)
91VSn19 Highway to Metagranite   10.327 −68.092 U/Pb 501 25 Sisson et al. (2005)
Valencia
91VSn15 Highway to Augen gneiss   10.362 −68.092 U/Pb 512.4 12.9 Sisson et al. (2005)
Valencia
CH-86-1 South of Metagranite   10.558 −67.397 Rb/Sr 1560 83 Sisson et al. (2005)
Chichiriviche
PM86-1 El Avila National Metagranite   10.568 −67.008 Rb/Sr 1560 83 Sisson et al. (2005)

The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  61


Park
PM86-2 El Avila National Metagranite   10.576 −66.998 Rb/Sr 1560 83 Sisson et al. (2005)
Park
SVC-7 Pena de Mora Gneiss Zircon 10.593 −66.927 FT 17.4 2.1 Kohn et al. (1984)
SVC-13 Pena de Mora Gneiss Zircon 10.558 −66.925 FT 17.5 1.7 Kohn et al. (1984)
SVC-11 Pena de Mora Gneiss Zircon 10.533 −66.928 FT 18.4 1.9 Kohn et al. (1984)
SVC-8 Pena de Mora Gneiss Zircon 10.607 −66.924 FT 18.8 2.1 Kohn et al. (1984)
SV-81-4 Pena de Mora Gneiss Zircon 10.566 −66.993 FT 18.9 1.9 Kohn et al. (1984)
SV-81-6 Pena de Mora Gneiss Zircon 10.540 −66.970 FT 23.9 1.7 Kohn et al. (1984)
V-92-24 Cata Trondhjemite dike   10.495 −67.010 U/Pb 511.3 1.8 Sisson et al. (2005)

Cordillera de la Costa Belt (II)


Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
V-94-49 Choroni Granite Apatite 10.356 −67.600 FT 13.6 1.4 Sinton et al. (1998)
V-94-49 Choroni Granite Zircon 10.356 −67.600 FT 14.1 1.7 Sinton et al. (1998)
V-94-49 Choroni Granite   10.306 −67.616 U/Pb 471 23 Sinton et al. (1998)
SVC-19 Choroni Gneiss Apatite 10.348 −67.589 FT 17.5 1.9 Kiser (1994)

(continued)
10/27/15 10:03 AM
13880_ch02_ptg01_039-084.indd 62

62  ALTAMIRA and Burke


Cordillera de la Costa Belt (II) (continued)
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
SVC-20 Choroni Gneiss Zircon 10.343 −67.577 FT 21.9 2.9 Kiser (1994)
SVC-17 Choroni Gneiss Zircon 10.363 −67.586 FT 22.3 2.3 Kiser (1994)
SVC-15 Choroni Gneiss Zircon 10.362 −67.568 FT 24.1 3 Kiser (1994)
  Choroni Metagranite Biotite 10.357 −67.648 K/Ar 30 1.9 Santamaria and Schubert
(1974)
  Choroni Metagranite Biotite 10.357 −67.648 K/Ar 30 1.8 Santamaria and Schubert
(1974)
  Cantagallo Gabbro Whole rock     K/Ar 65 5 Santamaria and Schubert
(1974)
  Cantagallo Gabbro Whole rock     K/Ar 66 5 Santamaria and Schubert
(1974)
  Cantagallo Gabbro Whole rock     K/Ar 67 6 Santamaria and Schubert
(1974)
  Cantagallo Gabbro Whole rock     K/Ar 67 6 Santamaria and Schubert
(1974)
SVC-17 Choroni Gneiss Sphene 10.363 −67.586 FT 126 15 Kiser (1994)
DF-1203 Sebastapol Gneiss Muscovite 10.393 −67.042 K/Ar 41 2 Olmeta (1968)
CC-100 Guaremal Granite Biotite     K/Ar 32 2 Santamaria and Schubert
(1974)
CC-101 Guaremal Granite Biotite     K/Ar 31 1.8 Santamaria and Schubert
(1974)
CC-102 Guaremal Granite Biotite     K/Ar 33 2 Santamaria and Schubert
(1974)
CC-103 Guaremal Granite Biotite     K/Ar 30 1.9 Santamaria and Schubert
(1974)
CC-104 Guaremal Granite Biotite     K/Ar 30 1.8 Santamaria and Schubert
(1974)
CC-104 Guaremal Granite Biotite 10.363 −68.000 Rb/Sr 79 5 Loubet et al. (1985)
CC-105 Oritapo Diorite Biotite     K/Ar 76 3.9 Santamaria and Schubert
(1974)
CC-106 Oritapo Diorite Biotite     K/Ar 77 4 Santamaria and Schubert
(1974)
  Guaremal Orthogneiss Biotite 10.363 −68.000 K/Ar 33 3 Hess (1966)
  Guaremal Orthogneiss Biotite 10.363 −68.000 K/Ar 32 2 Santamaria and Schubert
(1974)
  Guaremal Orthogneiss Biotite 10.363 −68.000 K/Ar 33 2 Santamaria and Schubert
(1974)
SVC-24 Tovar Granite–gneiss Zircon 10.423 −67.281 FT 16.4 2.1 Kiser (1994)
  Puerto Cabello Garnet amphibolite Amphibole     K/Ar 32.4 1.2 Kohn et al. (1984)
  Cabo Codera Garnet amphibolite Amphibole 10.567 −66.061 K/Ar 155 7 Kohn et al. (1984)
10/27/15 10:03 AM
13880_ch02_ptg01_039-084.indd 63

  Oricao Garnet amphibolite Amphibole 10.584 −66.483 K/Ar 735 30 Kohn et al. (1984)
  Cabo Codera Garnet amphibolite Amphibole 10.567 −66.061 K/Ar 753 31 Kohn et al. (1984)
Gu. 2115 Tiara Formation Actinolite–metatuff Whole rock     K/Ar 100 10 Piburn (1968)

Cordillera de la Costa Belt (III)


Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
PA-119 Araya P. (El Augen gneiss Hornblende     K/Ar 128 11 Priem et al. (1966)
Mango-Dona
Juana)
PA-119 Araya P. (El Augen gneiss Whole rock     K/Ar 53 3 Priem et al. (1966)
Mango-Dona
Juana)
40
CaOf-01 Playa Medina Ultramafic Hornblende 10.714 −63.013 Ar/39Ar 148.5 29.9 Altamira-Areyan (2009)
40 39
CaOf-02 San Juan de Unare Ultramafic Whole rock 10.749 −62.741 Ar/ Ar 90 2 Altamira-Areyan (2009)
PA-117 Carupano area Dacite Whole rock     K/Ar 5 0.5 Priem et al. (1966)
PA-118 Carupano area Dacite Sanidine     K/Ar 5 0.6 Priem et al. (1966)
40
ParP-117 Paria (Boca del Gneiss Hornblende 10.730 −61.859 Ar/39Ar 97.3 14.1 Altamira-Areyan (2009)
Dragon)

The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  63


Villa de Cura Blueschist Belt–Las Hermanas Formation, Chacao Complex, and Basement in Serrania del Interior
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
40
AAVC-04176 Villa de Cura Blueschist White mica 10.035 −67.166 Ar/39Ar (2s) 86.2 3.3 Altamira-Areyan (2009)
Group
40
AAVC-04182 Villa de Cura Blueschist White mica 10.040 −67.167 Ar/39Ar (2s) 86.2 3.7 Altamira-Areyan (2009)
Group
40
AAVC-04228 Villa de Cura Blueschist Hornblende 9.883 −67.674 Ar/39Ar (2s) 93.4 5.4 Altamira-Areyan (2009)
Group
40
AAVC-04243 Villa de Cura Blueschist Hornblende 9.827 −67.834 Ar/39Ar (2s) 90.7 6.1 Altamira-Areyan (2009)
Group
40
AAVC-04251 Villa de Cura Blueschist Hornblende 9.902 −67.781 Ar/39Ar (2s) 91.7 4.6 Altamira-Areyan (2009)
Group
40
AAVC-04258 Villa de Cura Blueschist Hornblende 10.088 −66.489 Ar/39Ar (2s) 90.7 8.9 Altamira-Areyan (2009)
Group
CB 78-35 Villa de Cura Blueschist Amphibole     K/Ar (1s) 77 5 Loubet et al. (1985)
Group
CB 78-35 Villa de Cura Blueschist Phengite     K/Ar (1s) 87 3 Loubet et al. (1985)
Group
CB 78-35 Villa de Cura Blueschist Plagioclase     K/Ar (1s) 99 5 Loubet et al. (1985)
Group
(continued)
10/27/15 10:03 AM
13880_ch02_ptg01_039-084.indd 64

64  ALTAMIRA and Burke


Villa de Cura Blueschist Belt–Las Hermanas Formation, Chacao Complex, and Basement in Serrania del Interior (continued)
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
VC 25 Villa de Cura Blueschist Amphibole     K/Ar (1s) 88.4 3.1 Loubet et al. (1985)
Group
VC 25 Villa de Cura Blueschist White mica     K/Ar (1s) 82.5 2.5 Loubet et al. (1985)
Group
40
AAVC-04168i Las Hermanas Metavolcanic Hornblende 9.983 −67.128 Ar/39Ar (2s) 96.5 3.9 Altamira-Areyan (2009)
Formation breccia (clast)
40
AAVC-04164 Las Hermanas Metavolcanic Hornblende 9.979 −67.129 Ar/39Ar (2s) 86.2 3.3 Altamira-Areyan (2009)
Formation breccia (matrix)
40
AAVC- Las Hermanas Metavolcanic Hornblende 9.979 −67.129 Ar/39Ar (2s) 136.9 5 Altamira-Areyan (2009)
04164Xeno Formation breccia (clast)
VC 26 Las Hermanas Metavolcanic Amphibole     K/Ar (1s) 90.8 2.9 Loubet et al. (1985)
Formation breccia (matrix)
VC 26 Las Hermanas Metavolcanic White     K/Ar (1s) 192 6 Loubet et al. (1985)
Formation breccia (matrix) mica–epidote
VC 26 Las Hermanas Metavolcanic Plagioclase     K/Ar (1s) 107 11 Loubet et al. (1985)
Formation breccia (matrix)
VC 27 Las Hermanas Metavolcanic Amphibole     K/Ar (1s) 101 58 Loubet et al. (1985)
Formation breccia (matrix)

Chacao
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
40
Chacao Chacao Ultramafic Hornblende 9.885 −67.420 Ar/39Ar (2s) 121.6 17.2 Altamira-Areyan (2009)
CH 16 Chacao   Plagioclase     K/Ar (1s) 49.8 2 Loubet et al. (1985)
CH 8 Chacao   Plagioclase     K/Ar (1s) 58.5 2 Loubet et al. (1985)
CH 11 Chacao Fresh Hornblende     K/Ar (1s) 91 3.5 Loubet et al. (1985)
CH 7 Chacao Unmetamorphosed Hornblende     K/Ar (1s) 98 4 Loubet et al. (1985)
CH 16 Chacao Ultramafic to mafic Hornblende     K/Ar (1s) 101 4 Loubet et al. (1985)
complex
Ch 8 Chacao   Hornblende     K/Ar (1s) 104 4 Loubet et al. (1985)
ANT 184 Chacao Biotite–plagioclase Hornblende     K/Ar (1s) 97.2 3.5 Hebeda et al. (1984)
hornblendite
ANT 183 Chacao Hornblendite Hornblende     K/Ar (1s) 99.2 3.5 Hebeda et al. (1984)
ANT 182 Chacao Quartz gabbro Hornblende     K/Ar (1s) 106.6 3.5 Hebeda et al. (1984)

Camatagua–El Tinaco Belt


Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
RT-87-1 Tinaco Complex Gneiss       Rb/Sr 945 178 Sisson et al. (2005)
RT-87-2 Tinaco Complex Gneiss       Rb/Sr 945 178 Sisson et al. (2005)
10/27/15 10:03 AM
13880_ch02_ptg01_039-084.indd 65

RT-87-4 Tinaco Complex Gneiss       Rb/Sr 945 178 Sisson et al. (2005)
RT-87-5 Tinaco Complex Gneiss       Rb/Sr 945 178 Sisson et al. (2005)
40
VTO82-132 Tinaco Complex Granulite–   9.859 −66.431 Ar/39Ar 147.4 0.3 Sisson et al. (2005)
amphibolite
SVC-56 Tinaco Complex Trondjhemite Apatite 9.752 −68.418 FT 6.1 1.3 Kohn et al. (1984)
SVC-48 Tinaco Complex Diorite Zircon 10.188 −67.290 FT 41.9 4.9 Kohn et al. (1984)
SVC-48 Tinaco Complex Diorite Zircon 10.188 −67.290 FT 43.4 5.6 Kohn et al. (1984)
SVC-27 Tinaco Complex Gneiss Zircon 10.534 −67.343 FT 49 5.8 Kohn et al. (1984)
942-a Aguadita gneiss; Biotite gneiss Biotite     K/Ar 112.4 3 Hess (1966)
Tinaco Complex
942-a Aguadita gneiss; Biotite gneiss Hornblende     K/Ar 117.5 3 Hess (1966)
Tinaco Complex
Ti-3582 North of Tacata Amphibolite gneiss Amphibole     K/Ar 204 12 Olmeta (1968)
Ti-3582 North of Tacata Amphibolite gneiss Actinolite     K/Ar 210 10 Olmeta (1968)
TQZ Tinaco gneiss Hornblende gneiss Hornblende     K/Ar 235.8 13 Hess in edit in Urbani
(1982)
TQZ Tinaco gneiss Hornblende gneiss Pyroxene     K/Ar 684 55 Hess in Kugler (1972)
TQZ Tinaco gneiss Hornblende gneiss Hornblende     K/Ar 191 15 Hess in edit in Urbani
(1982)
40
V94-62 coarse Tinaquillo Complex Hornblendite vein   9.913 −66.437 Ar/39Ar 173.9 1.6 Sisson et al. (2005)
40
V94-62 fine Tinaquillo Complex Hornblendite vein   9.913 −66.437 Ar/39Ar 154.7 0.3 Sisson et al. (2005)

The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  65


40
V94-63 Tinaquillo Complex Gabbroic pegmatite   9.913 −66.437 Ar/39Ar 153.9 2 Sisson et al. (2005)
40
V97-4 Tinaquillo Complex Gabbroic pegmatite   9.913 −66.437 Ar/39Ar 190.6 1.2 Sisson et al. (2005)
40
VTO82-88 Tinaquillo Complex Gabbroic pegmatite   9.916 −66.439 Ar/39Ar 165.2 1.4 Sisson et al. (2005)
T90-6 Tinaquillo massif Granulite Zircon     U/Pb 150 2 Seyler et al. (1998)
T90-2b Tinaquillo massif Granulite Zircon     U/Pb 125 1 Seyler et al. (1998)
T90-12 Tinaquillo massif Garnet granulite Garnet–     Nd/Sm 87.1 3.1 Seyler et al. (1998)
hornblende
whole rock
Co-4505 Timurato Hill Basalt Whole rock     K/Ar 64.2 2.4 Hess (1966)
Co-4505 Timurato Hill Basalt Pyroxene     K/Ar 77 8 Hess (1966)
  Mogote Hill Granite K–feldspar     K/Ar 270 10 Feo-Codecido (1963)

Serrania del Interior


Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
V94-50 Embalse de Guarico Zircon 9.821 −66.944 FT 70.1 9.7 Sisson et al. (2005)
Camatagua
V94-53 Embalse de Garrapata Apatite 9.938 −66.916 FT 14.2 2 Sisson et al. (2005)
Camatagua
V94-53 Valle Morin Garrapata Zircon 9.938 −66.916 FT 51.1 6.4 Sisson et al. (2005)
V94-54 Valle Morin Garrapata Apatite 9.938 −66.914 FT 13.5 1.8 Sisson et al. (2005)

(continued)
10/27/15 10:03 AM
13880_ch02_ptg01_039-084.indd 66

66  ALTAMIRA and Burke


Serrania del Interior (continued)
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
V94-54 Valle Morin Garrapata Zircon 9.938 −66.914 FT 60.7 7.5 Sisson et al. (2005)
V94-55 Valle Morin Garrapata Apatite 9.938 −66.940 FT 15.8 2.8 Sisson et al. (2005)
V94-55 Valle Morin Garrapata Zircon 9.938 −66.940 FT 64.3 6.2 Sisson et al. (2005)
V94-56 Embalse de Guarico Zircon 9.813 −66.950 FT 53.9 13.3 Sisson et al. (2005)
Camatagua
V94-57 Embalse de Guarico Zircon 9.822 −66.942 FT 61.4 8.6 Sisson et al. (2005)
Camatagua
V94-59 East of Palmarito Quebradon Zircon 9.741 −66.592 FT 73.4 8.9 Sisson et al. (2005)
V94-60 North of Taguay Mucaria Zircon 9.821 −66.649 FT 85.1 19.7 Sisson et al. (2005)
V94-61 Altagracia de Quiamare Zircon 9.864 −66.354 FT 62.8 16.8 Sisson et al. (2005)
Orituco
GXB-2 Serrania del Granite Muscovite 9.302 66.939 K/Ar 330   Feo-Codecido et al. (1984)
Interior
GXB-8 Serrania del Granite Muscovite 9.368 66.939 K/Ar 321   Feo-Codecido et al. (1984)
Interior

Tobago
Sample Location Rock Type Mineral Lat Long Technique Age (Ma) ± (Ma) Reference
40 39
2D-14 Tobago Hornblende gabbro Hornblende     Ar/ Ar 91.4 2.2 Snoke et al. (1990)
40
1C-30 Tobago Hornblende Hornblende     Ar/39Ar 102.9 1.1 Snoke et al. (1990)
melagabbro
40
1C-29 Tobago Hornblende Hornblende     Ar/39Ar 102.8 1.2 Snoke et al. (1990)
porphyry
40
1C-5 Tobago Poikilitic hornblende Hornblende     Ar/39Ar 104.7 1.6 Snoke et al. (1990)
diorite
40
2D-35 Tobago Hornblende–gabbro Hornblende     Ar/39Ar 103.6 1.4 Snoke et al. (1990)
pegmatite
40
T827 Tobago Hornblende– Hornblende     Ar/39Ar 104.2 1.3 Snoke et al. (1990)
plagioclase
porphyry
40
DR245 Tobago A clast of Hornblende     Ar/39Ar ≥ 120   Snoke et al. (1990)
hornblende phyryc
andesite
10/27/15 10:03 AM
The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  67

faults, thrust faults, pull-aparts, and folding on vari- were involved in the eastward transform motion of the
ous scales (Hackley, 2005). Anticlines similar to those CARIB plate with respect to South America (Figure 9)
of the South Caribbean fold belt, and in our interpre- before the Maracaibo Block formed. There are two such
tation forming part of that belt, have been mapped at faults in the Guajira Peninsula, the Cuisa and Oca faults.
outcrop on islands (Pijpers, 1933; Helmers and Beets, The Cuisa fault crosses the peninsula, ending abruptly
1977; Beets, 1996) and offshore on the basis of their 100 km (62.1 mi) off the west coast. The western end of
bathymetric and shallow seismic expression (Silver the Oca fault lies at the northwest corner of the Marac-
et al., 1975; Clark et al., 2008, figure 6; and the “fam- aibo Block where the Oca fault meets the Santa Marta
ily 3 fault and fold structures” of Gorney et al., 2007). fault at an acute angle (Figure 9). The Oca fault extends
These anticlines occupy a belt about 150 km (93.2 mi) eastward across the peninsula, but cannot be traced as
wide south of the bounding thrust of the Maracaibo far as the east coast. We here suggest that the Oca and
Block (Figure 10). Folds (ca 20 3 10 km2) on Aruba and Cuisa faults were the same fault before the Maracaibo
Curacao that involve the late Miocene to Pliocene (ca Block formed and that they have been offset from one
7–1.8 Ma) aged Seroe Domi carbonate rocks (Jackson another during the past 15 Ma by left-lateral motion on
and Robinson, 1994) form part of that population of a northwest extension of the Burro Negro fault (Figure
anticlines (Figure 11). Because our interest here is in 9). An Eocene phase of right-lateral motion on that fault
the relationship of the Maracaibo Block to the PBZ, we has been described by Escalona and Mann (2006), but
concentrate mainly on the northwestern and north- our concern here is with a late Miocene to Recent time
eastern parts of the block and their structural evolu- of active left-lateral movement on the fault. (Evidence of
tion during the past 15 My. Those areas have been that movement in epicenters on the fault and in defor-
strongly modified as the block has developed. mation of late Miocene rocks is seen in figures 6 and 10
of Escalona and Mann, 2006.)
Northwestern Part of the Maracaibo Block The Burro Negro fault does not crop out on the
Prominent E–W-trending strike-slip faults on the Guajira Guajira Peninsula, although indications of its loca-
and Paraguana Peninsulas have been interpreted by tion can be discerned in epicenters, seismic reflection
Gomez (2001) to be representative of the complex of data (Gomez, 2001), and bathymetry. Gomez (2001)
strike-slip faults within the South CARIB PBZ that also observed a set of N–W-trending structures that

Figure 10. Active


deformation in part
of the South Carib-
bean deformed belt
illustrated here on a
bathymetry base by
showing anticlines that
deform young Seroe
Domi carbonate rocks
(late Miocene and
Pliocene in age Jackson
and Robinson, 1994),
mapped onshore in
Aruba, Curacao, and
Bonaire. Folds offshore
from sparker surveys of
Silver et al. (1975)
are similar although
not identical.
20 km (12.4 mi)

13880_ch02_ptg01_039-084.indd 67 10/27/15 10:03 AM


68  ALTAMIRA and Burke

Figure 11. Geological map


and cross section of Aruba
(Beets, 1996). The late
Neogene Seroe Domi
Formation dips from 10° to
30°. Those dips have been
considered depositional,
but because they extend
over 2 km (1.2 mi) and be-
cause they are concordant
with a mapped offshore
structure of Silver et al.
(1975) (see Figure 10),
we attribute them to active
folding. That folding ex-
poses the ca 90 Ma Aruba
batholith and the “Aruba
Lava,” which we consider
from its structure (see
the cross section in this
figure) and compositional
­variation to represent part
of the subduction complex
of the Great Arc of the
­Caribbean. 10 km (6.2 mi)

Figure 12. Rose diagram showing strikes


of steeply dipping joints from LANDSAT
images of the upper Guajira Peninsula that
Gomez (2001) interpreted as Riedel shears
in the N–S-shortening South Caribbean
deformed belt. The peak azimuth at ca 45°
is close to that of the Burro Negro fault ca
100 km (62.1 mi) to the southwest.

he identified on Landsat imagery of the whole of the Northeastern Part of the Maracaibo Block
upper peninsula (1) as steep joints and (2) as faults Earthquake distributions and mechanisms as well as GPS
with small throws (either normal faults or strike-slip results (Perez et al., 1997; Perez and ­Mendoza, 1998; Weber
faults in offshore seismic reflection data for late Mio- et al., 2001; Trenkamp et al., 2002; C
­ olmenares and Zoback,
cene sedimentary rocks (Figure 12). Gomez (2001) inter- 2003) show that the northeastern part of the Maracaibo
preted the consistent azimuth of the structures (N 45° Block is not symmetrical with the northwestern part, as it
W) to indicate N–S compression in the Guajira part of would be if the Bocono fault formed the eastern bound-
the ­Maracaibo Block and suggested that the structures, ary of the block (­ Figure 13). We instead follow Perez et al.
the trend of which roughly parallels that of the Burro (2001a) in finding that the block may be better regarded as
Negro and Santa Marta strike-slip faults, indicate that bounded by the Victoria fault farther to the ESE. That fault
they represent Riedel shears with respect to the north- has a much more nearly E–W azimuth than the Bocono
ward escape of the Maracaibo Block in this region. fault. We interpret the asymmetry as reflecting the fact

13880_ch02_ptg01_039-084.indd 68 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  69

Figure 13. Faults, folds, pull-


aparts, and GPS results for the
Maracaibo Block (gray shading).
The sketch below the main figure
emphasizes that ­movements
on the left-lateral Santa Marta
(SM) and Burro Negro (BN)
strike-slip faults oppose the
sense of ­motion of the Carib-
bean plate with respect to South
America, but that movement
on the La ­Victoria (LV) fault
adds to that motion. As a result,
the ­Maracaibo Block extends
farther to the east than has been
commonly assumed. Six GPS
measurements on islands east of
64° W and north of 11° N indicate
that the influence of the Marac-
aibo Block may even be felt as far
to the northeast as Barbados at
59.5° W. GPS results from Perez
et al. (2001), Trenkamp et al.
(2002), and Weber et al. (2001).
200 km (124.3 mi)

that while the Santa Marta and Burro Negro faults are reducing the E–W extent of the offshore island archi-
left-lateral faults, the Victoria fault is a right-lateral fault. pelago within the Maracaibo Block (i.e., from Los
As Figure 13 shows, movement of the CARIB to the east Monjes to La O ­ rchilla), closing the Cariaco trench, and
is largely in a direction opposite to that on the Santa Marta accommodating movement on the La Victoria fault
and Burro Negro faults. That accounts for the elevation provide a representation of the north coast of South
of the Sierra Nevada de Santa Marta nearly 6 km (3.7 mi) America at ca 15 Ma ­(Figure 14), and reveals the struc-
above sea level. Movement on the La V ­ ictoria fault adds ture within the ­Ribbon Continent on the north coast of
to, rather than opposes, that of the CARIB. GPS azimuths South America as it was before it became modified as
close to, but north of the Victoria fault provide the best result of the events in the Maracaibo Block.
indication of the extent of the Maracaibo Block to the east
(Perez et al., 2001b) (­Figure 13).
The Cariaco trench lies at the edge of the Maracaibo Structure of the Ribbon Continent in the Transform
Block and adjacent to the Victoria fault (Figure 9). It PBZ of the North Coast of South America
is not a normal pull-apart basin. We date the initia-
tion of the Cariaco trench to the time of origin of the Restoration of the structure of the transform PBZ
Maracaibo Block at 15 Ma. East of the Maracaibo as it existed before the development of the Marac-
Block, movement with respect to South America in the aibo Block reveals the existence of three belts within
PBZ is, as it has been for ca 65 My, due east. East–west the PBZ, each characterized by distinctive rock types
movement continues today as shown by GPS meas- (Figure 8). A fourth belt of related foreland basins lies
urements and earthquake mechanisms (Perez et al., immediately to the south of and adjacent to the PBZ. A
1997; Perez and Mendoza, 1998; Weber et al., 2001). We striking contrast between the Ribbon Continent on the
found it convenient (Figure 14) to restore movement in west coast of South America and the R ­ ibbon Continent
the whole of that eastern part of the PBZ to conditions of the north coast of South America is that extensive
at 15 Ma so as to match conditions farther to the west, areas containing CCOP fragments are confined to the
in which we have removed the effects of the penetra- west coast. Fragments of the CCOP on the north coast
tion of the PBZ by the escaping Maracaibo Block. That of South America are few and are restricted to slivers
involved the closure of the Cariaco trench. within the subduction complexes that outcrop within
Retro-deforming the ca 100-km (62.1-mi) offsets of the Northern Belt of the Ribbon Continent (Figure 14).
the Bocono and Santa Marta faults and the ca 75-km Structure within the PBZ at 15 Ma was complex,
(46.6-mi) offset on the Burro Negro fault, as well as and we do not here address complexities such as

13880_ch02_ptg01_039-084.indd 69 10/27/15 10:03 AM


70  ALTAMIRA and Burke

Figure 14. The Ribbon Continent restored to its 15 Ma position in Venezuela. Removal of offsets related to Maracaibo Block move-
ments reveals three W–E-trending belts of the Ribbon Continent. South Caribbean plate transform plate boundary zone faults bounded
the Ribbon Continent to the north and separated the Caribbean–Colombian Oceanic Plateau (CCOP) from rocks of the Great Arc.
Two faults of that set are shown that have offset La Blanquilla (depth to Moho: 24 km [14.9 mi]) from the Aves Swell and Los Testigos
(depth to Moho: 23 km [14.3 mi]) from the Lesser Antilles. Exposures in the Northern Belt south of the faults are dominantly of sub-
volcanic igneous intrusions and subduction complex rocks of the Great Arc of the Caribbean. The former crop out more widely on
islands within the active anticlines of the South Caribbean deformed belt. The southern boundary of the Northern Belt is formed by
faults of the PAROCOSSEP fault system in this area including the Cuisa, Oca, San Sebastien, and El Pilar faults. The Central Belt consists
entirely of subduction complex rocks of the Great Arc in numerous mountainous regions of Venezuela. The Southern Belt consists of
the folded and thrust Serrania del Interior and along strike (and presently within the Maracaibo Block) its western extension the Lara
nappe province. The northern part of the Serrania del Interior includes subduction complex rocks from the Central Belt thrust to the
south. The figure shows an estimated location (A–A') for the Ribbon Continent–South American boundary that is based on depths to
the Moho from Niu et al. (2007) (in large bold numerals). We have modified depths to Moho from Niu et al. (2007) (1) by restoring
horizontal offsets of the past 15 My on the Bocono and La Victoria faults and (2) by removing sedimentary rock thicknesses mainly
from Summa et al. (2003) and Feo-Codecido et al. (1984) (in small numerals) from receiver function depths to Moho to yield crystal-
line rock crustal thickness estimates. The figure shows that both the Central Belt and the Serrania del Interior have been thrust over the
South American continental margin to an extent that varies along strike. Clark et al. (2008, figure 6) showed similar thrusting of Great
Arc and Serrania del Interior rocks onto South America based on seismic velocity and reflection seismic analyses at ca 64° W. The belt
of foreland basins south of the Serrania del Interior is not shown in this figure. 1000 km (621.4 mi)

the Falcon–Bonaire Basin, but simply draw atten- Northern Belt with Arc Plutons and
tion to a ca 1400-km (870 mi)-long strike-slip fault the Subduction Complex Rocks
“COSSEP” (Cuisa–Oca–San Sebastian–El Pilar, see A ca 100-km (62.1-mi)-wide belt is dominated by sub-
Figure 15) fault as a prominent feature that divides volcanic plutons of the Great Arc and associated sub-
the ­Northern Belt from the Central Belt of the Ribbon duction complex rocks, mainly in green schist and
Continent within the PBZ at ca 10.5° N. (Figure 14). prehnite–pumpellyite facies. Folding and consequent
From north to south, three belts within the PBZ, all of structural elevation during the past 10 My within the
which trend close to E–W, are the (1) Northern Belt, ­Maracaibo Block has had the effect of exposing larger
(2) Central Belt, and (3) Southern Belt. areas of intrusives and relatively smaller areas of

13880_ch02_ptg01_039-084.indd 70 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  71

Figure 15. The Ribbon Continent in its


present position offset by faults of the
Maracaibo Block. The Serrania del Interior
straddles the Great Arc–South American
continent ­boundary (A–A' and B–B') as
a result of ­secondary ­thrusting from the
southern ­margin of the plate boundary
zone as indicated in Figure 16. The Cuisa
and Oca faults are interpreted to be
offset parts of a single fault active before
15 Ma. Depths to the Moho from Niu et
al. (2007) (in large ­numerals). We have
­modified depths to Moho from Niu et al.
(2007) by removing sedimentary rock
thicknesses mainly from Summa et al.
(2003) and ­Feo-Codecido et al. (1984)
(in small numerals) from receiver function
depths to Moho to yield crystalline rock
crustal thickness estimates. 1000 km
(621.4 mi)

subduction complex rocks on islands within the block Southern Belt Dominated by Cretaceous and
than are exposed farther to the south in the Guajira Cenozoic Sedimentary Rocks That Were Deposited
and P
­ araguana Peninsulas and farther to the east out- on the Rifted Margin of the North Coast of South
side the Maracaibo Block (Figure 9). In those areas America and Have Been Involved in Thrusting from
Great Arc intrusions comprise relatively smaller pro- the Southern Margin of the PBZ
portions of the outcrop. A belt ca 50–100 km (31.1–62.1 mi) wide is dominated
by Cretaceous to Cenozoic sedimentary rocks that were
Central Belt originally deposited on South American continental base-
A 50- to 100-km (31.1- to 62.1-mi)-wide belt is domi- ment at the rifted continental margin of South America
nated by subduction complex metamorphic rocks of after the Yucatan Block moved away. These rocks have
the Great Arc, mainly in prehnite–pumpellyite and been emplaced southward farther onto the continent by
green schist facies; however, blue schist facies occur secondary thrusting at the southern margin of the PBZ.
in one 50-km (31.1-mi)-wide by 100-km (62.1-mi)-long
lens. Slivers and knockers of serpentinite (after harz- Foreland Basin Belt
burgite), as well as basalt, blue schist, and eclogite, are A 100–150-km (62.1–93.2 mi)-wide belt developed south
sporadically distributed within the rocks of the sub- of the Southern Belt in response to loading of the conti-
duction complex. The subduction complex belt crops nental margin by rocks emplaced by thrusting from the
out in mainly mountainous areas such as the Cordil- Southern Belt of the PBZ. The oldest sedimentary rocks
lera de la Costa, inner part of the Serrania del Interior, in individual parts of the foreland basin belt become
Caucagua– El Tinaco Belt, Paracotos Belt, and Villa progressively younger from the west (in the Maracaibo
de Cura Belt, all on the ­Venezuelan mainland or in Basin, which was initiated in the late P­ aleocene ca 57 Ma)
the Northern Range of ­Trinidad. The belt exposes no to the east (in the Maturin Basin, initiated in the Miocene
Great Arc intrusions, but it does contain isolated frag- at ca 20 Ma). A summary of the geology of individual
ments, mainly of granodiorite, torn from the main- areas on the mainland and islands within the belts of the
land of South America that we correlate with rocks of PBZ, emphasizing aspects that have helped us in estab-
similar ages in the ­basement of the Cretaceous Andean lishing the evolution of the Ribbon Continent.
Arc of Ecuador and the equivalent Polymetamorphic
Complex of western Colombia. Rocks in both areas
have yielded Proterozoic, Paleozoic, and Triassic high- Northern Belt
temperature closure isotopic ages and younger low-
temperature closure (on micas 40Ar/39Ar ages) and The locations of formations are shown on Figure 9
fission track ages (on apatites and zircons). except for the Ruma, Siapana, and Caucagua–El

13880_ch02_ptg01_039-084.indd 71 10/27/15 10:03 AM


72  ALTAMIRA and Burke

Tinaco lithounits, which are too small in outcrop to The oldest fossils of the Guajira Peninsula, in the
show on the maps in this chapter. Caju “Formation” are late Jurassic ammonites, and
the youngest are in Cretaceous rocks of Turonian to
Campanian age (ca 93–70 Ma). The age range is con-
Guajira Peninsula sistent with that established for the Great Arc subduc-
tion complexes in Ecuador and western Colombia.
The Guajira Peninsula is not only the westernmost The occurrence of a La Luna–Querecual type “brown
part of the Ribbon Continent on the north coast of limestone” within the Great Arc rocks is particularly
South America but is also the area that shows the clos- interesting because that rock type is often considered
est structural resemblance to the west coast structure confined to mainland NW South America. Because
of the Ribbon Continent as seen in the Romeral suture the La Luna Formation is a representative of oceanic
and fault zone of the west coast of South America in anoxic episode 3 (Rey et al., 2004), we suggest that
Colombia (Gomez, 2001). The Upper Guajira Penin- its development on the Great Arc in a near equatorial
sula exposes prominent faults, including the 100-km latitude is not surprising. We consider Precambrian
(62.1-mi)-long E–W-trending trans-peninsula Cuisa igneous rocks, of which the Siapana body is the best
fault (Figure  15), that resemble those within the described, to resemble the fragments of the polymeta-
Romeral suture and fault zone of the west coast, with morphic complex basement of the Cretaceous Andean
the important difference that those of the Guajira trend Arc of western Colombia that were first mechanically
E–W, not N–S. This difference is readily explained by embodied into the Romeral suture and fault zone by
the paleomagnetic results of Macdonald and Opdyke faulting and were then carried by strike-slip motion to
(1972) who showed that the Cretaceous rocks they the north and east. Jurassic-aged rims to Proterozoic
studied in the Guajira, which we identify as Great Arc zircons in the Siapana body are consistent with that
of the Caribbean rocks, had become magnetized ca 20° interpretation (Molina et al., 2006). A younger intru-
to the south of their present latitude of ca 12° N, that sion has a published Cenozoic K/Ar whole-rock age
is, in the approximate latitude of Ecuador (at ca 2° S (Lockwood, 1966), but we consider that does not rep-
now and at 10° S at 75 Ma; Luzieux et al., 2006) rotated resent the time of its intrusion.
ca 90° clockwise to their present position.
Rocks of the Ruma Metamorphic zone of Alva- Paraguana Peninsula
rez (1967) include quartzites and phyllites that were MacDonald (1968) described low-grade metamorphic
derived from sandstones and carbonaceous shales, as rocks in prehnite–pumpellyite facies including phyl-
well as slivers of serpentinite and gabbro we interpret lites, graphitic schists, meta-arenites, and metacon-
to be representative of the subduction complex of the glomerates, with sparse carbonate rocks of the Pueblo
Great Arc in the Guajira. Blueschists are not known Nuevo Formation from the center of the Paraguana
in the outcrop in the Guajira, but an occurrence is Peninsula that contain ammonites of Kimmeridgian age
reported as a boulder in Cenozoic gravel thought to (ca 150 Ma). He suggested that this unit correlates with
be derived locally from underlying Mesozoic subduc- the Cocinas Formation of the Guajira Peninsula. We
tion complex rocks (Green et al., 1968). We interpret consider both to represent parts of the subduction com-
the Cretaceous and Jurassic rocks of La Guajira found plex of the Great Arc of the Caribbean. Both ­Gonzalez
in the Ruma Metamorphic zone and in the Cocinas de Juana et al. (1980) and Stephan (1985) suggested cor-
platform and Guajira trough units to be Great Arc relation of the rocks of the Pueblo Nuevo F ­ ormation
­subduction complex rocks that correspond to rocks with subduction complex rocks of the C ­ ordillera de la
in the Arquia and Quebradagrande Complexes of the Costa to the east. We suggest that granitic (U/Pb titan-
west coast. Descriptions of rocks of the Alta Guajira ite) ages of 265 Ma and 262 Ma reported by Martin-­
have involved attempts to establish stratigraphy and Bellizzia (1968) are from rocks transported north from
thicknesses (Renz, 1960; Rollins, 1965; Duque-Caro the west coast of South America.
and Reyes, 1999; Gomez, 2001). Because these rocks Farther to the southwest on the Paraguana Penin-
appear to represent the subduction complex environ- sula, a zoned mafic–ultramafic intrusive and volcanic
ment, we regard the estimated thicknesses published complex (Martin-Bellizzia, 1972) shows similarities to
as inappropriate. However, the timing of deposition, the sub-volcanic Great Arc plutonic suite of Tobago
which is constrained by fossils, reveals a duration that (Snoke et al., 2001).
corresponds well with the known history of the Great
Arc of the Caribbean. Ages as old as Triassic have been Los Monjes
assigned to lithounits without fossils on the basis of an Los Monjes exposes basalts in green schist facies
analogy with the La Quinta Formation in the Merida (Martin-Bellizzia, 1972; Ostos et al., 2005b) and has
Andes (Renz, 1960), but we doubt those assignations. yielded K/Ar whole-rock ages of 116 Ma and 114 Ma

13880_ch02_ptg01_039-084.indd 72 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  73

on a granitic rock (Santamaria and Schubert, 1974). It Curacao


too may represent rocks of the Great Arc. The Curacao Lava consists of “over 5 km (3.1 mi) of pil-
low basalts, reworked hyaloclastites, diabase sills and
Aruba a thin succession of siliceous shales and limestones”
About 10 km2 (6.2 mi2) of the island is underlain by a (Jackson and Robinson, 1994), and includes a pelagic
unit mapped as the Aruba Lava Formation of which layer with late–middle Albian (ca 104 Ma) ammonites.
another small area (ca 1 km2 [0.6 mi2]) has been iden- We interpret this “lava” to represent subduction com-
tified 10 km (6.2 mi) away at the island’s northern tip plex rocks of the Great Arc. MORB compositions have
(Beets, 1996). The Aruba Lava Formation is inappro- been identified, and the CCOP may be represented
priately named. It consists of a 3 km (1.9 mi) stack of by komatiites, some of which show spinifex textures,
rocks typical of the Great Arc subduction complex that as do CCOP rocks on Isla Gorgona off the west coast
are generally metamorphosed to prehnite–pumpellyite of Colombia (Echeverria, 1980). The Curacao Lava is
facies (­ Figure 11). The unit includes such rock types as overlain by a sedimentary unit, the Knip Formation,
basalts (both in pillows and in sheets), diabases, basaltic that is up to 2000 m (6561.7 ft) thick, and in which tur-
lapilli tuffs, volcaniclastic conglomerates, sandstones, bidites are abundant. We also consider the Knip Unit
pelagic cherts, cherty limestones, and fine-grained tur- to be part of the subduction complex. Knip Formation
bidites with intercalated pebbly mudstones that have fossils indicate a range of late Cretaceous ages includ-
yielded ammonites of possibly Turonian (93? Ma) age ing late Santonian to Campanian (85–75 Ma) (Jack-
(Helmers and Beets, 1977). If, as has been suggested son and Robinson, 1994). An overlying turbiditic and
(White et al., 1999), the CCOP is represented among conglomeratic unit, the Midden Curacao Formation,
the basaltic rocks of the Aruba Lava Formation, then it which we also consider to be part of the subduction
is as slivers in the subduction complex. There are geo- complex, resembles the underlying Knip Formation,
chemical grounds for considering that the CCOP is so but has been regarded as Paleocene (ca 65–55 Ma) on
represented (White et al., 1999), which has the tectonic the basis of what are described as “poorly preserved
implication that the Aruba Lava Formation subduction planktic foraminifera” (Jackson and Robinson, 1994).
complex includes material subducted under the Great Younger sedimentary rocks from Eocene to Recent in
Arc after the CCOP eruption date (ca 89 Ma). age on Curacao are dominantly carbonates (Jackson
That is consistent with a suggestion by White et and Robinson, 1994).
al. (1999) that the Aruba batholith was formed in the
Great Arc of the Caribbean from ocean-floor rocks Bonaire
that contained the CCOP. The Aruba batholith under- On Bonaire, the Washikemba, and Rincon Formation
lies most of the island (ca 200 km2 [124.3 mi2] in four that exceed 5 km (3.1 mi) in thickness, we also assign
separate outcrops) and has a composition that could to the Great Arc subduction complex. They include
reflect subduction of the CCOP. Altamira-Areyan et al. in prehnite–pumpellyite and zeolite facies “pillow
(in press) obtained 40Ar/39Ar igneous intrusive ages on lavas, lava flows, shallow intrusions and subaqueous
hornblendes of ca 90 Ma and 94 Ma and also on horn- pyroclastic flows that alternate with pelagic and vol-
blendes from a diabase of the “Aruba Lava F ­ ormation.” caniclastic sediments” (Jackson and Robinson, 1994).
Although Aruba is an island of only ca 250 km 2 Fossils are of late Albian (ca 100 Ma) and middle–late
(155.3 mi2) in area, its geology typifies the history of the Maastrichtian (75–66 Ma) ages. Pelagic sedimentary
Great Arc intrusion–­dominated belt of the Ribbon Con- rocks include cherts, radiolarites, and cherty shales
tinent on the north side of South America in two ways: (Jackson and Robinson, 1994). Compositionally the
(1) occurrences of CCOP rocks are scarce in the belt, and volcanic rocks embrace compositions from mafic to
where they do occur are in slivers within the subduction felsic and are representative of the primitive island
complex, and (2) metaigneous and metasedimentary arc series of Donnelly and Rogers (1978) and Donnelly
rocks identified as representing various environments et al. (1990). The overlying Seobi Blanco Formation has
in earlier papers and maps (Beets and MacGillavry, been assigned to the Paleocene but is unfossiliferous.
1977; ­Helmers and Beets, 1977; Beets, 1996; White et al., It has yielded boulders, thought to be of South Ameri-
1999; Wright and Wyld, 2004) have generally turned can origin, with U/Pb zircon ages of 1150 Ma (Priem
out to be recognizably subduction complex rocks of et al., 1986), which is an age more readily matched in
the Great Arc, often in prehnite–pumpellyite and green the Andes of the west coast of South America than in
schist facies. Those rocks do represent a range of envi- the Venezuelan basement.
ronments but only environments that are to be expected
in a subduction complex where simultaneous deposi- Los Roques
tion and tectonism is likely to juxtapose igneous, sedi- On Gran Roque, the largest of these tiny islets, a ca
mentary, and metamorphic rocks. 3-km (1.8-mi)-long by 1-km (0.6-mi)-wide strip of

13880_ch02_ptg01_039-084.indd 73 10/27/15 10:03 AM


74  ALTAMIRA and Burke

outcrops, exposes metavolcanic rocks reminiscent of ratios, suggest a volcanic arc origin. If that interpre-
those of the Great Arc that are cut by quartz–diorite tation is valid, we consider that arc to have been the
and aplite–­pegmatite dikes (ca 66 Ma K/T boundary, Great Arc of the Caribbean and the intrusive ages to
see Table 1). Ostos and Sisson (2005) obtained trace record a time when the Great Arc was in the Pacific
­element data consistent with the conclusion of San- Ocean before it had collided with the CCOP. Granitic
tamaria and Schubert (1974), that the islands overlay augen gneiss at El Salado has a volcanic arc composi-
rocks of a volcanic island arc underlain by oceanic tional affinity (Ostos and Sisson, 2005). A U/Pb zircon
crust; however, the rocks could also represent arc age of 86 Ma shows it to have been intruded into the
intrusions and subduction complex rocks. Great Arc before the arc collided with the west coast
of South America, close to the time when the Great
La Blanquilla Arc collided with the CCOP. Last, the Los Robles
La Blanquilla, the most eastern island in the Maracaibo Group is a low-grade metasedimentary sequence of
Block (Figure 9), has yielded an igneous age of 59 Ma quartz–chlorite phyllite, marble, dolomitic marble,
(Levander et al., 2006) that we attribute to its deriva- calcareous phyllite, and metaconglomerate without
tion from the southern tip of the Aves Swell. We do not knockers or amphibolite (Ostos and Sisson, 2005).
consider it to be a part of the Great Arc that collided Margarita exposes Pliocene and middle Eocene
with the west coast of South America, but to represent ­s edimentary rocks and the Punta Garnero Group
the rotated Lesser Antillean tip of the part of the Great (Hunter, 1978) of limestones, shales, and graywackes
Arc that entered the Atlantic Ocean on the CARIB that have ­e xperienced brittle deformation but no
and traveled to the east (Figure 8). This age matches ­m etamorphism. Miocene sedimentary rocks show
the youngest igneous ages of the Aves Swell phase of faulting but are relatively undeformed.
the Great Arc that ceased to be active at the time of In summary, igneous rock intrusive ages (with the
formation of the Grenada Basin during the Paleocene exception of those of diabase dikes) and related struc-
(ca 65–55 Ma) (Figure 8). tures on Margarita that record steps 1–4 of Stockert et
al.’s (1995) “integrated tectonic evolution” sequence of
Margarita events are ca 86 Ma or older, and relate to times before
The island of Margarita lies outside and to the east the Great Arc collided with the west coast of South
of the Maracaibo Block (Figure 9), in the area in America (at ca 70 Ma). However, structures dateable
which Great Arc intrusions crop out much less than by using lower temperature closure systems, which
do subduction complex rocks. Rocks assigned to the record steps 5 to 12, are all younger than 66 Ma and
Cretaceous consist entirely of subduction complex “suggest deformation at a transform margin” (see
material in which serpentinite tectonic slices (after Stockert et al., 1995, Table 1). That is consistent with
harzburgite) occupy more than 10% of the outcrop. our attribution of that deformation to movement in
Three metamorphic units are distinguished. The first, the PBZ of the Ribbon Continent.
the Juan Griego “Group,” a metasedimentary unit
with quartzo-feldspathic, marble, and metaquartzite Trinidad: Northern Range
lithologies, including numerous eclogite and amphi- This area might be better treated with the Araya–Paria
bolite knockers and lenses. A granitic augen gneiss at Peninsula of Venezuela, but we place our treatment
Macanao has yielded six zircon fractions that gave a here because it lies to the north of the El Pilar sector
U/Pb discordia plot with an upper intercept of 315 of the PAROCOSSEP fault system. That also enables
(+35/−24) Ma (Stockert et al., 1995). We consider this us to treat Venezuela as an entity and to associate
gneiss to be tectonically incorporated into the sub- ­Trinidad with Tobago.
duction complex from a mainland of South America Green schist facies rocks making up the North-
environment such as the Polymetamorphic Complex ern Range of Trinidad, a ca 20-km (12.5-mi)-wide
of Colombia. The second (Rinconada “Group”) is E–W-trending strip of rugged and highly vegetated
composed of mafic rocks metamorphosed at eclog- hills, have been identified as having been formed in
ite facies conditions. It has been suggested to be of a subduction complex environment because of their
MORB composition and considered to represent for- close links to the rocks of the Araya–Paria Peninsula
mer oceanic crust (Bocchio et al., 1990). Trondjhemites in neighboring Venezuela (Ostos and Sisson, 2005).
in the Rinconada Group at Metasiete and Guayacan Although knockers of serpentinite and eclogite such
that have yielded U/Pb zircon ages in the 114–105 as those that are known in the Araya–Paria Penin-
Ma range (Stockert et al., 1995) might be considered sula have not been recognized in Trinidad, a ca 87-Ma
to represent ocean-floor rocks; however, Ostos and (whole-rock K/Ar) basaltic body at Sans Souci (Wadge
Sisson (2005), using trace element and Sr isotopic and MacDonald, 1985) has been suggested to have

13880_ch02_ptg01_039-084.indd 74 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  75

either been generated at a spreading center or to be Delta from the north coast of South America. Since
a sliver of the CCOP. Minor volcanic horizons have then, as the Lesser Antillean Arc continued to sweep
also been recognized in some of the other Northern east, the Cordillera de la Costa of Venezuela has risen
Range units (Donovan, 1994). The green schists of the within the migrating transform PBZ and the Orinoco
Northern Range are mainly pelitic but more siliceous has been diverted eastward into its present course. The
rocks and carbonate rocks also crop out. As elsewhere, huge sediment volume of the Barbados accretionary
efforts have been made to establish thicknesses, suc- prism resulted from the subduction zone picking up
cessions, and stratigraphies in the Northern Range, three separate deep-sea fans. Sandstones from the two
and these have been fruitful to the extent that faunas older fans are known in outcrop and in the subsurface
of late Jurassic (ca 150 Ma) and Barremian (130–125 of Barbados where Eocene and Miocene fossilifer-
Ma) age have been identified. ous deep-sea fan sandstones have been distinguished
(Baldwin, 1986; Burke, 1988; Speed, 1994).
Tobago
Tobago is recognized to be “a fragment of the accreted
Mesozoic oceanic arc of the Southern Caribbean” The Central, Subduction Complex–dominated, Belt of
(Snoke et al., 2001). We identify that arc to be the Great the Southern CARIB PBZ
Arc of the Caribbean. Three east–west-striking belts
of Cretaceous arc rocks cross the island. (1) The north The Central Belt of the PBZ crops out only on the Ven-
coast schists are metamorphosed volcanogenic rocks, ezuelan mainland. No Great Arc sub-volcanic plutonic
mainly in green schist facies, that we identify as sub- rocks crop out in the Central Belt, which lies south of
duction complex rocks of the Great Arc. A 40Ar/39Ar both the Paraguana Peninsula and the pre-15-Ma loca-
age on a hornblende indicates an age of ca 120 Ma tion of the PAROCOSSEP fault (Figure 9). South of
(Aptian) or older. (2) Rocks of the Composite ­Plutonic those boundaries the Ribbon Continent is represented
Suite cross the center of the island in a belt about only by (1) rocks that can be assigned to the subduction
3 km (1.8 mi) wide. They intrude the north coast complex environment of the Great Arc of the Caribbean
schists and are themselves deformed. (3) The rocks of and (2) slivers up to ca 10 km (6.2 mi) long, mainly of
the Tobago volcanic group that occupy the southern granite and gneiss, that have sometimes in the past been
half of the island are “consanguineous” with the Com- regarded as basement to the rocks of the Central Belt.
posite Plutonic Suite (Snoke et al., 2001). Isotopic ages We interpret these rocks to include (1) metamorphosed
of the Composite Plutonic Suite and Tobago volcanic Great Arc igneous rocks and (2) tectonic fragments of
group and fossil ages for the latter indicate igneous South American continental basement that we consider
activity at ca 104 Ma (roughly middle Albian). to have been caught up from such sources as the Poly-
metamorphic Complex of Colombia during late Creta-
Indeterminate Character of the Structure of the ceous (ca 75 Ma to ca 65 Ma) transpression, as the Ribbon
Ribbon Continent in the South Caribbean PBZ to Continent PBZ moved to the north along the west coast
the East of Trinidad and Tobago of South America (Figure 1). The southern boundary
The South Caribbean PBZ must extend ca 300 km of the Central Belt in Venezuela is well defined on the
(186.5 mi) ­f arther to the east and be linked to the basis of the receiver function results of Niu et al. (2007),
Lesser Antilles subduction zone plate boundary in the who concluded that stations with depths to Moho of
neighborhood of Long 58° W, but it is not possible to 30 km (18.6 mi) (or less) overlie island arc rocks (i.e.,
establish structure in that region because of the excep- Great Arc of the Caribbean rocks) and that stations with
tional sediment thickness of the overlying Orinoco depths to Moho of ca 40 km (24.8 mi) (or more) over-
River deep-sea fan and of the adjacent Barbados accre- lie South American continental crust. We show results
tionary prism into which that fan merges (Figure 14). of Niu et al. (2007) slightly modified in two ways in
The huge volume of the Barbados accretionary prism Figure 14:
is a direct result of the evolution of the Ribbon Conti- (1) Where information is available we have deducted
nent. As the Lesser Antillean segment of the Great Arc sedimentary rock thicknesses, using Summa et al.
swept to the east during the past 65 My, the southern (2003) and Feo-Codecido et al. (1984), so that we have
end of its subduction zone first picked up the deep-sea indicated approximate crystalline crustal thicknesses
fan of an early Magdalena River delta from the Atlan- rather than total depths to Moho. This has enabled us
tic-type north coast of South America. That was dur- to smooth the mapped shape of the southern boundary
ing the Eocene at ca 50 Ma. Later during the Oligocene of the central zone published by Niu et al. (2007).
at ca 30 Ma the subduction zone, as it traveled to the (2) We have removed the effects of the movements
east, picked up the deep-sea fan of the then Orinoco of the Maracaibo Block so that the boundary defined by

13880_ch02_ptg01_039-084.indd 75 10/27/15 10:03 AM


76  ALTAMIRA and Burke

receiver functions is shown approximately as it would lies within the Caucagua–El Tinaco Belt. It has been
have been at ca 15 Ma before the Maracaibo Block suggested to represent either of Jurassic continental
began to move. That also results in a smoother bound- margin-rift rocks or ocean-floor rocks, but “a MORB
ary and addresses the issue raised by Niu et al. (2007), protolith is suggested by variation diagrams and REE
that three continental receiver function thicknesses patterns” (Ostos and Sisson, 2005). Knockers of blue
(38, 38, and 37 km [23.6, 23.6, and 23 mi]) appear to schist, eclogite, serpentinite, and basalts are sporadi-
lie within the Great Arc region. The restored structure cally distributed in some but not all of the belts. Ser-
(1) aligns the ultramafic and mafic bodies at Siquisique, pentinites ornament many thrust and strike-slip faults
within the Maracaibo Block, with a line of ultramafic between and within the blocks of the Central Belt, and
outcrops close to the Victoria fault and with closure all the belts “are thought to be allochthonous” with
of the Cariaco trench, adds coherence to the align- respect to South America (Unger et al., 2005).
ment of ultramafic rocks on the pre-15-Ma structure 2. The subduction complex of the Central Belt in Ven-
map, and (2) aligns the Lara nappes from within the ezuela has yielded fossils in the age range between mid-
Maracaibo Block with the rest of the thrust belt of Jurassic (ca 170 Ma) and late Cretaceous (­Campanian to
the Serrania del Interior (Figure 14). Maastrichtian ca 80–65 Ma) (Ostos et al., 2005b). That
Results of BOLIVAR program research in the is consistent with the range of Great Arc sedimentary
Central Belt and reviews of earlier petrological and depositional ages. Isotopic ages on protoliths are few
geochemical research in a recently published book but are consistent with an origin in the Great Arc.
(Ave-Lallemant and Sisson, 2005) are consistent with 3. Isotopic ages for HP/LT metamorphism in the
the Ribbon Continent model for the belt: Central Belt are consistent with a cessation of subduc-
1. Subduction complex rocks are recognized within tion-related metamorphism no later than ca 70 Ma.
the Central Belt of Venezuela in: (i) the Cordillera de 4. The question of how the blocks of the C ­ entral
la Costa (Figure 9), which is “built on a subduction Belt have been juxtaposed has attracted attention
mélange of mid- to late Cretaceous age” (Sisson et al., because of the diversity of metamorphic facies
2005). (ii) The Paracotos Belt that consists of Campa- (HP/LT eclogite to LP/LT prehnite–pumpellyite
nian to Maastrichtian (80–65 Ma) sedimentary rocks facies and lower) among the blocks. The Villa de
with slices of serpentinite, gabbro, pillow basalts, Cura Block has, for example, been suggested to be a
and radiolarian cherts. The Paracotos Belt (Figure 9, “klippe,” and the possibility that it has been thrust
inset A) exposes “sedimentary structures akin to fly- from the north over blocks closer to the coast has
sch” (Ostos et al., 2005a). (iii) The Villa de Cura Belt been raised (Martin-­B ellizzia, 1972). However, we
(Figure 9), the protolith of which has been suggested prefer the interpretation of Lugo (2000) (referred to
to be of Aptian–Albian age, includes metachert and by Ostos et al., 2005b) that the Villa de Cura Block
metagraywacke units (ca 120–97 Ma), four 40Ar/39Ar represents an imbricate thrust sheet within the
ages in amphiboles gave 90.7 ± 6.1 Ma, 91.7 ± 4.6 ­subduction complex.
Ma, 93.4 ± 5.4 Ma, and 90.7 ± 8.9 Ma; and two ages The juxtaposition of a variety of metamorphic facies
of white micas gave 86.2 ± 3.7 Ma and 86.2 ± 3.3 Ma among blocks of the Ribbon Continent is consistent
(see Table 1). The Villa de Cura Belt has been shown with the observation of the widespread occurrence of
to geochemically represent an oceanic, tholeiitic island tectonically incorporated South American continental
arc with its much-studied blue schists representing rocks in subduction complex blocks of the Central Belt.
two forearc slivers (Unger et al., 2005). (iv) The Las Those occurrences, which are known as far to the south
­Hermanas Formation, which includes volcanic breccia, as Ecuador and southern Colombia, show that the Rib-
ash tuff, lithic tuff, and lava in prehnite–pumpellyite bon Continent has been in contact with South America
metamorphic facies. Las Hermanas rocks are thought since it initially collided with the continent by ca 70 Ma
to represent either island arc-or back arc-magmatism (Figure 7C). Blocks of various subduction complex
(Ostos and Sisson, 2005). A volcanic breccia included rocks with various metamorphic facies represent
in this formation gave an 40Ar/39Ar age of 136.9 ± 5.0 equilibration at depths of ca 10–60 km (6.2–37.3 mi).
Ma for hornblende concentrates on a tonalitic clast. The roughly 50-km (31-mi) range in displayed depth is
Another sample from the same unit gave an age of 96.5 very small, compared with the distances of hundreds
± 3.9 Ma; the matrix fraction gave an 40Ar/39Ar age of to thousands of km that the blocks have traveled first
131.3 ± 6.9 Ma. The Caucagua–El Tinaco Belt consists northward and then eastward. No special process is
of gneisses that Ostos and Sisson (2005), mainly on required for the exhumation of HP/LT rocks in sub-
the basis of rare-earth element compositions, consid- duction zones, although it may be possible to relate
ered to have formed in a “volcanic or magmatic arc.” a particular occurrence to a particular environment
The Tinaquillo Peridotite ­Complex (Ostos et al., 2005a) such as a transition in an arc from a convergent to a

13880_ch02_ptg01_039-084.indd 76 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  77

transform boundary (e.g., Ave-Lallemant, 1996). Exhu-


mation of blue schists, by a variety of processes, is a
part of the normal process of subduction complex evo-
lution (e.g., Karig, 1980).
Establishing the exact timing of the j­uxtaposition
of individual blocks depends on a variety of
­c onsiderations including the determination of
­low-temperature closure timings by a variety of
­methods. Results from Venezuela (Sisson et al., 2005)
show complexity, because the methods are relatively
insensitive to low-temperature change in exhumation
history, and have not yet yielded a coherent picture.

Southern Belt

This belt is dominated by Cretaceous and Cenozoic


sedimentary rocks that were deposited on the rifted
margin of the north coast of South America and have
been involved in thrusting from the southern margin
of the PBZ. These rocks occupy the Serrania del Inte-
rior and occur also within the Maracaibo Block as the
western extension of the Serrania in the Lara nappes
(Figures 14 and 15). In their more northern locations,
Serrania del Interior folds and thrusts show struc-
tural continuity with the southern thrusts of the Cen-
tral Belt that expose subduction complex rocks, while
Figure 16. Cross section at ca 64° W modified from Clark
farther south the folds and thrusts of the Serrania
et al. (2008) by (1) extension of the section to the regional
expose rifted margin sedimentary rocks in both silici-
seismogenic maximum depth in the plate boundary zone of
clastic and carbonate facies that were deposited on the ca 80 km (49.8 mi); (2) extending the secondary thrusting
north coast of South America from the late Jurassic of the ­Volcanic Arc (the Great Arc of the Caribbean) over
time when the Yucatan Peninsula rifted away (Pindell, the Caribbean–Colombian Oceanic Plateau (CCOP) in the
1994). The extended continental margin on which these ­Venezuelan Basin; (3) emphasizing the relative importance
sedimentary rocks were deposited has been thrust to of the PAROCOSSEP fault system, which is here locally the El
the south onto South America by secondary shorten- Pilar (EP) fault; and (4) emphasizing the thrusting of rocks
ing associated with the eastward propagation in the of the Great Arc, the sedimentary rocks of the rifted South
Ribbon Continent PBZ. The amount, and probably also American Margin, and the South American basement near
the azimuth of shortening, has varied with location for the southern end of the section. 10 km (6.2 mi)
several reasons. For example, the continental margin
is likely to have been both irregular and wider from (Hackley, 2005), and that has been interpreted either
north to south in some places than in others. The stack- (1) by applying the results of Riedel’s (1929) experi-
ing of the thrusts has probably also been irregular. ment, as indicating secondary folding related to the
The southern boundary of the Ribbon Continent, E–W trend of the transform Ribbon Continent PBZ
as mapped by Niu et al. (2007) using receiver func- (Robertson and Burke, 1989), or (2) by applying an
tions and as modified by us (Figure 15), indicates “oblique convergence” model indicating the nor-
the ­Serrania del Interior straddles the edge of South mal to the azimuth of the approach of a hypothetical
America. That is because of the varied extent of thrust- ­converging object.
ing at the edge of the PBZ (Figure 16 modified from
figure 6 in Clark et al., 2008). In some areas subduction
complex rocks in outcrop as well as rifted continental Foreland Basin Belt
margin rocks have been thrust tens of kilometers onto
the continent. A belt of foreland basins south of the Serrania del
Fold axis and thrust strike trends in the Serrania Interior extends from the Maracaibo Basin in the west
del Interior typically have an azimuth of ca N 60° E (Lugo and Mann, 1995) to the Maturin Basin in the east

13880_ch02_ptg01_039-084.indd 77 10/27/15 10:03 AM


78  ALTAMIRA and Burke

(Summa et al., 2003) (upper Cretaceous to Cenozoic   5. As the Ribbon Continent moved, its transpres-
autochthonous units in Figure 9) for a distance of more sional relationship to the Andean Margin of
than 700 km (435 mi). The belt has formed in response to ­Colombia led to the narrowing of the Great Arc
the loading of the edge of the South American continent material into the ca 600-km (372.9 mi)-long and ca
by rocks thrust onto the continental margin in the Serra- 100-km (62.1-mi)-wide Romeral suture and fault
nia del Interior. Foreland basins are commonly modeled zone. In that fault zone fragments from the base-
as responses to loads applied by thrusting in a direction ment to the Andean Margin (the “Polymetamor-
normal to the greatest length of the basin, but that is not phic Complex”) were tectonically incorporated
what has happened in Venezuela. The load has been into the Great Arc subduction complex (between ca
applied progressively from west to east as the tip of the 70 Ma and ca 65 Ma, during Maestrichtian times).
Ribbon Continent PBZ propagated in that direction.   6. At ca 65 Ma the Great Arc, carrying the CCOP be-
The elastic flexure of the continental lithosphere below hind it, began to enter the Atlantic Ocean, as a new
the basin depends on where the load has been applied CARIB separated from the Farallon plate.
and not the direction from which the load came.   7. Fragments of the Ribbon Continent began to be
Jacome et al. (2003a,b), (2005), (2008), Govers and carried to the east where they can be identified
Wortel (2005), and Clark et al. (2008) have drawn from characteristic lithologies and ages (e.g., hav-
attention to the complexities in structural develop- ing igneous and HP/LT ages > 70 Ma) as parts of
ment likely to be associated with what Govers and the Great Arc of the Caribbean that had collided
Wortel (2005) called a “subduction-transform edge with the west coast of South America.
propagator” (STEP). Phenomena such as shallow   8. Three W–E-trending belts characterize the Ribbon
mantle convection–influenced structures in addition Continent of the north coast of South America.
to elastic loads are suggested to characterize the STEP   9. The Northern Belt consists of Great Arc sub-­
environment. For that reason establishing the controls volcanic igneous intrusions and subduction com-
on both the Serrania del Interior thrust belt and Vene- plex rocks mainly exposed in islands off the north
zuelan foreland basin evolution is likely to be difficult. coast of South America. In some places we have
identified rocks as representing the Great Arc sub-
duction complex only from their published, in-
Conclusions cluding geochemical, descriptions. Rocks with an
origin in the CCOP have been only locally iden-
  1. Recognition of the nearly simultaneous cessation tified. They occur as slivers, some of which are a
of volcanic arc igneous and HP/LT metamorphic kilometer or more in length within the subduction
activity shortly before ca 70 Ma (Campanian time) complex.
in rocks of the Great Arc of the Caribbean, now dis- 10. The Central Belt consists entirely of Great Arc of
tributed over a distance of ca 3000 km (1864.1 mi) the Caribbean subduction complex rocks in the
close to the NW margin of South America from mountainous areas of Venezuela. A minor but sig-
the Gulf of Guayaquil to Tobago, has enabled us nificant part of the complexes consists of granitic
to recognize a “Ribbon Continent of the NW South and gneissic rocks of Paleozoic and Proterozoic
American Cordillera.” age. Those rocks have, in some cases, been con-
  2. The assembly of the Ribbon Continent, which is re- sidered to represent basement to the complexes,
corded in the forearc of the Ecuadorian Andes, was but we interpret them to be fragments of the South
established in a two-stage process. The first stage in- American continent, perhaps from the Polymeta-
volved the collision of the Great Arc of the Caribbean morphic Complex of Colombia tectonically incor-
with the CCOP at ca 84 Ma (late Santonian times). porated into the subduction complex. If that is the
  3. The second stage, following polarity reversal of case, the subduction complex rocks must have been
the Great Arc at its collision with the CCOP, in- in contact with the continent at the time of their in-
volved collision of the combined Great Arc and corporation. That was from ca 70 Ma.
CCOP with the Ecuadorian Margin, probably then 11. The third belt of the Ribbon Continent on the
an active Andean Margin at ca 70 Ma. north coast of South America consists of the fold-
  4. Immediately following the second collision (at ca and-thrust belt of the Serrania del Interior which
70 Ma), the combined CCOP and Great Arc began mainly involves sedimentary rocks deposited
to travel to the north in a ca 200-km (124.2 mi)- on the rifted margin of South America between
wide transpressional PBZ occupied by a newly as- late Jurassic and Cenozoic times. The Serrania
sembled Ribbon Continent consisting of fragments del Interior formed by secondary thrusting as
of both the CCOP and the Great Arc. the Ribbon Continent traveled eastward in the

13880_ch02_ptg01_039-084.indd 78 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  79

transform PBZ progressively from west to east. Altamira-­Areyan, A., P. Copeland, and D. Foster, 40Ar/39Ar
The Serrania straddles the margin of the South hornblende ages from Venezuela and the offshore ­islands:
American continent to a varying degree. in press.
12. The structure of the north coast of South America Alvarez, W., 1967, Geology of the Simarua and Carpintero
areas, Guajira Peninsula, Colombia: Ph.D. Thesis,
has been radically perturbed during the past 15
­Princeton, NJ.
My by the northward and eastward escape of the
Aspden, J. A., W. J. McCourt, and M. Brook, 1987, Geo-
Maracaibo Block which, in response to the Panama metrical control of subduction-related magmatism: The
Arc collision, has cut right across the PBZ. Remov- ­M esozoic and Cenozoic plutonic history of western
ing the effects of movement of the Maracaibo Block ­Colombia: GSL Journal, v. 144, p. 893–905.
has been essential in establishing the structure of Ave-Lallemant, H. G., 1996, Displacement partitioning and
the Ribbon Continent on the north coast of South arc-parallel extension; example from the southeastern
America. Our restoration has been quite crude. Caribbean Plate Margin, in G. E. Bebout, D. W. Scholl, S.
More refined restoration would be useful. H. Kirby, and J. P. Platt, eds., Subduction: Top to b ­ ottom:
13. We have characterized rocks and structures in Geophysical monograph series 96: Washington, DC,
many areas in this chapter. Testing our model by American Geophysical Union, p. 113–118.
Ave-Lallemant, H. G., and V. B. Sisson, 2005, Caribbean-
field and laboratory work involving age determi-
S o u t h A m e r i c a n p l a t e i n t e r a c t i o n s , Ve n e z u e l a :
nations, and by major and trace element geochem-
GSA Special Paper 394, 346 p.
istry is an obvious need for the future. Baldwin, S. L., 1986, Fission track evidence for the source
14. Isotopic work has established that volcanic arc of accreted sandstones, Barbados, in T. M. Harrison and
igneous and HP/LT metamorphic activity ended K. Burke, eds., Tectonics: Washington, DC, American
shortly before ca 70 Ma (Campanian time) in rocks ­Geophysical Union, 457 p.
of the Great Arc of the Caribbean that are now dis- Barat, F., B. Mercier de Lepinay, M. Sosson, C. Muller,
tributed over a distance of ca 3000 km (1864.1 mi) P. O. Baumgartner, and C. Baumgartner-Mora, 2014,
close to the NW margin of South America from the Transition from the Farallon plate subduction to the
Gulf of Guayaquil to Tobago (Table 1). That result ­collision between South and Central America: Geological
has enabled us to recognize a Ribbon Continent evolution of the Panama Isthmus: Tectonophysics, v. 622,
p. 145–167.
that includes the Northwestern South American
Beardsley, A. G., and H. G. Ave-Lallemant, 2007, Oblique
Cordillera.
collision and accretion of the Netherlands Leeward Antil-
les to South America: Tectonics, v. 26, p. 1–16.
Beets, D. J., 1996, Geological map of Aruba, scale 1: 50,000:
Acknowledgments Foundation for Scientific Research in the Caribbean Region
and the Geological Survey of The Netherlands (RGD).
Much of the work described in the chapter was per- Beets, D. J., and H. J. MacGillavry, 1977, Outline of the
formed while the first author was a Ph.D. student in Cretaceous and Early Tertiary history of Curaçao,
the Earth and Atmospheric Sciences Department in Bonaire and Aruba, in H. J. MacGillavry, ed., Eighth
the University of Houston. Thanks to Professors Peter Caribbean geological conference guide to the field ex-
cursions on Curaçao, Bonaire, and Aruba: GUA Papers
Copeland, Jack Casey, and David Foster (of the Univer-
of Geology 10: Amsterdam, The Netherlands, Stichting
sity of Florida) for their support toward completion of
GUA, p. 1–6.
that degree. Participation in the Broadband Onshore– Bocchio, R., L. De Capitani, G. Liborio, W. V. Maresch, and
Offshore Lithospheric Investigation of Venezuela and A. Mottana, 1990, The eclogite-bearing series of Isla
the Antilles Arc region “BOLIVAR” NSF-funded pro- ­M argarita, Venezuela: Geochemistry of metabasic
ject, and interaction with scientists involved in that ­lithologies in the La Rinconada and Juan Griego Group:
project in the field and laboratory, particularly Pro- Lithos, v. 25, p. 55–69.
fessors Hans Ave-Lallemant and Virginia Sisson (Rice Bucher, W. H., 1952, Geological structure and orogenic
University), is greatly appreciated. We would also like history of Venezuela: GSA Memoir 94, 113 p.
to thank Joan Flinch and Joshua Rosenfeld for provid- Burke, K., 1988, Tectonic evolution of the Caribbean:
ing a comprehensive review of the manuscript. ­Annual Review of Earth and Planetary Sciences, v. 16,
p. 201–230.
Burke, K., 1996, The African plate, South African Journal of
Geology, South Africa, Bureau for Scientific Publications
REFERENCES CITED at the Foundation for Education, Science and Technology:
Pretoria, South Africa, p. 339–409.
Altamira-Areyan, A., 2009, The Ribbon Continent of North- Burke, K., 2007, Dancing continents: Science, v. 318, 1385 p.
western South America: Ph.D. Thesis, University of Hou- Burke, K., and J. M. Cannon, 2014, Plume-plate interaction:
ston, Houston, Texas, 193 p. Canadian Journal of Earth Science, v. 51, p. 208–221.

13880_ch02_ptg01_039-084.indd 79 10/27/15 10:03 AM


80  ALTAMIRA and Burke

Calle, B., H. Gonzalez, R. De La Pena, E. Escorce, and Shell Exploradora y ­Productora de Colombia B. V.,
J. Durango, O. Ramirez, E. Alvarez, M. Calderon, 131 p.
J. Alvarez, G. Guarin, C. Rodriguez, J. Munosz, Echeverria, L. M., 1980, Tertiary or Mesozoic komatiites
and J. Duran, 1980, Mapa Geologico de Colombia—Escala from Gorgona Island, Colombia: Field relations and geo-
1:100.000, Plancha 166—Jerico: Bogota, INGEOMINAS. chemistry: Contributions to Mineralogy and Petrology, v.
Centeno-Garcia, E., M. Guerrero-Suastegui, and O. Talavera- 73, p. 253–266.
Mendoza, 2008, The Guerrero Composite Terrane of Ernst, R. E., and K. L. Buchan, 2003, Recognizing mantle
western Mexico: Collision and subsequent rifting in plumes in the geological record: Annual Review of Earth
a supra-subduction zone, in A. Draut, P. D. Clift, and and Planetary Sciences, v. 29, p. 469–523.
D. W. Scholl, eds., Formation and applications of the sedi- Escalona, A., and P. Mann, 2006, Tectonic controls of the
mentary record in arc collision zones: GSA Special Paper, right-lateral Burro Negro tear fault on Paleogene struc-
v. 436, p. 279–308. ture and stratigraphy, northeastern Maracaibo Basin:
Clark, S. A., A. Levander, M. B. Magnani, and C. Zelt, 2008, AAPG Bulletin, v. 90, p. 479–504.
Negligible convergence and lithospheric tearing along Farris, D. W., C. Jaramillo, G. Bayona, S. A. ­Restrepo-Moreno,
the Caribbean–South American plate boundary at 64°W: C. Montes, A. Cardona, A. Mora, R. J. Speakman, M.
Tectonics, v. 27, doi:10.1029/2008TC002328. D. Glascock, and V. Valencia, 2011, Fracturing of the
Coates, A. G., L. S. Collins, M. P. Aubry, and W. A. Berggren, ­Panamanian Isthmus during initial collision with South
2004, The geology of Darien, Panama, and the late America: Geology, v. 39, p. 1007-1010.
Miocene-Pliocene collision of the Panama Arc with Feininger, T., 1980, Eclogite and related high pressure
northwestern South America: GSA Bulletin, v. 116, ­regional metamorphic rocks from the Andes of Ecuador:
p. 1327–1344. Journal of Petrology, v. 21, p. 107–140.
Colmenares, L., and M. D. Zoback, 2003, Stress field Feo-Codecido, G., F. D. J. Smith, N. Aboud, and E. Di
and seismotectonics of northern South America, in ­G iacomo, 1984, Basement and Paleozoic rocks of
M. D. Zoback, ed., Geology Boulder: GSA, p. 721–724. the ­Venezuelan Llanos basins, in W. E. Bonini, R. B.
Correa-Martinez, A. M., 2007, Petrogênese e evolução do ­Hargraves, and R. Shagam, eds., The Caribbean-South
ofiolito de Aburra, Cordilheira Central dos Andes Co- American plate boundary and regional tectonics:
lombianos: Doutorado Thesis, Universidade de Brasilia, GSA Memoir 162, p. 175–187.
Brasilia, 178 p. Garnero, E. J., T. Lay, and A. K. McNamara, 2007,
Dalmayrac, B., and P. Molnar, 1981, Parallel thrust and Implications of lower mantle structural heterogeneity
normal faulting in Peru and constraints on the state for existence and nature of whole mantle plumes, in
of stress: Earth and Planetary Science Letters, v. 55, G. R. Foulger and D. M. Jurdy, eds., Plates, plumes,
p. 473–481. and planetary processes: GSA Special Paper 430,
Donnelly, T. W., 1994, The Caribbean sea floor, in S. K. Dono- p. 79–102.
van, and T. A. Jackson, eds., Caribbean geology: An intro- Gomez, I., 2001, Structural style and e­ volution of the Cuisa
duction: Kingston, Jamaica: University of the West Indies fault system, Guajira, Colombia: M.Sc. T
­ hesis, University
Publisher’s Association, p. 41–64. of Houston, Houston, Texas, 147 p.
Donnelly, T. W., D. Beets, M. J. Carr, T. Jackson, G. Klaver, Gonzalez, H., 1980, Geologia de las planchas 167 (Sonson)
J. Lewis, R. Maury, H. Schellekens, A. L. Smith, G. Wadge, y 187 (Salamina): Boletin Geologico Ingeominas, v. 23,
and D. Westercamp, 1990, History and tectonic setting of 174 p.
Caribbean magmatism, in G. Dengo and J. E. Case, eds., Gonzalez, H., 1993, Mapa Geologico del ­Departamento de
The Caribbean region: The geology of North America Caldas: Bogota, INGEOMINAS, 62 p.
v. H, p. 339–374. Gonzalez, H., 2001, Geologia de las planchas 206, Manizales
Donnelly, T. W., and J. J. W. Rogers, 1978, The distribution of y 225, Nevado del Ruiz, Escala 1:100.000, Memoria expli-
igneous rock suites throughout the Caribbean: Geologie cativa. Bogota, INGEOMINAS, 92 p.
en Mijnbouw, v. 57, p. 151–162. Gonzalez, H., and A. Nunez, 1991, Mapa geologico generali-
Donovan, S. K., 1994, Trinidad, in S. K. Donovan and T. A. zado del Departamento del Quindio: Bogota, INGEOMI-
Jackson, eds., Caribbean geology: An introduction: King- NAS, 42 p.
ston, Jamaica, University of the West Indies Publisher’s Gonzalez de Juana, C., J. M. Iturralde de Arozena, and
Association, p. 209–228. X. Picard Cadillat, 1980, Geologia de Venezuela y de
Droxler, A. W., K. C. Burke, A. D. Cunningham, A. C. Hine, sus cuencas petroliferas (Geology of Venezuela and its
E. Rosencrantz, D. S. Duncan, P. Hallock, and E. Robin- petroliferous basins): Caracas, Ediciones FONINVES,
son, 1998, Caribbean constraints on circulation between 1031 p.
Atlantic and Pacific Oceans over the past 40 million years, Goossens, P. J., and J. W. I. Rose, 1973, Chemical composi-
in T. J. Crowley and K. C. Burke, eds., Tectonic boundary tion and age determination of tholeiitic rocks in the
conditions for climate reconstructions: New York, Oxford Basic Igneous Complex, Ecuador: GSA Bulletin, v. 84,
University Press, p. 169–191. p. 1043–1051.
Duque-Caro, H., and R. Reyes, 1999, Seismic biostratigra- Gorney, D., A. Escalona, P. Mann, M. B. Magnani,
phy study of the Guajira region (onshore and offshore), A. Levander, G. Christeson, C. A. Zelt, M. Schmitz,
Proprietary information from Texas Petroleum Company S. Clark, M. C. Guedez, M. Bezada, Y. Arogunmati,

13880_ch02_ptg01_039-084.indd 80 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  81

T. Aitken, and A. Beardsley, 2007, Chronology of Ce- the Caribbean: Hydrocarbon habitats, basin formation,
nozoic tectonic events in western Venezuela and the and plate tectonics, AAPG Memoir 79, p. 735–749.
Leeward Antilles based on integration of offshore seis- Jacome, M. I., K. Rondon, and A. Carballo, 2005, Integrated
mic reflection data and on-land geology: AAPG Bul- geodynamic modelling of the Cordillera Central thrust
letin, v. 91, p. 653–684. belt and the Guarico Basin, North-Central region, Ven-
Govers, R., and M. J. R. Wortel, 2005, Lithosphere tearing ezuela, in 6th International Symposium on Andean Geo-
at STEP faults: Response to edges of subduction zones: dynamics, p. 400–403.
Earth and Planetary Science Letters, v. 236, p. 505–523. Jacome, M. I., K. Rondon, M. Schmitz, C. Izarra, and E. Viera,
Green, D. H., J. P. Lockwood, and E. Kiss, 1968, Eclogite and 2008, Integrated seismic, flexural and gravimetric model-
almandine-jadeite-quartz rock from the Guajira Penin- ling of the Coastal Cordillera thrust belt and the Guarico
sula, Colombia, South America: The American Mineralo- Basin, North-Central Region, Venezuela: Tectonophysics,
gist, v. 53, p. 1320–1335. v. 459, p. 27–37, doi:10.1016/j.tecto.2008.03.008.
Green, T. H., and A. Ringwood, 1968, Genesis of the calc- Jaillard, E., M. Ordonez, J. Suarez, J. Toro, D. Iza, and W.
alkaline igneous rock suite: Contributions to Mineralogy Lugo, 2004, Stratigraphy of the late Cretaceous–Paleo-
and Petrology, v. 18, p. 105–162. gene deposits of the cordillera occidental of central
Gutscher, M. A., W. Spakman, H. Bijwaard, and E. R. ecuador: Geodynamic implications: Journal of South
­Engdahl, 2000, Geodynamics of flat subduction: Seismic- American Earth Sciences, v. 17, p. 49–58.
ity and tomographic constraints from the Andean Mar- Johnston, S. T., 2008, The cordilleran ribbon continent of
gin: Tectonics, v. 19, p. 814–833. North America: Annual Review of Earth and Planetary
Hackley, P. C., 2005, Geologic shaded relief map of Sciences, v. 36, p. 495–530.
­Venezuela, in F. Urbani, A. W. Karlsen, and C. P. Garrity, Karig, D. E., 1980, Material transport within accretionary
eds., Open-file report: Reston, Virginia, USGS. prisms and the “knocker” problem: Journal of Geology,
Helmers, H., and D. J. Beets, 1977, Geology of the v. 88, p. 27–39.
­Cretaceous of Aruba, in H. J. MacGillavry, ed., Eighth Kerr, A. C., J. Tarney, G. F. Marriner, A. Nivia, and A. D.
Caribbean geological conference guide to the field Saunders, 1997, The Caribbean-Colombian Cretaceous
­e xcursions on Curaçao, Bonaire, and Aruba: GUA Igneous Province: The internal anatomy of an oceanic
­P apers of Geology 10: Amsterdam, The Netherlands, plateau, in J. J. Mahoney and M. F. Coffin, eds., Large
Stichting GUA, p. 29–35. igneous provinces: Continental, oceanic, and planetary
Hildebrand, R. S., 2009, Did westward subduction cause flood volcanism: Geophysical Monograph 100: Washing-
Cretaceous–Tertiary orogeny in the North American ton, DC, American Geophysical Union, p. 123–144.
Cordillera? GSA Special Paper 457, p. 1–71. Leroy, S., A. Mauffret, P. Patriat, and B. M. de Lepinay, 2000,
Hildebrand, R. S., 2013, Mesozoic assembly of the North An alternative interpretation of the Cayman trough evo-
American Cordillera: GSA Special Paper 495, 169 p. lution from a reidentification of magnetic anomalies: Ge-
Hildebrand, R. S., and J. B. Whalen, 2014, Arc and slab failure ophysical Journal International, v. 141, p. 539–557.
magmatism in Cordilleran batholiths I–The Cretaceous Levander, A., M. Schmitz, P. Mann, G. Christeson, J. Wright,
Coastal batholith of Peru and its role in South American H. Ave-Lallemant, C. Zelt, J. Pindell, G. Pavlis, and F.
orogenesis and hemispheric subduction flip: Geoscience Niu, 2006, BOLIVAR: Investigating island arc accretion
Canada, v. 41, p. 255–282. along the southeastern Caribbean plate boundary: GSA
Hoernle, K., P. van den Bogaard, R. Werner, B. Lissinna, Abstracts with Programs, p. 210.
and F. Hauff, 2002, Missing history (16–71 Ma) of the Litherland, M., J. Aspden, and R. A. Jemielita, 1994, The meta-
Galapagos hotspot: Implications for the tectonic and morphic belts of Ecuador: British Geological Survey Over-
biological evolution of the Americas: Geology, v. 30, seas Memoir 11, p. 147.
p. 795–798. Lockwood, J. P., 1966, Geology of the Serrania de Jarara area:
Hunter, V. F., 1978, Notes on the Tertiary stratigraphy of Ph.D. Thesis, Guajira Peninsula, Colombia: Princeton
Margarita Island: Geologie en Mijnbouw, v. 57, University, 237 p.
p. 189–192. Lugo, J., and P. Mann, 1995, Jurassic-Eocene tectonic evolu-
Jackson, T. A., and E. Robinson, 1994, The Netherlands and tion of Maracaibo Basin, Venezuela, in A. J. Tankard, R. S.
Venezuelan Antilles, in S. K. Donovan and T. A. Jackson, Soruco, and H. J. Welsink, eds., Petroleum basins of South
eds., Caribbean geology: An introduction: Kingston, America: GSA Memoir 62, p. 699–725.
Jamaica, University of the West Indies Publisher ’s Luzieux, L. D. A., F. Heller, R. Spikings, C. F. Vallejo, and
Association, p. 249–263. W. Winkler, 2006, Origin and Cretaceous tectonic his-
Jacome, M. I., N. Kusznir, F. Audemard, and S. Flint, 2003a, tory of the coastal Ecuadorian forearc between 1 degrees
Formation of the Maturin foreland basin, eastern Vene- N and 3 degrees S: Paleomagnetic, radiometric and fos-
zuela: Thrust sheet loading or subduction dynamic topog- sil evidence: Earth and Planetary Science Letters, v. 249,
raphy: Tectonics, v. 22, 1 p. doi:10.1029/2002TC001381. p. 400–414.
Jacome, M. I., N. Kusznir, F. Audemard, and S. Flint, 2003b, MacDonald, W. D., 1968, Estratigrafia, estructura y meta-
Tectono-stratigraphic evolution of the Maturin foreland morfismo, rocas del Jurasico Superior, Peninsula de Para-
basin: Eastern Venezuela, in C. Bartolini, R. T. Buffler, guana, Venezuela: Boletin de Geologia (Caracas), v. 9, p.
and J. Blickwede, eds., The Circum-Gulf of Mexico and 441–458.

13880_ch02_ptg01_039-084.indd 81 10/27/15 10:03 AM


82  ALTAMIRA and Burke

MacDonald, W. D., and N. D. Opdyke, 1972, Tectonic rota- Ostos, M., H. G. Ave-Lallemant, and V. B. Sisson, 2005a, The
tions suggested by paleomagnetic results from northern alpine-type Tinaquillo peridotite complex, Venezuela:
Colombia, South America: Journal of Geophysical Re- Fragment of a Jurassic rift zone? in H. G. Ave-­Lallemant
search, v. 77, p. 5720–5730. and V. B. Sisson, eds., Caribbean/South American
Martin-Bellizzia, C., 1968, Edades isotopicas de rocas ven- plate interaction, Venezuela: GSA, Special Paper 394,
ezolanas: Boletin de Geologia, Ministerio de Minas e Hi- p. 207–222.
drocarburos, Caracas, v. 10, p. 356–383. Ostos, M., and V. B. Sisson, 2005, Geochemistry and
Martin-Bellizzia, C., 1972, Sistema montañoso del Caribe, ­t ectonic setting of igneous and metaigneous rocks
corde Sur de la Placa Caribe ¿es una Cordillera Aloctona? of northern Venezuela, in H. G. Ave-Lallemant and
in C. Petzall, ed., Memoirs 6th Caribbean Geological Con- V. B. Sisson, eds., Caribbean-South American plate
ference, Margarita, Venezuela, p. 247–258. interactions, Venezuela: GSA, Special Paper 394,
Maya, M., and H. Gonzalez, 1996, Unidades litodemicas en p. 119–155.
la Cordillera Central de Colombia: Boletin Geologico, Ostos, M., F. Yoris, and H. G. Ave-Lallemant, 2005b,
Ingeominas, v. 35, p. 43–57. ­Overview of the southeast Caribbean/South ­American
McCourt, W. J., and J. A. Aspden, 1983, A plate tectonic plate boundary zone, in H. G. Ave-Lallemant and
model for the Phanerozoic evolution of central and V. B. Sisson, eds., Caribbean-South American plate
southern Colombia, in 10th Caribbean Geological Confer- interactions, Venezuela: GSA, Special Paper 394,
ence Transactions, INGEOMINAS, p. 38–47. p. 53–90.
McCourt, W. J., J. A. Aspden, and M. Brook, 1984, New geo- Perez, O. J., R. Bilham, R. Bendick, N. Hernandez, M. Hoyer,
logical and geochronological data from the Colombian J. R. Velandia, C. Moncayo, and M. Kozuch, 2001a, Ve-
Andes: Continental growth by multiple accretion: GSL locidad relativa entre las placas del Caribe y Suramerica
Journal, v. 141, p. 835–841. a partir de observaciones dentro del sistema de posicion-
McCourt, W. J., and T. Feininger, 1984, High pressure meta- amiento global (GPS) en el norte de Venezuela: Intercien-
morphic rocks in the Central Cordillera of Colombia: Brit- cia, v. 26, p. 69–74.
ish Geological Survey Report Series, v. 16, p. 28–35. Perez, O. J., R. Bilham, R. Bendick, J. R. Velandia, N. Her-
McKenzie, D., and N. Weiss, 1975, Speculations on the ther- nandez, C. Moncayo, M. Hoyer, and M. Kozuch, 2001b,
mal and tectonic history of the Earth: Geophysical Jour- Velocity field across the southern Caribbean plate
nal International, v. 42, p. 131–174. boundary and estimates of Caribbean/South-American
Mejia, M., E. Alvarez, H. Gonzalez, and E. Grosse, 1983a, plate motion using GPS geodesy 1994–2000: Geophysi-
Mapa Geologico de Colombia - Escala 1:100.000, Plancha cal Research Letters, v. 28, p. 2987–2990.
146 - Medellin Occidental: Bogota, INGEOMINAS. Perez, O. J., and J. S. Mendoza, 1998, Sismicidad y tectonica
Mejia, M., E. Alvarez, H. Gonzalez, and E. Grosse, 1983b, en Venezuela y areas vecinas: Fisica de la Tierra, v. 10, p.
Mapa Geologicode Colombia – Escala 1:100.000, Plancha 87–110.
130 – Santa Fe de Antioquia: Bogota, INGEOMINAS. Perez, O. J., C. Sanz, and G. Lagos, 1997, Microseismicity,
Molina, A. C., U. G. Cordani, and W. D. MacDonald, 2006, tectonics and seismic potential in southern Caribbean
Tectonic correlations of pre-Mesozoic crust from the and northern Venezuela: Journal of Seismology, v. 1,
northern termination of the Colombian Andes, Caribbean p. 15–28.
region: Journal of South American Earth Sciences, v. 21, Pijpers, P. J., 1933, Geology and paleontology of Bonaire:
p. 337–354. Physiographisch-geologische reeks 8, Utrecht: Geog. en
Montes, C., G. Bayona, A. Cardona, D. M. Buchs, C. A. Silva, Geol. Mededeel, s Uitgevers-mijv, 103 p.
S. Moron, N. Hoyos, D. A. Ramirez, C. A. Jaramillo, and Pindell, J. L., 1994, Evolution of the Gulf of Mexico and
V. Valencia, 2012, Arc-continent collision and orocline for- the Caribbean, in S. K. Donovan and T. A. Jackson, eds.,
mation: Closing of the Central American seaway: Journal Caribbean geology: An introduction: Kingston, Jamaica,
of Geophysical Research: Solid Earth, v. 117, p. B04105. University of the West Indies Publisher ’s Association,
Mosquera, D., 1978, Geologia del Cuadrangulo K-8: Informe p. 13–39.
1763 (unpublished): Bogota, INGEOMINAS, 63 p. Pindell, J. L., S. C. Cande, W. C. Pitman, III, D. B. Rowley, J.
Niu, F., T. Baldwin, G. Pavlis, F. Vernon, H. Rendon, M. F. Dewey, J. LaBrecque, and W. Haxby, 1988, A plate-kin-
Bezada, and A. Levander, 2007, Receiver function study ematics framework for models of Caribbean evolution:
of the crustal structure of the southeastern Caribbean Tectonophysics, v. 155, p. 121–138.
plate boundary and Venezuela: Journal of Geophysical Pindell, J., and L. Kennan, 2001, Kinematic evolution of the
Research: Solid Earth, v. 112, doi:10.1029/2006JB004802. Gulf of Mexico and Caribbean, GCSSEPM Foundation
Nivia, A., G. F. Marriner, A. C. Kerr, and J. Tarney, 2006, The 21st Annual Research Conference Transactions, Petro-
Quebradagrande Complex: A Lower Cretaceous ensialic leum Systems of Deep-Water Basins, p. 193–220.
marginal basin in the Central Cordillera of the Colombian Pindell, J., L. Kennan, W. V. Maresch, K.-P. Stanek, G. Draper,
Andes: Journal of South American Earth Sciences, v. 21, and R. Higgs, 2005, Plate-kinematics and crustal dy-
p. 423–436. namics of circum–Caribbean arc-continent interac-
Orrego, A., H. Cepeda, and G. I. Rodriguez, 1980, Esquistos tions: Tectonic controls on basin development in
glaucofanicos en el area de Jambalo: Geologia Norandina, proto-­Caribbean ­margins, in H. G. Ave-Lallemant and
v. 10, p. 161–202. V. B. Sisson, eds., Caribbean-South American plate

13880_ch02_ptg01_039-084.indd 82 10/27/15 10:03 AM


The Ribbon Continent of South America in Ecuador, Colombia, and Venezuela  83

interactions, V ­ enezuela: GSA, Special Paper 394, Santamaria, F., and C. Schubert, 1974, Geochemistry
p. 7–52. and geochronology of the southern Caribbean-north-
Priem, H. N. A., D. J. Beets, and E. A. T. Verdurmen, 1986, ern Venezuela plate boundary: GSA Bulletin, v. 85,
Precambrian rocks in an early Tertiary conglomerate on p. 1085–1098.
Bonaire, Netherlands Antilles (southern Caribbean bor- Sengor, A. M. C., and B. A. Natalin, 1996, Paleotectonics of
derland): Evidence for a 300 km eastward displacement Asia: Fragment of a synthesis, in A. Yin and T. M. Har-
relative to the South American mainland? Geologie en Mi- rison, eds., The tectonics of Asia: New York, Cambridge
jnbouw, v. 65, p. 35–40. University Press, p. 486–640.
Ratschbacher, L., L. Franz, M. Min, R. Bachmann, U. Mar- Silver, E. A., J. E. Case, and H. J. MacGillavry, 1975,
tens, K. Stanek, K. Stubner, B. K. Nelson, U. Herrmann, ­Geophysical study of the Venezuelan borderland: GSA
and B. Weber, 2009, The North American-Caribbean Bulletin, v. 86, p. 213–226.
plate boundary in Mexico-Guatemala-Honduras, in K. H. Sinton, C. W., R. A. Duncan, M. Storey, J. Lewis, and J. J. Es-
James, M. A. Lorente, and J. L. Pindell, eds., The origin trada, 1998, An oceanic flood basalt province within the
and evolution of the Caribbean plate: GSL, Special Publi- Caribbean plate: Earth and Planetary Science Letters,
cations 328, p. 219–293. v. 155, p. 221–235.
Renz, O., 1960, Geologia de la parte sureste de la Pen- Sisson, V. B., H. G. Ave-Lallemant, M. Ostos, A. E. Blythe, L.
insula de la Guajira (Republica de Venezuela), in 3rd W. Snee, P. Copeland, J. E. Wright, R. A. Donelick, and L.
Venezuelan Geological Congress: Special Publication R. Guth, 2005, Overview of radiometric ages in three al-
3, p. 317–374. lochthonous belts of northern Venezuela: Old ones, new
Restrepo, J. J., and J. F. Toussaint, 1982, Metamorfismos su- ones, and their impact on regional geology, in H. G. Ave-
perpuestos en la Cordillera Central de Colombia: Actas Lallemant and V. B. Sisson, eds., Caribbean-South Ameri-
del V Congreso Latinoamericano de Geologia, v. 3, p. can plate interactions, Venezuela: GSA, Special Paper 394,
505–512. p. 91–118.
Restrepo, A., J. F. Toussaint, H. Gonzalez, U. G. Cordani, K. Snoke, A. W., D. W. Rowe, J. D. Yule, and G. Wadge, 2001,
Kawashita, E. Linares, and C. Parica, 1991, Precisiones ge- Petrologic and structural history of Tobago, West Indies:
ocronologicas sobre el Occidente Colombiano: Simposio A fragment of the accreted Mesozoic Oceanic-Arc of the
sobre magmatismo andino y su marco tectonico: Mani- Southern Caribbean: GSA, Special Paper 354, 54 p.
zales, v. 1, p. 1–21. Speed, R. C., 1985, Cenozoic collision of the Lesser-Antilles
Rey, O., J. A. Simo, and M. A. Lorente, 2004, A record of Arc and continental South America and the origin of the
long- and short-term environmental and climatic change El Pilar fault: Tectonics, v. 4, p. 41–69.
during OAE3: La Luna Formation, Late Cretaceous Speed, R. C., 1994, Barbados and the Lesser Antilles forearc,
­(Santonian–early Campanian), Venezuela: Sedimentary in S. K. Donovan and T. A. Jackson, eds., Caribbean ge-
Geology, v. 170, p. 85–105. ology: An introduction: Kingston, Jamaica, University of
Reynaud, C., E. Jaillard, H. Lapierre, M. Mamberti, and the West Indies Publisher’s Association, p. 179–192.
G. H. Mascle, 1999, Oceanic plateau and island arcs of Stephan, J. F., 1985, Andes et Chaine Carïbe sur la trans-
southwestern Ecuador: Their place in the geodynamic versale de Barquisimeto (Venezuela): Evolution geo-
evolution of northwestern South America: Tectonophys- dynamique (Andes and Caribbean chain south of the
ics, v. 307, p. 235–254. Barquisimeto passage (Venezuela)), in A. Mascle, ed., Ge-
Riedel, W., 1929, Zur Mechanik geologischer Brucherschei- odynamique des Carïbes: Paris, France, Editions Technip,
nungen, Zentralblatt für Mineralogie, Geologie und p. 505–529.
Paläontologie, Abt. 1929B, p. 354–368. Stockert, B., W. V. Maresch, M. Brix, C. Kaiser, A. Toetz, R.
Robertson, P., and K. Burke, 1989, Evolution of southern Kluge, and G. Kruckhans-Lueder, 1995, Crustal history
­C aribbean plate boundary, vicinity of Trinidad and of Margarita Island (Venezuela) in detail: Constraint on
­Tobago: AAPG Bulletin, v. 73, p. 490–509. the Caribbean plate-tectonic scenario: Geology, v. 23,
Rogers, R., P. Mann, P. Emmet, and M. Venable, 2007, p. 787–790.
Colon fold-thrust belt of Honduras: Evidence for late Summa, L. L., E. D. Goodman, M. Richardson, I. O. Norton,
­C retaceous collision between the continental Chortis and A. R. Green, 2003, Hydrocarbon systems of north-
Block, in P. Mann, ed., Geologic and tectonic ­development eastern Venezuela; plate through molecular scale-­analysis
of the Caribbean plate in Northern Central America: of the genesis and evolution of the eastern Venezuela
GSA, Special Paper 428, p. 129–150. ­Basin: Marine and Petroleum Geology, v. 20, p. 323–349.
Rollins, J. F., 1965, Stratigraphy and structure of the Goajira Torsvik, T. H., and K. Burke, 2014, Large igneous province
Peninsula, northwestern Venezuela and northeastern locations and their connections with the core mantle
­Colombia: University of Nebraska studies: new series 30, boundary: EGU General Assembly Conference Abstracts,
102 p. 2881 p.
Rutherford, E., K. Burke, and J. Lytwyn, 2001, Tectonic Torsvik, T. H., M. A. Smethurst, K. Burke, and B. Stein-
history of Sumba Island, Indonesia, since the Late berger, 2006, Large igneous provinces generated from
Cretaceous and its rapid escape into the forearc in the margins of the large low-velocity provinces in the
the Miocene: Journal of Asian Earth Sciences, v. 19, deep mantle: Geophysical Journal International, v. 167,
p. 453–479. p. 1447–1460.

13880_ch02_ptg01_039-084.indd 83 10/27/15 10:03 AM


84  ALTAMIRA and Burke

Toussaint, J. F., and J. J. Restrepo, 1989, Acreciones sucesivas in W. E. Bonini, R. B. Hargraves, and R. Shagam,
en Colombia: Un nuevo modelo de evolucion geologica: eds., The Caribbean-South American plate bound-
V Congreso Colombiano de Geologia, v. I, p. 127–146. a r y a n d re g i o n a l t e c t o n i c s : G S A M e m o i r 1 6 2 ,
Trenkamp, R., J. N. Kellogg, J. T. Freymueller, and H. P. Mora, p. 189–252.
2002, Wide plate margin deformation, southern Central Villagomez, D., and R. Spikings, 2013, Thermochronology
America and northwestern South America, CASA GPS ob- and tectonics of the Central and Western Cordilleras of
servations: Journal of South American Earth Sciences, v. 15, Colombia: Early Cretaceous–Tertiary evolution of the
p. 157–171. Northern Andes: ­Lithos, v. 160, p. 228–249.
Unger, L. M., V. B. Sisson, and H. G. Avé-Lallemant, 2005, Wadge, G., and K. Burke, 1983, Neogene Caribbean plate
Geochemical evidence for island-arc origin of the rotation and associated Central American tectonic evolu-
Villa de Cura blueschist belt, Venezuela, in H. G. Avé- tion: Tectonics, v. 2, p. 633–643.
Lallemant, and V. B. Sisson, eds., Caribbean-South Ameri- Wadge, G., and R. MacDonald, 1985, Cretaceous tholeiites
can plate interactions, Venezuela: GSA, Special Paper 394, of the northern continental margin of South America; the
p. 223–249. Sans Souci Formation of Trinidad: Journal of the GSL,
Vallejo, C., R. A. Spikings, L. Luzieux, W. Winkler, D. Chew, v. 142, p. 297–308.
and L. Page, 2006, The early interaction between the Car- Weber, J. C., T. H. Dixon, C. DeMets, W. B. Ambeh,
ibbean Plateau and the NW South American plate: Terra P. Jansma, G. Mattioli, J. Saleh, G. Sella, R. Bilham, and
Nova, v. 18, p. 264–269. O. Perez, 2001, GPS estimate of relative motion between
Van der Hilst, R. D., and P. Mann, 1994, Tectonic implications the Caribbean and South American plates, and geologic
of tomographic images of subducted lithosphere beneath implications for Trinidad and Venezuela: Geology, v. 29,
northwestern South America: Geology, v. 22, p. 451–454. p. 75–78.
Van Melle, J., W. Vilema, B. Faure-Brac, M. Ordonez, H. White, R. V., J. Tarney, A. C. Kerr, A. D. Saunders, P. D.
Lapierre, N. Jimenez, E. Jaillard, and M. Garcia, 2008, Pre- Kempton, M. S. Pringle, and G. T. Klaver, 1999, Modifica-
collision evolution of the Pinon oceanic terrane of SW Ec- tion of an oceanic plateau, Aruba, Dutch Caribbean: Im-
uador: Stratigraphy and geochemistry of the “Calentura plications for the generation of continental crust: L ­ ithos,
Formation”: Bulletin de La Societe Geologique de France, v. 46, p. 43–68.
v. 179, p. 433–443. Wilson, J. T., 1966, Are the structures of the Caribbean and
Velasco-Sanchez, M. L., and K. Mendoza, 2003, Estudio geo- Scotia arc regions analogous to ice rafting? Earth and
logico del miembro Calentura en el flanco oriental de la Planetary Science Letters, v. 1, p. 335–338.
Cordillera Chongon Colonche: Escuela Superior Politec- Wright, J., and S. J. Wyld, 2004, Aruba and Curacao:
nica del Litoral, Guayaquil, 182 p. ­Remnants of a collided Pacific Oceanic Plateau? Initial
Vierbuchen, R., 1984, The geology of the El Pilar fault geologic results from the BOLIVAR Project: American
zone and adjacent areas in northeastern Venezuela, Geophysical Union, Fall Meeting 2004.

13880_ch02_ptg01_039-084.indd 84 10/27/15 10:03 AM

You might also like