You are on page 1of 74

Captulo I

CARACTERSTICAS GEOFSICAS DE LA CORTEZA A LO LARGO DEL


LMITE DE PLACAS
Michael Schmitz, Alan Levander, Fenglin Niu, Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt,
Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

Resumen. Entre los aos 1984 y 2004 se realizaron


estudios ssmicos de gran ngulo en Venezuela con
el propsito de determinar el espesor de la corteza
y las velocidades de propagacin dentro de la
misma. Durante los aos 2003-2005, se instalaron
estaciones sismolgicas temporales en el marco
de los proyectos BOLVAR y GEODINOS, cuyos
datos se analizaron junto con los grabados por
las estaciones sismolgicas de la red sismolgica
nacional aplicando el mtodo de funciones
receptoras. En esta investigacin se analizan los
datos y resultados obtenidos en los diferentes
estudios y se interpretan los resultados principales
y las posibles diferencias en la parte continental.
Los estudios ssmicos corticales de gran ngulo
estn enfocados en 4 perfiles norte-sur que cruzan
el lmite de placas entre el Caribe y Sudamrica
desde la Cuenca de Venezuela en el Norte hasta
las cuencas de antepas en el Sur. El rasgo principal
en todos estos perfiles es una disminucin del
espesor cortical relacionado con los sistemas de
fallas transcurrentes (Oca-Ancn, San-Sebastin,
El Pilar) que van de aproximadamente 35-40 km de
espesor al Sur de estos sistemas de fallas hasta 2530 km al Norte de los mismos. La misma tendencia
del adelgazamiento cortical hacia el Norte se puede
observar en los mapas de espesores corticales
derivados del anlisis de funciones receptoras. En
ambos casos hay una coincidencia satisfactoria
con la disminucin del espesor cortical al Norte del
sistema de fallas transcurrentes en el occidente
y el oriente de Venezuela. Sin embargo, para la
regin centro-norte, la disminucin comienza a
unos 100 km al Sur en la cercana del perfil en
67O, pero tiene el mismo comportamiento en la
parte ms oriental de la Cordillera de la Costa.
Las diferencias existentes entre los resultados
de funciones receptoras de algunas estaciones
sismolgicas, as como entre los modelos ssmicos,
podran relacionarse con la naturaleza de la corteza
inferior en la zona, posiblemente un producto de la
subduccin de la placa Suramericana bajo la placa
Caribe. En el oriente hay una coincidencia entre
los principales rasgos corticales en los diferentes
estudios, con un gran espesor cortical debajo de la
Cuenca de Maturn de ms de 45 km, posiblemente
un producto de una duplicacin de corteza inferior,
as como espesores corticales de 40-45 km en el

Escudo de Guayana. Esto podra ser responsable


de las diferencias observadas en las profundidades
para algunas estaciones sismolgicas. Para explicar
las diferencias entre los estudios de funciones
receptoras, algunos aspectos que se deben
considerar son los perfiles de velocidad usados en
la conversin a profundidad de las funciones de
receptor y la relacin Vp/Vs, as como el efecto de
una interfaz inclinada en lugar de horizontal, sobre
el anlisis de las funciones receptoras.
Palabras clave: Ssmica de gran ngulo, Funciones
receptoras, Corteza, Placa del Caribe, Venezuela.
Extendend Abstract. In the years 1984 to 2004
wide angle seismic measurements were done
in Venezuela aiming at determining the crustal
thickness and the corresponding seismic velocities.
From 2003 to 2005 temporary seismological
stations were installed in the scope of the BOLVAR
and GEODINOS projects (Figure1); whose data
were analyzed together with the data of the national
seismological network using receiver functions
(Ammon et al. 1990). In this work we analyze
the data and results obtained in the different
studies, and the principal results and the possible
differences in the continental part are interpreted.
The seismic wide-angle crustal studies are focused
along 4 north-south profiles, which cross the
Caribbean South America plate boundary from
the Venezuela Basin in the north up to the foreland
basins in the south. The main feature common to
these profiles is a crustal thinning related to the
Oca-Ancn, San Sebastin, El Pilar transcurrent
fault systems, from about 35-40 km south of them
to 25-30 km north of them. The same tendency in
decrease of the crustal thickness is observed in the
maps derived from receiver function analysis. In the
west, a crustal thickness of more than 40 km was
derived from measurements in the year 1985 for
the East Coast of Maracaibo Lake (Figure 2), which
coincide with results from receiver function analysis
(Figure 3), with exception to TERV station, for which
a crustal thickness of only 27-29 km was derived
(Niu et al. 2007; Quinteros et al. 2009). Further
east, along longitude 70W, the main feature is a
crustal thinning related to the Falcn Basin, or to
the Oca-Ancn strike slip fault system (Figures 4

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

and 5; Bezada et al. 2008; Gudez, 2007). For the


Coastal Cordillera of the central region, with the
profile along longitude 67W, a crustal thickness of
34 to 37 km was derived (Figures 6 and 7; Jcome
et al. 2008; Schmitz et al. 2008), with an alternative
interpretation of the deep reflections from the lower
crust, which implies a crust up to 40 km thickness
below the coastal Cordillera (Figure 8; Magnani et al.
2009). Further east, Gudez et al. (2003) obtained
a crustal thickness of 40 km at the coast (Figures 9
and 10). From the receiver function analysis, the 3
seismological stations from the Coastal Cordillera
show values ranging from 25 (TURV) to 45 (BIRV)
km (Figure 11), with a possible second value for the
station BIRV (Figure 12), which coincides roughly
with the values for TURV, and the intracrustal
reflector obtained by Magnani et al. (2009).
Serpentinized material in the lower crust, emplaced
within the collision process between the Caribbean
and South American plates, could be responsible for
the distinct features obtained by the seismological
studies. The crustal structure in eastern Venezuela
was studied along 2 north-south seismic profiles, at
longitudes 65 and 64W, which allowed to derive
crustal thickness of 45 and more than 50 km for the
Maturin Basin, respectively (Figures 13 and 14).
The nature of this deep crust might be associated
with lower crustal material attached during the
subduction process of the Atlantic plate beneath
the Caribbean. The steep step in Moho morphology
suggested by Clark et al. (2008), related to the El
Pilar fault system, might be somewhat smoother
considering alternative interpretations (Figures
15 and 16). At the northern edge of the Guayana
Shield, crustal thickness varies between 39 and 46
km (Figure 17), which is compatible with the depth
values derived from receiver function analysis
(Figures 18 to 20). Close to the coastline, depth
values from receiver functions are slightly lower
than from active seismics, only for the station CRUV
at the coast exist strong differences between the
analysis from Niu et al. (2007) and Quinteros et
al. (2009) (Figure 18). Approaches to integrate the
information derived from active and passive seismic
methods lead generally to good coherence, but with
some differences at distinct locations (Figures 21
and 22). To explain the differences between the
different receiver function analyses, some aspects
to consider could be the velocity profiles used for
the depth conversion, the Vp/Vs relation, as well as
a possible inclination of the discontinuity.
Keywords: Wide angle seismic, Receiver functions,
Crust, Caribbean plate, Venezuela.

INTRODUCCIN
El conocimiento de la estructura de la litosfera es
fundamental para el entendimiento de los procesos
geodinmicos que rigen el desarrollo tectnico de
una regin. La historia reciente de interaccin de las
placas del Caribe y de Sudamrica est marcada por
una colisin oblicua en los ltimos 55 Ma (Dewey
& Pindell, 1986), con un movimiento relativo de la
placa del Caribe de 2 cm/ao respecto a Sudamrica
(Weber et al. 2001), la cual ha originado un sistema
de fallas transcurrentes en el Norte del continente
(Schubert, 1984; Audemard et al. 2000) (Figura 1).
En el marco de los proyectos BOLVAR (Broadband
Ocean-Land Investigations of Venezuela and the
Antilles arc Region) y GEODINOS (Geodinmica
reciente del lmite norte de la placa Sudamericana)
(Levander et al. 2006), se realizaron investigaciones
sismolgicas utilizando fuentes activas y pasivas
con el propsito de determinar las estructuras
litosfricas generadas por la interaccin de las
placas del Caribe y de Sudamrica. Imgenes
previas de esta zona de interaccin, a travs de
mtodos sismolgicos pasivos, fueron obtenidas
principalmente por trabajos tomogrficos con
fuentes locales (Bosch, 1997) o telessmicas (van
der Hilst, 1990; VanDecar et al. 2003) con una
resolucin limitada. Las primeras mediciones
ssmicas de refraccin profunda se realizaron en
la zona noroccidental de Venezuela (Gajardo et al.
1986) en dos perfiles ubicados en la Costa Oriental
del Lago de Maracaibo (COLM). Posteriormente,
se desarrollaron proyectos de ssmica profunda
con fuentes activas en el Escudo de Guayana con
el proyecto ECOGUAY (Schmitz et al. 2002), en
la cuenca oriental de Venezuela con el proyecto
ECCO (Schmitz et al. 2005) y en la zona central
con el proyecto Mar y Tierra (Gudez et al.
2003). Los proyectos ECOGUAY y Mar y Tierra
usaron como fuente de energa voladuras mineras
y registros de caones de aire, respectivamente.
Para complementar la informacin estructural
sobre la conformacin de la litosfera en Venezuela,
se utilizaron estudios gravimtricos para la
determinacin del espesor cortical a travs de
perfiles regionales (Orihuela & Ruiz, 1990; Bosch
& Rodrguez, 1992; Sousa et al. 2005).
Las investigaciones sismolgicas activas y pasivas

Captulo I - Caractersticas Geofsicas de la Corteza a lo Largo del Lmite de Placas

realizadas en el marco de los proyectos BOLIVAR


y GEODINOS en el ao 2004, permitieron
estimar los espesores corticales y los principales
cuerpos litosfricos, as como su dinmica en
el rea de interaccin entre las placas del Caribe

y Sudamrica. En el presente trabajo, as como


in extenso en Schmitz (2011), se analizan los
resultados obtenidos y se discuten las posibles
diferencias existentes entre los resultados derivados
por los diferentes mtodos.

Figura 1. Mapa del sureste de la placa del Caribe con el lmite difuso de placas que va desde el Cinturn Deformado del Sur del
Caribe hasta el lmite norte de las cuencas de antepas (Barinas-Apure; Gurico; Maturn). Las islas de las Antillas Menores son:
A-Aruba; C-Curacao; B-Bonaire; LA-Las Aves; LR-Los Roques; O-Orchila; LB-La Blanquilla; H-Los Hermanos; T-Testigos. CAR LIP es
la Gran Provincia gnea del Caribe; las lneas slidas indican los perfiles marinos de reflexin y los perfiles de refraccin en tierra;
crculos blancos indican OBS a lo largo de los perfiles de refraccin marina; estrellas disparos en tierra; tringulos indican las
estaciones sismolgicas en tierra y hexgonos los OBSIP en mar.
Figure 1. Map of the southeastern Caribbean with the diffuse plate boundary extending from the Southern Caribbean Deformed
Belt to the northern limit of the foreland basins (Barinas-Apure; Gurico; Maturn). The Leeward Antilles islands are: A-Aruba;
C-Curacao; B-Bonaire; LA-Las Aves; LR-Los Roques; O-Orchila; LB-La Blanquilla; H-Los Hermanos; T-Testigos. CAR LIP is the
Caribbean Large Igneous Province; red lines indicate marine reflection profiles and refraction profiles on land, white circles the
marine refraction profiles; stars indicate land shot points, triangles seismological stations on land and hexagons OBSIP in the
Caribbean.

CONTEXTO TECTNICO
Las cadenas montaosas y las cuencas asociadas
en el Norte de Venezuela son el producto de la
interaccin tectnica entre las placas del Caribe
y de Suramrica a nivel litosfrico. Presentamos
primero una visin regional de la interaccin de
ambas placas y, posteriormente, las estructuras y
la geologa de la Cordillera de la Costa y Serrana
del Interior, de las cuencas de Falcn, Gurico y
Maturn y de las Antillas Holandesas y del Escudo
de Guayana.
Interaccin de las placas Caribe y Suramrica
El movimiento actual de la Placa del Caribe con

respecto a Suramrica, casi constante por al menos


55 Ma (Dewey & Pindell, 1986), ha generado
una amplia zona de regmenes tectnicos de
transcurrencia, compresin y extensin (Audemard
et al. 2000). El margen sur de la Placa del Caribe
ha sido considerado por mucho tiempo como una
zona angosta de transcurrencia pura (Molnar &
Sykes, 1969; Prez & Aggarwal, 1981). Trabajos
ms recientes reconocen la subduccin de la placa
ocenica del Caribe en el occidente de Venezuela
debajo del continente suramericano (Bloque de
Maracaibo) (van der Hilst & Mann, 1994). En el
Este de Venezuela, especficamente debajo de la
Serrana del Interior, se presencia un cambio en
la polaridad de la subduccin, de manera de que
la continuacin de la litsfera suramericana, la

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

placa ocenica del Atlntico, subduce debajo de


la placa ocenica del Caribe (Russo & Speed,
1994; Jcome et al. 2003). En este modelo,
el movimiento transcurrente constituye una
componente menor de la convergencia oblicua;
un movimiento transcurrente menor (menor a 150
km) ha sido reportado por Audemard & Giraldo
(1997), en un ambiente de particin de fuerzas con
convergencia oblicua entre las placas del Caribe y
de Suramrica. Datos marinos de refraccin ssmica
y de gravimetra determinaron el Moho debajo de
la Cuenca de Venezuela en la Placa del Caribe a una
profundidad entre 15 y 20 km, siendo esta corteza
ocenica ms gruesa que la normal (Officer et al.
1959; Edgar et al. 1971; Biju-Duval et al. 1978;
Diebold et al. 1981; Case et al. 1990; Diebold
et al. 1999). Basndose en datos ssmicos de
reflexin, perforaciones y observaciones de campo
en complejos gneo-mficos obducidos, Donnelly
et al. (1990) y Donnelly (1994) sugirieron que
esta corteza gruesa fue producto de la erupcin de
grandes flujos de basaltos ocurridos cercano a la
Cresta de Beata alrededor de 85 Ma.
Varias interpretaciones han sido propuestas para el
origen de la Placa del Caribe; la ms ampliamente
aceptada propone que la placa se form en la
regin del Pacifico y fue llevada a su posicin
actual mediante un sistema de fallas transcurrentes,
que involucra desplazamiento lateral de ms de
4000 km (Wilson, 1966; Malfait & Dinkelman,
1972; Pindell & Dewey, 1982; Burke et al. 1984;
Pindell & Barrett, 1990; Pindell et al. 2006). La
otra interpretacin importante sugiere la formacin
ms cercana a su posicin actual, entre las dos
Amricas (Ball et al. 1969; Sykes et al. 1982;
Donnelly, 1985; Meschede, 1998; Meschede &
Frisch, 1998). Orihuela et al. (2012) analizaron las
bandas magnticas en el Caribe Oriental, asociando
el centro de esparcimiento ocenico con un centro
de apertura lenta, tal como lo es el Atlntico.
El Norte de Venezuela
Al menos 4 diferentes eventos geodinmicos han
contribuido a la formacin de las estructuras en
esta regin: 1) Orognico del Paleozoico; 2) Fase
de rifting asociado con la apertura de Pangea en
el Jursico y Cretcico temprano; 3) Desarrollo
diacrnico del margen pasivo desde el Cretcico

(Oeste) hasta el Palegeno (Este); 4) Colisin


oblicua entre el Caribe y Suramrica, que gener el
cinturn de corrimiento de la Cordillera de la Costa
y la cuenca de antepas de Gurico en el Centro y el
cinturn de corrimiento de la Serrana del Interior
Oriental y la Cuenca de antepas de Maturn en el
Este.
Cinturn de corrimiento de la Cordillera de la
Costa y cuenca de antepas de Gurico
Durante el Eoceno Medio, la Placa del Caribe
colision con la parte nor-central de Venezuela
emplazando la Cordillera de la Costa en su lugar.
La carga tectnica asociada con el cinturn de
corrimiento gener la flexin de la litosfera, lo que
permiti la generacin de la cuenca de Gurico,
con 7 km de sedimentos en el Oeste y 7 km en el
Este (Erlich & Barrett, 1990; Jcome et al. 2008).
El cinturn de corrimiento de la Cordillera de la
Costa est compuesto por diferentes provincias
geolgicas que contienen (de Norte a Sur): rocas
bsicas y ultrabsicas del Jursico tardo y Cretcico
temprano, roca del basamento del Precmbrico y
Paleozoico, fragmentos de la corteza inferior y del
manto superior del Jursico y Cretcico, secuencias
volcano-sedimentarias y rocas baslticas a riolticas,
y secuencias de sedimentos molsicos y de flysch
del Cretcico tardo a Paleoceno (Beck, 1986;
Bellizzia, 1986; Navarro et al. 1988; Donnelly et
al. 1990; Ostos, 1990; Giunta et al. 2003).
Cinturn de corrimiento de la Serrana del
Interior Oriental y la Cuenca de antepas de
Maturn
El cinturn de corrimiento de la Serrana del
Interior Oriental, es un cinturn de fallas y
corrimientos con vergencia haca el sureste y
azimut N70E (Rossi, 1985). Est limitado por 4
estructuras tectnicas principales: en el Norte, la
falla transcurrente dextral de El Pilar; en el Sur
el frente de deformacin; y en el Este y Oeste las
fallas transcurrentes de Los Bajos y San Francisco,
respectivamente (Audemard et al. 2000). La
Serrana est compuesta principalmente por rocas
del Cretcico y Terciario depositados en un margen
pasivo (Lilliu, 1990). Estas rocas fueron levantadas
por la convergencia oblicua noroeste-sureste entre
las placas del Caribe y Suramrica.

Captulo I - Caractersticas Geofsicas de la Corteza a lo Largo del Lmite de Placas

La cuenca de antepas de Maturn se extiende


desde el frente de deformacin en el Norte hasta
el ro Orinoco en el Sur, que marca el lmite norte
de los afloramientos del basamento del Escudo
de Guayana (Gonzlez de Juana et al. 1980). La
cuenca, cuyos sedimentos provienen del Sur
(Escudo de Guayana), del Norte (Serrana del
Interior en levantamiento) y del Oeste, limita en el
Oeste con la cuenca de Gurico y en el Este con
el ocano Atlntico. Como cuenca antepas, fue
rellenada con 8-13 km de sedimentos negenos,
acomodados como resultado de la carga de los
paquetes de corrimiento, as como de la topografa
producto de la dinmica de subduccin continental,
que oblig a la litsfera continental suramericana a
flexurarse haca abajo entre el Escudo de Guayana
en el Sur y el Cinturn de corrimiento de la Serrana
del Interior en el Norte (Roure et al. 1994; Jcome
et al. 2003).
Escudo de Guayana
Al Sur de las cuencas de antepas de Maturn
y Gurico, aflora la corteza continental del
Precmbrico en el Escudo de Guayana, compuesta
principalmente de rocas metasedimentarias y
metagneas en la facies de anfibolita y granulita,
que han sido intrusionadas por granitos (FeoCodecido et al. 1984), con edades comprendidas
entre 3.600 y 800 Ma (Gonzlez de Juana et al.
1980). El espesor cortical en el lmite norte del
escudo ha sido determinado entre 38 y 45 km (Niu
et al. 2007 y Schmitz et al. 2002, respectivamente).
Cuenca de Falcn y Antillas Holandesas
La Cuenca de Falcn, ubicada en el noroeste de
Venezuela con una superficie de aproximadamente
36.000 km2, tiene un basamento formado por rocas
emplazadas durante la colisin de las placas del
Caribe y de Suramrica en el Paleoceno a Eoceno
Temprano (Audemard, 1995; Baquero et al. 2009),
posiblemente sobre un basamento metamrfico
de alto grado (Grande & Urbani, 2009). La
cuenca se abri durante el Oligoceno, producto
de extensin que gener originalmente estructuras
elongadas en direccin NE-SO que comunicaron
con la cuenca profunda de Bonaire. En el eje de
la cuenca se observan intrusiones de composicin
basltica (Muessig, 1978) que sugieren, junto con

informacin gravimtrica y ssmica reciente, un


adelgazamiento cortical asociado con la apertura
de la cuenca (Muessig, 1984; Rodrguez & Sousa,
2003; Bezada et al. 2008). El origen de la cuenca
an no est resuelto; primeras teoras sugieren su
formacin como cuenca de traccin en un sistema
de transcurrencia dextral (Muessig, 1984) con la
formacin de las islas de Aruba, Curaao y Bonaire
(Antillas Holandesas) en una zona aledaa durante
el Eoceno. Las islas fueron separadas debido a un
rgimen extensivo de direccin este-oeste, tambin
responsable de la apertura de las cuencas de Falcn
y Bonaire (Boesi & Goddard, 1991; Macellari,
1995). El desplazamiento total a lo largo del sistema
transcurrente dextral de Oca-Ancn, es estimado
en menos de 150 km, lo que no concuerda con la
creacin de estas fallas en un sistema de cuenca
de traccin similar a los de la Cuenca de Falcn
(Audemard & Giraldo, 1997). Por ello, Audemard
(1995) propone un origen de tras-arco para la
apertura de la cuenca, con extensin ms bien nortesur que este-oeste en el Eoceno Tardo; la cuenca de
Falcn fue invertida posterior al Mioceno Medio en
un rgimen tectnico compresivo que la convirti
en un alto estructural.
SSMICA DE
VENEZUELA

GRAN

NGULO

EN

Los perfiles ssmicos de gran ngulo se basan en


mediciones de fuentes activas, generalmente
explosiones qumicas, a lo largo de perfiles que
cruzan las estructuras que se van a analizar, con
receptores ubicados a lo largo de estos perfiles. Se
instalan los equipos independientes a una distancia
entre varios kilmetros hasta cientos de metros
entre ellos, a lo largo de un perfil, para grabar varias
voladuras con explosivo a lo largo del mismo, con
el fin de obtener disparos reversos en las distancias
importantes de anlisis, generalmente cercano a las
distancias crticas de las reflexiones de la interfaz
corteza manto o discontinuidad de Mohorovii
(Moho), que se ubican entre 40 y 100 km de
distancia en la parte continental. Otra modalidad
de registro consiste en la grabacin de seales
de caones de aire comprimido, disparados a lo
largo de perfiles en el mar, en estaciones en tierra
o en estaciones en el fondo del mar. La parte ms
importante del anlisis es la identificacin de las
diferentes fases observadas en la seccin ssmica.

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

Las principales fases observadas en las secciones


ssmicas corticales son: Pg para la refraccin del
basamento cristalino, PmP para la reflexin del
Moho y Pn para la refraccin del manto superior, y
se aplican generalmente velocidades de reduccin
para su mejor reconocimiento. Se utiliza la tcnica
del modelado directo para datos de poca cobertura
o inversin para datos muy densos. Ambas tcnicas
se basan en el mtodo de trazado de rayos, con el
cual se calculan los tiempos de viaje de las ondas a
travs de un modelo bidimensional. Los tiempos de
viaje calculados se comparan con los observados y
se modifica el modelo iterativamente hasta lograr un
ajuste satisfactorio entre los tiempos calculados y
observados (Zelt & Smith, 1992). El resultado final
del modelado lo compone el modelo estructural
de la corteza con las respectivas velocidades de
propagacin.
Desde finales de los aos 1990, en los proyectos
ECOGUAY (Schmitz et al. 2002) y Mar y Tierra
(Gudez et al. 2003), se usaron como fuente de
energa voladuras mineras y registros de caones
de aire, respectivamente. En ambos casos, se
pudieron determinar solamente las caractersticas
principales de la corteza, debido a la poca densidad
de los puntos de registro. En el Escudo de Guayana,
la corteza tiene un espesor de 45 km, con una
velocidad promedio de la corteza de 6,5 km/s y una
corteza inferior de elevada velocidad ssmica entre
20 y 30 km de profundidad (Schmitz et al. 2002).
El espesor de la corteza en la zona centro-norte es
de 36 - 39 km cercano a la costa, disminuyendo
hacia la parte ocenica en la placa del Caribe. La
velocidad promedio de la corteza oscila entre 5,9
y 6,3 km/s, sin que se pueda observar una zona de
elevada velocidad en la corteza inferior (Gudez et
al. 2003).
En el proyecto ECCO se emple un mayor nmero
de equipos de registro y un nmero mayor de
voladuras que permitieron una mejor resolucin
a nivel de corteza superior y el seguimiento de
la discontinuidad de Moho entre el Escudo de
Guayana y la costa en Barcelona (Schmitz et al.
2005). La profundidad mxima de los sedimentos
de la cuenca oriental es de 13 km en la parte norte
del perfil y el espesor de la corteza vara entre 35
km en el Norte y 39 km al Norte del ro Orinoco,
con una disminucin de la velocidad promedio de

la corteza de 6,5 km/s en el Escudo de Guayana a


5,8 km/s en la cuenca oriental.
En el marco de los proyectos BOLVAR y
GEODINOS, se realizaron en el ao 2004 estudios
ssmicos con 4 perfiles norte-sur, combinados de
ssmica de refraccin/reflexin de gran ngulo, en
el Norte de Venezuela entre las longitudes 64O y
70O (Figura 1), con la finalidad de investigar la
geodinmica de la compleja zona de interaccin
entre las placas Caribe-Suramrica (Levander et al.
2006). Se utilizaron disparos de caones de aire en
mar y explosiones qumicas en tierra como fuentes,
550 grabadores Texan, 50 OBS y las estaciones
cercanas a la costa norte de la Red Sismolgica
Venezolana como receptores; los resultados se
analizan ms adelante por regiones.
ANLISIS DE FUNCIONES RECEPTORAS
EN VENEZUELA
Con la instalacin de la Red Sismolgica Nacional
(RSN) modernizada y equipada con sensores de
banda ancha Guralp CMG-40T con una respuesta de
velocidad plana de 1 s a 30 s, se gener la base para
el clculo de funciones receptoras en Venezuela. En
los aos 2003-2005, las 35 estaciones de la RSN
fueron complementadas en el centro-oriente del
pas con 27 estaciones de banda ancha en tierra
(PASSCAL) y 14 OBS en la parte marina (OBSIP),
en el marco del proyecto BOLIVAR (Levander et
al. 2006; Figura 1).
La funcin receptora es una herramienta
matemtica que permite calcular la respuesta
relativa de la estructura del subsuelo debajo de
una estacin sismolgica de tres componentes,
a partir de una serie de tiempos, analizando las
conversiones y las reverberaciones de las ondas
P y S en las estructuras que se encuentran bajo el
receptor (Ammon et al. 1990). La funcin receptora
se obtiene removiendo del registro los efectos de
la fuente, los efectos instrumentales y los efectos
de la trayectoria por el manto. Las reverberaciones
son mejor observadas en eventos con distancias
epicentrales de 30 a 95 (Ammon, 1991), en los
cuales el tiempo de viaje y las distancias recorridas
por las ondas son suficientemente grandes para que
un tren de ondas P, de varios minutos, se registre
antes que la primera onda S. Como el viaje de

Captulo I - Caractersticas Geofsicas de la Corteza a lo Largo del Lmite de Placas

las ondas S es ms lento que el de las ondas P, la


profundidad de la discontinuidad se puede medir
directamente calculndola por la diferencia de los
tiempos de llegada de la onda P y la fase convertida
Ps (conversin de la fase P a Sv en una interfaz,
principalmente Moho), conociendo el modelo de
velocidades. Langston (1979) incluye la informacin
de la fase Ps utilizando la relacin en el dominio
de frecuencia compleja y la transformada inversa
para devolverla al dominio de tiempo. Para su
deconvolucin, utiliz un mtodo de estabilizacin
con el filtro nivel de agua y un filtro gaussiano
pasa-bajos. En estudios recientes, han estimado los
espesores de la corteza por medio de la relacin
Vp/Vs (Zandt & Ammon, 1995; Zhu & Kanamori,
2000; Niu & James, 2002) y los tiempos de llegadas
de las diferentes fases convertidas P a S (PpPs y
PpSs+PsP). El valor calculado para el espesor de
la corteza es altamente dependiente de la relacin
Vp/Vs (Zhu & Kanamori, 2000). Esta ambigedad
puede ser reducida utilizando las dos principales
fases, las cuales suministran un contraste adicional
para poder estimar tanto el espesor H como Vp/Vs.
La relacin Vp/Vs, permite, adems, caracterizar
la composicin de la corteza en profundidad por
medio del parmetro elstico conocido como la
relacin de Poisson.
Aplicando la metodologa de funciones receptoras
en Venezuela, Niu et al. (2007) analizaron 313
eventos telessmicos con distancias epicentrales
entre 30 y 90, y magnitudes (Mw) mayores
a 5. El mapa de espesores corticales resultante
muestra gran variacin en la profundidad de Moho,
desde 16 km bajo la Cuenca de Bonaire hasta
ms de 50 km en los Andes venezolanos y en el
nororiente venezolano (bajo las cuencas de Gurico
y Maturn). En el Cratn de Guayana, el espesor
cortical alcanza los 38 km. Quinteros et al. (2008;
2009) detallaron los anlisis de funciones receptoras
en las estaciones del noroccidente y todo el pas,
respectivamente, y los respectivos resultados sern
discutidos a continuacin junto con los resultados
de la ssmica de gran ngulo por regiones.

ESTRUCTURAS
REGIONES

CORTICALES

POR

Occidente
En la Costa Oriental del Lago de Maracaibo
(COLM; Figura 1), Gajardo et al. (1986) y Castejn
et al. (1986) determinaron un espesor cortical de
43 km con una disminucin importante del espesor
cortical hacia el noroeste. Ambas interpretaciones
coinciden en el espesor cortical de ms de 40 km
(Figura 2); sin embargo, Gudez et al. (2003) no
consideraron las evidencias para una disminucin
del espesor hacia el noroeste tan importante como
la planteada por Castejn et al. (1986), al interpretar
llegadas aproximadamente 1 s posteriores en la
seccin Butaque SE como reflexin del Moho
(PmP).

Figura 2. Seccin de registro de la explosin de Altagracia


(arriba, izquierda) y Butaque (arriba, derecha) de los
perfiles de la Costa Oriental del Lago de Maracaibo, con la
identificacin de fases (celeste) de Castejn et al. (1986) y la
re-interpretacin (rojo) de Gudez et al. (2003). Modelado
bidimensional de trazados de rayos (abajo, izquierda) y
modelo de velocidades (abajo, derecha) del perfil Costa
Oriental del Lago de Maracaibo NO-SE (Gudez et al.
2003). La lnea negra en el modelo de velocidades indica la
configuracin del Moho obtenido por Castejn et al. (1986).
Figure 2. Record section Altagracia (top, left) and Butaque
(top, right) of the profiles of the Costa Oriental del Lago de
Maracaibo, with the phase identification (light blue) after
Castejn et al. (1986), and the re-interpretation (red) of
Gudez et al. (2003). 2D raytracing model (bottom, left) and
velocity model (bottom, right) from the NW-SE Costa Oriental
del Lago de Maracaibo profile (Gudez et al. 2003). The black
line in the velocity model indicates the position of the Moho
after Castejn et al. (1986).

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

Los espesores corticales obtenidos por Quinteros


et al. (2008) por anlisis de funciones receptoras
indicaron valores entre 34 (DABV) y 38 (QARV)
km (Figura 3), aproximadamente 50 km al Este
del perfil ssmico, en coincidencia con los valores
obtenidos por Niu et al. (2007). Las tendencias de los
espesores corticales obtenidos con ambos mtodos
muestran pocas diferencias, considerando que no
coinciden directamente las zonas investigadas.
Unos 100 km ms al Este, a lo largo de la longitud
70O, se realiz un perfil ssmico en el marco de
los proyectos BOLVAR y GEODINOS (Figura
1). A travs de las reflexiones del Moho del
disparo Aracua (Figura 4), se puede comprobar la
disminucin de la profundidad del Moho en Falcn
(Figura 5), planteada por Rodrguez & Sousa
(2003) sobre la base del modelado gravimtrico.

Figura 3. Funciones receptoras apiladas del occidente


(izquierda) y las islas (derecha) de Venezuela (Niu et al.
2007). Los cuadradros negros indican el espesor de Moho
asignado, signos de interrogacin indican un segundo mximo,
la asignacin de la profundidad se realiza en consistencia con
estaciones vecinas. Los valores de espesores a la derecha de
la funcin receptora corresponden a los obtenidos por Niu et
al. (2007) y a la izquierda por Quinteros et al. (2009), con +
y se indican diferencias de 2 o ms km respecto a los valores
de Niu et al. (2007); depth = profundidad. En las islas, la lnea
de referencia corresponde a 30 km de profundidad, en el resto
de las zonas a 40 km.
Figure 3. Stacked receiver functions from Occidente (left) and
the islands (right) of Venezuela (Niu et al. 2007). Square dots
indicate the Moho depths, question marks a second maximum;
depth assignment in accordance with neighboring stations.
Depth values at the right of the receiver functions correspond
to Niu et al. (2007), at the left to Quinteros et al. (2009); +
and indicate differences of 2 or more km with respect to the
values obtained by Niu et al. (2007). For the island stations,
the reference line corresponds to 30 km depth, for all the other
stations it is 40 km.

Figura 4. Seccin ssmica del disparo Aracua (Serrana de


Falcn) hacia Norte (lnea de costa aproximadamente a 45 km
desde donde disminuye claramente la densidad de equipos
de registro) y Sur (final del perfil en Barquisimeto) (Bezada et
al. 2008). La diferencia en la distancia crtica de la reflexin
PmP (32 km al Norte y 55 km al Sur) y los tiempos reducidos
respectivos (5,5 s al Norte y 6,2 s al Sur) es indicativa clara
para la disminucin del espesor cortical al Norte. Offset =
Distancia.
Figure 4. Record section of shot Aracua (Serrana of Falcn)
towards north (coast line approximately at profile km 45; north
of it, station spacing is much higher) and south (profile ends in
Barquisimeto) (Bezada et al. 2008). The difference in the critical
distance of the PmP reflection (32 km to the north and 55 km to
the south) and the respective reduced times (5,5 s to the north
and 6,2 s to the south) give hints to the decrease of the crustal
thickness to the north.

Figura 5. Trazado de rayos (arriba) y ajuste entre tiempos


observados y calculados (abajo) a lo largo del perfil 70O
(Bezada et al. 2008). La distancia de 0 km corresponde a 12N
(aproximadamente costa norte de la pennsula de Paraguan);
el disparo Aracua (Figura 4) est ubicado a 110 km. Se nota
la diferencia en el espesor al Norte y Sur del disparo. Distance
= Distancia
Figure 5. Ray tracing (top) and observed and calculated travel
times (bottom) along profile 70W (Bezada et al. 2008). The
0 distance corresponds to 12N (approximately north coast of
Paraguan peninsula); the shot Aracua (Figure 4) is located at
110 km. See the difference in crustal thickness north and south
of the shot point.

Los modelos presentados por Bezada et al.


(2008; Figura 5) y Gudez (2007) coinciden en
el adelgazamiento cortical al Norte de Falcn,
con algunas diferencias en su localizacin exacta.
Gudez (2007) relaciona el adelgazamiento cortical

Captulo I - Caractersticas Geofsicas de la Corteza a lo Largo del Lmite de Placas

a la falla de Oca-Ancn y compara este fenmeno


con lo observado en otros perfiles que cruzan
este lmite de placas en 67O (Magnani et al.
2009) y en 64O (Clark, 2007), donde se observa
una disminucin del espesor cortical al Norte del
sistema de fallas transcurrentes, a distancias de 40
(64O) a 70 (67O) km al Norte de la falla principal,
una distancia en el mismo orden que la observada
en Falcn (60 km al Norte de la falla). En este caso,
lo excepcional no sera el adelgazamiento de la
corteza, sino el aumento del espesor hacia el Norte
a 35 km (al Norte de la pennsula de Paraguan;
Figura 5), que podra estar relacionado con la
corteza de la pennsula de Paraguan. El bloque
de Aruba, caracterizado por un alto gravimtrico
(Bezada et al. 2008), se ubica unos 40 km ms al
Norte.
En esta regin hay una densidad baja de estaciones
sismolgicas, ya que se encuentra fuera de la zona
cubierta por las estaciones porttiles del proyecto
BOLVAR (Figura 1). La estacin MONV en
la pennsula de Paraguan muestra un valor de
26 km (Niu et al. 2007; Quinteros et al. 2009;
Figura 3), que es unos 8 km menos que en el
perfil ssmico. Este valor podra ser influenciado
por el adelgazamiento cortical al Sur, donde el
espesor mnimo est en el mismo orden (Figura
5). Directamente encima del adelgazamiento
cortical no existe una estacin sismolgica, pero la
estacin JACV, aproximadamente 100 km al Este,
en el rumbo del adelgazamiento, tiene un valor de
26 (Niu et al. 2007; Figura 3) a 28 (Quinteros et
al. 2009) km. La estacin SIQV al Sur del perfil
ssmico est en el mismo orden de profundidad
(aproximadamente 35 km).
Unos 50 km al Sur del perfil, en la estacin TERV,
se observa un espesor de solo 27 km (Quinteros et
al. 2009; Niu et al. 2007; Figura 3), lo que es 7 - 10
km menor que lo obtenido con la ssmica activa y
en las dems estaciones ubicadas en los Andes. Los
registros de esta estacin (TERV) arrojaron una
funcin receptora bastante clara, derivada, adems,
de una mayor cantidad de eventos en comparacin
con las otras estaciones. Hay que tomar en cuenta
que la estacin se ubica cerca del contacto SurcoAndes, es decir, en el lmite entre la Cadena
Andina y la Cadena Caribe; podra suponerse algn
rgimen de deformacin que sea responsable de

este bajo espesor de corteza. La ubicacin tampoco


coincide directamente con la ubicacin del perfil
ssmico, por lo cual podra considerarse variaciones
laterales del espesor.
Centro-norte
El segundo perfil de los proyectos BOLVAR y
GEODINOS se ubica lo largo de la longitud 67O
(Figura 2). Del anlisis de los disparos en tierra
(Figura 6), se infiere un Moho relativamente plano
con espesores entre 34 y 37 km en la zona cubierta
por las reflexiones (Figura 7). En los registros de
los disparos en tierra, la reflexin del Moho en
direccin norte est caracterizada por un inicio
difuso sin una llegada marcada (Jcome et al.
2008). Hay indicios para una reflexin intracortical
(Pi) proveniente de unos 18 km de profundidad en
la parte sur del cinturn deformado de la Cordillera
de la Costa (Figuras 6 y 7). Una interpretacin
diferente de las reflexiones de la corteza inferior
y Moho de los disparos Ortiz y Calabozo (Figuras
6 y 7; Magnani et al. 2009), considera el inicio
de la reflectividad profunda como reflexiones
provenientes de un reflector que marca el tope de
la corteza inferior con una fuerte inclinacin al Sur.
El Moho es interpretado como zona de terminacin
de la reflectividad de la corteza inferior.

Figura 6. Secciones ssmicas de los disparos Ortiz (arriba)


y Calabozo (abajo), que proporcionan informacin de la
estructura cortical del cinturn deformado de la Cordillera
de la Costa a lo largo del perfil 67O (Jcome et al. 2008).
El lmite norte de los registros marca la lnea de costa. En
ambos registros, la reflexin del Moho en direccin norte est
caracterizada por un inicio difuso sin una llegada marcada. Las
lneas rojas indican la interpretacin del Moho como zona de

10

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

terminacin de la reflectividad de la corteza inferior (Magnani


et al. 2009); Distance = Distancia.
Figure 6. Record sections of shot points Ortiz (top) and
Calabozo (bottom), which provide information of the crustal
structure of the Cordillera de la Costa deformed belt along
profile 67W (Jcome et al. 2008). The northern limit of
the records indicates the coast line. In both records, the
Moho reflection is characterized by a diffuse record without
remarkable onsets. The red lines mark the interpretation of the
Moho as the end of the zone of reflectivity of the lower crust
(Magnani et al. 2009).

Una comparacin de los dos modelos corticales


(Figura 8) muestra que los puntos de reflexin entre
225 y 275 km en la parte sur del perfil (Magnani
et al. 2009) coinciden con el Moho propuesto
por Schmitz et al. (2008), mientras una serie de
reflectores entre km 125 y 175, provenientes de
llegadas observadas de los disparos de caones de
aire en los Texan, parecen reforzar la solucin de
un Moho ms inclinado. El modelo resultante de
velocidades de Magnani et al. (2009) muestra una
disminucin del espesor cortical de 44 km debajo
de la Cordillera de la Costa a 24 km debajo de la
Cuenca de Bonaire, una disminucin de 20 km en
100 km de distancia. Esta disminucin del espesor
cortical est asociada con el sistema de fallas
transcurrentes de San Sebastin.

entre 34 y 37 km en la zona cubierta por las reflexiones en la


parte continental. Modelo alterno de trazado de rayos (abajo)
de Magnani et al. (2009) considerando a un reflector flotante
en la corteza inferior y el Moho a mayor profundidad;
Distance = Distancia; Depth = Profundidad.
Figure 7. Ray tracing (center) and observed and calculated
travel times (top) along profile 67W (Schmitz et al. 2008).
TURV corresponds to the coast line, O y C are shot points Ortiz
and Calabozo, respectively. The model considers a relatively
flat Moho with a thickness between 34 and 37 km in the
continental region with reflection coverage. An alternative ray
tracing model of Magnani et al. (2009) (bottom) considers a
floating reflector in the lower crust and a deeper Moho.

Figura 8. Comparaciones del Moho a lo largo del perfil 67O


utilizando el mtodo de tomografa de reflexin de las fases
PmP (puntos de reflexin como crculos), Moho propuesto por
Magnani et al. (2009) (lnea negra gruesa) y por Schmitz et al.
(2008) (lnea negra ms delgada), dibujado donde difiera el
modelo (modificado de Magnani et al. 2009).
Distance = Distancia; Depth = Profundidad; Preferred
Moho = Moho preferido; First-arrival Moho (7.2, 7.4, 7.6
km/s contours) = Moho de primeras llegadas con isolneas de
7,2, 7,4 y 7,6 km/s; Reflection tomography Moho = Moho de
tomografa de reflexin.
Figure 8. Comparison of Moho along profile 67W using PmP
reflection tomography (reflection points indicated by circles),
Moho proposed by Magnani et al. (2009) (bold black line)
and by Schmitz et al. (2008) (thin black line), plotted where
the model is different (modified from Magnani et al. 2009).
7.27.6 km/s contours from the first-arrival velocity model.

El comportamiento del Moho propuesto por


Schmitz et al. (2008), para el cual coinciden las
profundidades debajo de la Cuenca de Bonaire,
presenta una diferencia considerable debajo de
la Cordillera de la Costa, ya que Magnani et al.
(2009) interpretan la zona debajo de las reflexiones
profundas como material de corteza inferior.
La disminucin del espesor cortical sera ms
moderada, pero an considerable con 10 km en
aproximadamente 120 km de distancia (Figura 8).

Figura 7. Trazado de rayos (centro) y ajuste entre tiempos


observados y calculados (arriba) a lo largo del perfil 67O
(Schmitz et al. 2008). TURV corresponde a la lnea de costa, O
y C marcan los disparos Ortiz y Calabozo, respectivamente. El
modelo considera un Moho relativamente plano con espesores

Unos 100 km hacia el Este, en la parte oriental de


la Cordillera de la Costa (Figura 9), se realizaron
registros de disparos de caones de aire en la
estacin BIRV (Gudez et al. 2003). Se obtuvo
una variacin fuerte del espesor de la corteza de

Captulo I - Caractersticas Geofsicas de la Corteza a lo Largo del Lmite de Placas

40 km (en la costa) a 25 km (unos 80 km hacia el


Norte) (Figura 10), que coincide con los resultados
obtenidos en estudios anteriores (Case, 1990;
Orihuela & Ruiz, 1990; Bosch & Rodrguez, 1992)
y con espesores ligeramente menores que los
valores mximos de profundidad cortical (42 km
debajo de la Cordillera de la Costa), propuestos por
Orihuela & Ruiz (1990).
Del anlisis de funciones receptoras, las 3
estaciones sismolgicas de la Cordillera de la
Costa (TURV, BIRV y se agrega tambin CUPV)
muestran valores muy dispersos, desde 25 km para
TURV, 40 km para BIRV y 43 km para CUPV
(Figura 11; Niu et al. 2007), los que coinciden con
los valores obtenidos por Quinteros et al. (2009),
con excepcin de BIRV con 45 km. Los valores de
BIRV y CUPV son congruentes con los encontrados
por Schmitz et al. (2008) mediante ssmica de gran
ngulo.

Birongo, CL: Lnea de costa.


Figure 10. Ray tracing (left, top) and observed and calculated
travel times (left, bottom) of profile Birongo-north (Gudez et
al. 2003). The velocity model (right) indicates a strong variation
of the crustal thickness 40 km at the coast to 25 km about 80
km further north. The red line identifies area of control of the
Moho reflection (PM); B: Birongo, CL: Coast line.

Figura 12. Espesores corticales y relacin Vp/Vs en la estacin


BIRV, usando los modos Ps, PpPs y PpSs. (izquierda), Ps y PpSs
(centro) y Ps y PpPs (derecha) (Amaz & Rojas, 2009). Aparte
del valor cortical de 45 km, existe un segundo valor de menor
claridad con 30 km.
Figure 12. Crustal thickness and Vp/Vs relation at station BIRV,
using the modes Ps, PpPs and PpSs. (left), Ps and PpSs (center)
and Ps and PpPs (right) (Amaz & Rojas, 2009). Despite the
value of 45 km crustal thickness, a less expressed second value
of 30 km exists.

En la parte oriental de la Cordillera de la Costa,


los datos ssmicos son escasos, hay indicios de un
espesor en el orden de 40 km (Gudez et al. 2003),
confirmado por los valores obtenidos a partir de
funciones receptoras por Niu et al. (2007) (40 km)
y Quinteros et al. (2009) (45 km).

Figura 9. Seccin ssmica del registro de los disparos de


caones de aire de la lnea 9 (CC-01G-09) del Caribe Central
en la estacin BIRV (Birongo) con llegadas de la Pg hasta 55
km y la reflexin del Moho (PM) entre 180 y 76 km (Gudez
et al. 2003).
Figure 9. Record section using the airgun shots from line 9 (CC01G-09) of Caribe Central at the BIRV (Birongo) station with
Pg arrivals up to 55 km and the Moho reflection (PM) between
180 and 76 km (Gudez et al. 2003).

Figura 10. Trazado de rayos (izquierda, arriba) y ajuste entre


tiempos observados y calculados (izquierda, abajo) del perfil
Birongo-norte (Gudez et al. 2003). El modelo de velocidades
(derecha) indica una variacin fuerte del espesor de 40 km
en la costa a 25 km unos 80 km hacia el Norte. La lnea roja
identifica la zona de control de la reflexin en el Moho (PM); B:

A lo largo de la Cordillera de la Costa se presenta un


cambio importante en la anomala de Bouguer, con
valores ligeramente positivos en la parte oriental, y
valores negativos en la parte occidental (Snchez
et al. 2010). Tanto el anlisis de las anomalas de
Bouguer, como de la dinmica de la Cordillera de
la Costa, con quebradas ms sedimentadas en la
parte occidental que oriental (Picard & Goddard,
1975), indican que este cambio est asociado con
el sistema de fallas oblicuas que cortan al sistema
de fallas de San Sebastin (Orihuela et al. 2011).
La parte oriental est sostenida dinmicamente, lo
que implicara una raz cortical menos profunda
que en la parte occidental. Sin embargo, los
resultados de las investigaciones sismolgicas
parecen indicar lo contrario. Indicios en los anlisis
de funciones receptoras, que permitiran identificar
una discontinuidad ms somera (30 km) en la parte
oriental de la Cordillera de la Costa (Figura 12),
ayudaran a conciliar la informacin sismolgica
existente con los modelos geodinmicos.

11

12

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

Una posibilidad de explicar la existencia de material


cortical en la corteza inferior de la Cordillera de
la Costa, sin que todas las discontinuidades se
desarrollen de manera completa, podra ser la
existencia de manto serpentinizado en la posicin
de la actual corteza inferior. Una situacin parecida
se presenta en la corteza inferior del antearco de los
Andes Centrales, donde gran parte del material de
la corteza inferior, con velocidades ssmicas entre
6,9 y 7,7 km/s, es interpretado como material de
origen mantlico (Schmitz et al. 1999). Para las
serpentinitas, al igual que para zonas de fusin
parcial, la relacin Vp/Vs debera ser alta, es decir
mayor a 1,75. El origen de los fluidos, requeridos
para el proceso de serpentinizacin, proviene en los
Andes Centrales de la placa de Nazca en subduccin
y, en especfico, del proceso de transformacin de
los basaltos de la corteza ocenica en eclogita.
La corteza inferior en la Cordillera de la Costa
muestra valores de velocidades ssmicas de 6,5-6,9
km/s (Magnani et al. 2009), ligeramente menores
que los indicados para el antearco en los Andes
Centrales. La relacin Vp/Vs, obtenida para la
estacin sismolgica TURV, es de 1,75 (Amaz
& Rojas, 2009), por lo que esta relacin podra
justificar la presencia de serpentinita en la corteza
inferior. Sin embargo, se presenta el problema de
hidratacin de las rocas; es decir, el origen de los
fluidos requeridos para la hidratacin del manto.
Sorensen et al. (2005) proponen una proveniencia
de una zona de subduccin para el cinturn
eclogtico de la Cordillera de la Costa, por lo que
esta zona potencialmente serpentinizada podra ser
un relicto geolgico de la fase de generacin de la
Cordillera de la Costa y las napas de Villa de Cura.
Dada la predominancia de material de corteza
ocenica en la zona, como por ejemplo las napas
de Villa de Cura (Unger et al. 2005) y partes de
la Cordillera de la Costa (Sorensen et al. 2005),
sometido a exhumacin en un complejo ambiente
de subduccin (Av Lallemant & Sisson, 2005), y
considerando la existencia de material de corteza
inferior obducido al nivel de corteza superior por
el sistema de fallas de San Sebastin (Figura 30;
Magnani et al. 2009), dicho material podra ser
tambin de origen ocenico o de corteza inferior,
tectnicamente puesto en el sitio de corteza inferior
debajo de la Cordillera de la Costa. La diferencia
en la evolucin entre este material y el continente

sudamericano propiamente dicho, podra ser


responsable de la capacidad de deteccin de las
respectivas discontinuidades con los mtodos
sismolgicos. Evidencia de este tipo de procesos
podra ser la reflectividad difusa proveniente de la
corteza profunda de las secciones ssmicas (Figura
6).
Las estaciones sismolgicas de la Serrana del
Interior Central (Figura 11) tienen un promedio
de profundidad de 31 km (Quinteros et al. 2009)
con valores de 34 km para FCPC, 33 (30) km
para JMPC, 29 km para LMPC, y 29 (23) km
para ROPC (valores entre parntesis de Niu et al.
2007). Hacia el Este, dichos valores contrastan
lateralmente con los altos valores (> 40 km) de
las estaciones BIRV y CUPV. Comparado con los
resultados de la ssmica, estn cercanos los valores
para las primeras dos estaciones, pero 5-10 km ms
someros respecto a los ltimos dos valores. Para el
noroeste de la Cuenca de Gurico (BAUV, ULPC)
los espesores son de 42 y 43 km (Niu et al. 2007)
y 38 y 35 km (Quinteros et al. 2009), siendo estos
ltimos ms cercanos a los valores observados en
los estudios de ssmica activa (Magnani et al. 2009;
Schmitz et al. 2008), que indican profundidades del
Moho entre 35 y 38 km respectivamente, para la
parte sur del perfil a lo largo de la longitud 67O.
Oriente
En el Oriente del pas se dispone de informacin
ssmica de dos perfiles norte-sur a lo largo de las
longitudes 65O y 64O, realizados en el marco de
los proyectos BOLVAR y GEODINOS (Figura
1). Del anlisis de los disparos en tierra del perfil
65O (Figura 13), se infiere un Moho de 45 km de
profundidad para la parte norte de la Cuenca de
Maturn, disminuyendo a 30 km hacia la costa del
Caribe en el Norte.

Captulo I - Caractersticas Geofsicas de la Corteza a lo Largo del Lmite de Placas

2008). El lmite norte de los registros marca la costa en CRUV


(Carpano). Puntos de disparo son: J = Jusepn y P = Pericoco.
Distance = Distancia; Depth = Profundidad.
Figure 14. Seismic record sections of shot points Jusepn (top,
left) y Pericoco (top, right), observed and calculated travel
times (bottom, left) and ray tracing (bottom, right) along profile
64W (Schmitz et al. 2008). The northern limit of the records
marks the coastline at CRUV (Carpano).
Shot points are: J = Jusepn y P = Pericoco.

Figura 13. Seccin ssmica del disparo San Mateo (arriba),


ajuste entre tiempos observados y calculados (centro) y trazado
de rayos (abajo) y a lo largo del perfil 65O (Schmitz et al.
2008). El lmite norte de los registros marca la costa en PCRV
(Puerto la Cruz). Puntos de disparo son: SM = San Mateo,
C = Cantaura, ET = El Tigre, L = Limo y CP = Ciudad Piar.
Distance = Distancia; Depth = Profundidad.
Figure 13. Seismic record section of shot point San Mateo (top),
observed and calculated travel times (center) and ray tracing
(bottom) along profile 65W (Schmitz et al. 2008). The northern
limit of the records marks the coastline at PCRV (Puerto la
Cruz). Shot points are: SM = San Mateo, C = Cantaura,
ET = El Tigre, L = Limo y CP = Ciudad Piar.

Unos 100 km ms hacia el Este, en el perfil 64O,


(Figura 14), se observa en la seccin ssmica
del disparo Jusepn la reflexin del Moho a una
profundidad de 40 km. Del disparo de Pericoco,
ubicado en la Cuenca de Maturn, se puede observar
dos reflexiones seguidas, la primera, con muy baja
amplitud, de una profundidad algo mayor de 40
km, y aproximadamente 1 s despus, sigue en la
seccin ssmica una segunda reflexin proveniente
de una profundidad mayor a 50 km.

Figura 14. Secciones ssmicas de los disparos Jusepn (arriba,


izquierda) y Pericoco (arriba, derecha), ajuste entre tiempos
observados y calculados (abajo, izquierda) y trazado de rayos
(abajo, derecha) a lo largo del perfil 64O (Schmitz et al.

La naturaleza de estas dos discontinuidades an est


en discusin. Schmitz et al. (2008) interpretan la
zona entre ambas discontinuidades como material
de corteza inferior que ha sido adherido a la corteza
en el proceso de interaccin de las placas del
Caribe y de Sudamrica. El origen de este material
ha sido atribuido a procesos relacionados con la
subduccin de la placa continental sudamericana
debajo del Caribe (Russo & Speed, 1994; Jcome
et al. 2003), la subduccin de la losa del atlntico
en el noreste (Molnar & Sykes, 1969) o del PaleoCaribe, subducido desde el Norte (Pindell &
Kennan, 2007).
En el anlisis de la seccin ssmica del disparo
Jusepn (Figura 15) de Clark et al. (2008), la
reflexin del Moho viene aproximadamente 1 s
despus de las llegadas del Moho interpretadas por
Schmitz et al. (2008) entre 90 y 70 km de distancia
y 9-10 s (velocidad de reduccin: 8 km/s). Por esta
razn, el Moho muestra en esta parte de la Serrana
del Interior una mayor profundidad que la propuesta
por Schmitz et al. (2008). Por otro lado, Clark et al.
(2008) no interpretan la segunda fase del disparo
Pericoco, por lo que el espesor cortical, planteado
por ellos, para el Norte de la Cuenca de Maturn,
queda ubicado a una profundidad ligeramente
menor (50 km) que el propuesto por Schmitz et
al. (2008) (53 km; Figura 16). En consecuencia, se
sugiere que el salto en la profundidad del Moho,
asociado con el sistema de fallas transcurrentes,
podra ser considerablemente menor que el
planteado por Clark et al. (2008; Figura 16).

13

14

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

(Schmitz et al. 2005), no conlleva a una variacin


importante del espesor cortical. Sin embargo,
debe tomarse en cuenta que los perfiles no tienen
disparos reversos, por lo que no hay control de las
velocidades absolutas ni de la inclinacin de los
estratos.

Figura 15. Seccin ssmica del disparo Jusepn (velocidad de


reduccin: 8 km/s); las fases son identificadas con barras azules
(Pg) y verdes (PmP), cuya altura corresponde al RMS de la
incertidumbre de la respectiva fase (Clark et al. 2008). Las
flechas indican la posicin de la PmP interpretada por Schmitz
et al. (2008) (Figura 14); offset = distancia.
Figure 15. Seismic record section of shot point Jusepn
(reduction velocity: 8 km/s); Examples of Pg (blue) and PmP
(green) picks are overlain, with heights of plus or minus the RMS
pick uncertainty for that phase (Clark et al. 2008). The arrows
indicate the position of the PmP interpreted by Schmitz et al.
(2008) (Figure 14).

Figura 16. Comparaciones del Moho utilizando el mtodo de


tomografa de reflexin PmP (reflexiones como puntos grises) y
Moho de inversin PmP/Pn de modelo estratificado (lnea negra
delgada; Clark et al. (2008). El Moho propuesto por Schmitz
et al. (2008) (lnea negra) est indicado como referencia
(modificado de Clark et al. 2008). Distance = Distancia;
Depth = Profundidad; Tomographic Moho = Moho
tomogrfico; Layered Moho = Moho estratificado.
Figure 16. Comparisons of the Moho using the method of PmP
reflection tomography (reflections as grey points),
and PmP/Pn layered inversion (thin black line; Clark et al.
2008). The Moho proposed by Schmitz et al. (2008) (black line)
is given for reference (modified from Clark et al. 2008).

Ms hacia el Sur, en el lmite norte del Escudo


de Guayana, se realizaron registros a partir de las
voladuras de las minas del Cerro Bolvar (Figura
17). El espesor de la corteza, determinado entre 39
y 46 km, se observa tanto en los registros de las
ondas P como de las ondas S (Schmitz et al. 2002).
An si se considera una correlacin ligeramente
diferente para el registro de Ciudad Piar hacia NNO

En los aos 2003-2005 exista una densa


distribucin de estaciones sismolgicas en esta
regin, gracias a las estaciones porttiles del
proyecto BOLVAR (Figura 1). Cercano a la costa,
los resultados obtenidos por anlisis de funciones
receptoras (Niu et al. 2007; Quinteros et al. 2009;
Figura 11), indican espesores entre 24 y 27 km,
ligeramente menores que los obtenidos por ssmica
de refraccin (Figuras 14 y 16), los cuales oscilan
entre 29 y 34 km de profundidad. Solamente para la
estacin CRUV hay mayores diferencias entre los
resultados de Niu et al. (2007) -24 km- y Quinteros
et al. (2009) -38 km-, respectivamente. Para las
dems estaciones, los resultados de ambos estudios
estn en el mismo orden (Figura 18).

Figura 17. Secciones ssmicas de las voladuras del Cerro


Bolvar en Ciudad Piar haca el Oeste (arriba), hacia el Este
(centro) y haca el NNO (abajo). Se muestran las correlaciones
de las ondas P (izquierda), de las ondas S (con velocidad de
reduccin de 3,46 km/s para las observaciones al Oeste y al
Este; centro), y los modelos 1D resultantes de las ondas P y de
las ondas S, basadas en la relacin Vp = 1,732 x Vs (derecha;
Schmitz et al. 2002). En el caso del perfil registrado al NNO
(abajo), se muestra una correlacin alternativa de la reflexin
del Moho (Schmitz et al. 2005). Los espesores de corteza para
el Norte del Escudo de Guayana varan entre 39 y 46 km.
Offset = distancia; depth = profundidad; velocity = velocidad;
line = lnea; shot = disparo; upper / lower crust = corteza
superior / inferior; mantle = manto; 1-D model from P
velocities = modelo 1D derivado de velocidades P.

Captulo I - Caractersticas Geofsicas de la Corteza a lo Largo del Lmite de Placas

Figure 17. Seismic record sections using the explosions of Cerro


Bolvar in Ciudad Piar towards west (top), east (center) and
NNW (bottom). Correlations of P-waves (left), and S-waves
(with reduction velocity of 3.46 km/s for the observations
to west and east; center), and the resulting 1-D models for
P-waves and S-waves, based on the relation Vp=1.732 x Vs
(right; Schmitz et al. 2002). In the case of the profile to NNW
(below), an alternative correlation of the Moho reflection is
shown (Schmitz et al. 2005). the crustal thicknesss for the north
of the Guayana Shield varies between 39 and 46 km.

Para la Cuenca de Maturn, existe una coincidencia


respecto a los grandes espesores corticales en
la parte oriental, mayores que 45 km, hasta un
mximo de 56 km de profundidad (Niu et al. 2007;
Schmitz et al. 2008; Quinteros et al. 2009; Figura
19). Las diferencias existentes entre los resultados
de funciones receptoras de algunas estaciones
sismolgicas, as como entre los modelos ssmicos
(Figuras 14 y 16), podran relacionarse con
la naturaleza de la corteza inferior en la zona,
posiblemente un producto de una duplicacin de
corteza inferior debido a la subduccin de la placa
Suramericana bajo la placa Caribe (Schmitz et al.
2008). Niu et al. (2007) identifican un segundo
valor de profundidad proveniente del anlisis de las
funciones receptoras (Figura 19), al cual asocian
con el espesor de los estratos sedimentarios, tal
como en el caso de la estacin MNPC, donde
este valor es algo menor de 10 km en la parte sur
de la cuenca, en coincidencia con los espesores
sedimentarios obtenidos por la ssmica de gran
ngulo (Schmitz et al. 2005; 2008).
Los resultados del anlisis de funciones receptoras
del lmite norte del Escudo de Guayana muestran
valores de espesores de corteza entre 38 y 45 km
(Figura 20; Quinteros et al. 2009), en el mismo
orden que los resultados de la ssmica (Figura 17).
Los espesores corticales determinados por Niu et
al. (2007), en promedio son de 4 km menos (Figura
20). Ms hacia el Sur, los valores decrecen, lo
que podra estar en contradiccin con resultados
obtenidos en el lmite sur del Escudo de Guayana (7
ms al Sur), donde Krger et al. (2002) identifican
un espesor cortical de 48-50 km, basado en anlisis
de funciones receptoras. Una de las razones para la
diferencia de los valores en el lmite norte podran
ser los altos valores de la velocidad promedio de
la corteza, que es de 6,5 km/s para el Escudo de

Guayana, decreciendo hacia el Norte en la medida


de que aumentan los espesores de sedimentos en
la Cuenca de Maturn hasta un valor de 6,0 km/s
(Schmitz et al. 2005).

Figura 18. Mapa del espesor cortical en el Norte de Venezuela


elaborado con los resultados del anlisis de funciones
receptoras (Niu et al. 2007) (Arriba) y Quinteros et al. (2009)
(Abajo); Venezuelan Andes = Andes de Mrida; Lower
Paleozoic Orogenic Belt = Cinturn Orognico del Paleozoico
Inferior; Coastal plains = Planicies costeras; profile = perfiles.
Figure 18. Maps showing the crustal thickness in northern
Venezuela based on receiver function analysis by Niu et al.
(2007; top) y Quinteros et al. (2009; bottom).

Figura 19. Funciones receptoras apiladas de las cuencas


antepas de Venezuela (Niu et al. 2007). Los cuadradros negros
indican el espesor de Moho asignado, signos de interrogacin
indican un segundo mximo, la asignacin de la profundidad
se realiza en consistencia con estaciones vecinas. La flecha
indica la base de los estratos sedimentarios en las cuencas.
Los valores de espesores a la derecha de la funcin receptora
corresponden a los obtenidos por Niu et al. (2007) y a la

15

16

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

izquierda por Quinteros et al. (2009), con + y se indican


diferencias de 2 o ms km y con ++ de 10 o ms km respecto
a los valores de Niu et al. (2007); depth = profundidad.
Figure 19. Stacked receiver functions for the foreland basins in
Venezuela (Niu et al. 2007). Square dots indicate the identified
Moho depth, traces with question marks show two peaks and
our identifications are based solely on the consistency with other
stations. The arrow indicates the bottom of the basin sediments.
Depth values at the right of the receiver function correspond
to the ones obtained by Niu et al. (2007), and at the left by
Quinteros et al. (2009), with + and of 2 or more km are
indicated, and with ++ of 10 or more km with respect to the
values obtained by Niu et al. (2007).

Figura 20. Funciones receptoras apiladas del Escudo de


Guayana de Quinteros et al. (2009) (izquierda) y Niu et
al. (2007) (centro y derecha). Los valores de espesores
determinados por Quinteros et al. (2009) que estn indicados
con +, se refieren a espesores mayores de 2 o ms km respecto
a los valores de Niu et al. (2007); depth = profundidad.
Figure 20. Stacked receiver functions for the Guayana Shield
from Quinteros et al. (2009) (left) and Niu et al. (2007) (center
and right). Depth values by Quinteros et al. (2009), with + and
of 2 or more km are indicated, and with ++ of 10 or more
km with respect to the values obtained by Niu et al. (2007).

INTERPRETACIN CONJUNTA DE DATOS


SISMOLGICOS ACTIVOS Y PASIVOS
Un primer intento de integracin de los datos
provenientes de estudios ssmicos de gran
ngulo, y del anlisis de funciones receptoras en
Venezuela, haba sido presentado por Bezada et al.
(2007). Los autores concluyen que los diferentes
mtodos sismolgicos aplicados al estudio del
lmite de placas Caribe-Suramrica proporcionan
una multitud de datos que no siempre son
consistentes entre s. En el mapa (Figura 21) se
integran los resultados consistentes entre ssmica
de gran ngulo y anlisis de funciones receptoras
provenientes del estudio de Niu et al. (2007),
excluyendo del mapa las funciones receptoras costa
afuera, que estn en clara contradiccin con los 4
perfiles de refraccin / reflexin de gran ngulo. El
mismo principio se aplica a las estaciones ubicadas

en la Cordillera de la Costa y sobre las cuencas


sedimentarias. Aun as, se identifican algunas zonas
donde existen inconsistencias, como por ejemplo
la estacin TERV en el occidente, al Sur del perfil
70O, que permite una interpolacin del occidente
de la Serrana del Interior y la parte norte de los
Andes con espesores corticales alrededor de 30
km. En el oriente, a lo largo del perfil 65O, los
espesores corticales alrededor de 30 km (aprox. 10
km menores que al Oeste y al Este) provienen de
los resultados ssmicos y podran reflejar la menor
expresin topogrfica de la Serrana del Interior
en esta regin. Los resultados ssmicos a lo largo
del perfil 64O se comparan con los resultados del
anlisis de funciones receptoras, y se colocan en
el contexto de los resultados del anlisis de ondas
superficiales (Figura 22), utilizando los valores del
mapa (Figura 21) como base.
Los rasgos principales de las estructuras corticales
obtenidos por los estudios ssmicos (Clark et al.
2008) coinciden con los resultados de las funciones
receptoras (Figura 22). Estudios de fase y amplitud
de ondas Rayleigh aportan datos adicionales sobre
la naturaleza de la corteza y el manto litosfrico. La
disminucin del espesor cortical entre la Serrana
del Interior en el Sur y la placa del Caribe en el
Norte parece tomar lugar ms hacia el Sur en los
datos provenientes de funciones receptoras que en
los datos ssmicos (Figura 22).
Miller et al. (2009) aplicaron una metodologa que
analiza velocidades de onda Rayleigh que varan de
forma lateral y azimutal. El modelo tomogrfico de
la zona de estudio consiste en una grilla de puntos
a los que se asignan valores de velocidad que son
sopesados por una funcin de distribucin gaussiana
y el frente de ondas proveniente de cada evento es
representado por dos ondas planas y su respectiva
fase, amplitud y direccin. Las variaciones de fase
y amplitud son aproximadas por esta representacin
y pueden ser invertidas. Con el anlisis de las ondas
superficiales, utilizando un rango de frecuencias de
onda Rayleigh de 5 a 50 mHz (perodo de 20 a 200
s) provenientes de 49 eventos de magnitud igual o
mayor a 5,1 a distancias epicentrales de 20 a 60,
muy parecido al utilizado por Miller et al. (2009).
Se logr obtener una imagen del manto superior
hasta profundidades de aproximadamente 200
km (Bezada et al. 2007; Figura 22), que muestra

Captulo I - Caractersticas Geofsicas de la Corteza a lo Largo del Lmite de Placas

el modelo de ondas superficiales en combinacin


con los resultados de funciones receptoras para
definir el espesor de la corteza. En la realizacin

del modelo se utiliz como control la profundidad


de Moho obtenida del mapa de espesores corticales
(Figura 21).

Figura 21. Espesor cortical obtenido a travs de la interpolacin de una combinacin de observaciones de ssmica activa y
funciones receptoras (Bezada et al. 2007). Los valores de funciones receptoras en las cuencas sedimentarias han sido excluidos de
este mapa y slo se muestran las estaciones sismolgicas cuyos resultados hayan sido considerados para el mapa integrado. Active
source = ssmica con fuentes activas; receiver function = estaciones con anlisis de funciones receptoras; la leyenda muestra el
espesor cortical en km.
Figura 21. Crustal thickness obtained from the combined interpolation of active seismic and receiver function observations (Bezada
et al. 2007). The values from the receiver functions in the sedimentary basins have been removed from the map, and only the
seismological stations that have been used for the integrated map
are shown (white dots).

Figura 22. Funciones receptoras apiladas en un perfil a lo


largo de la longitud 64O, desde la Cuenca de Venezuela en
el Norte hasta el Escudo de Guayana en el Sur (Bezada et al.
2007), incluyendo ssmica activa, funciones receptoras (arriba)
y velocidades de ondas superficiales como perturbaciones
(centro) y valores absolutos (abajo). Distance = distancia; depth
= profundidad; velocity = velocidad; latitude (deg North)
= latitud (grados norte); Root of Guayana Shield = raz del
Escudo de Guayana; Aves Ridge = Cuesta de Aves; subducted
Atlantic plate = place atlntica subducida; plate boundary
zone = zona de lmite de placas; deformed SA mantle = manto
sudamericano deformado.
Figure 22. Stacked receiver functions along a profile at 64W,
from the Venezuela Basin in the north to the Guayana Shield
in the south (Bezada et al. 2007), including receiver functions
plotted on results from active seismic(top), velocities from
surface waves as pertubations (center), and absolute values
(bottom).

17

18

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

DISCUSIN Y CONCLUSIONES
Se analizan los resultados de estudios ssmicos
corticales de gran ngulo realizados desde el ao
1984 hasta la fecha en Venezuela, en conjunto con
los resultados del anlisis de funciones receptoras
en las tres regiones: occidente, centro-norte y
oriente. Las estructuras obtenidas no siempre son
consistentes entre s, lo cual aplica tanto para la
comparacin de los resultados de los diferentes
mtodos (ssmica de gran ngulo y anlisis de
funciones receptoras) como para la comparacin de
los datos de diferentes autores utilizando el mismo
mtodo.
Basado en resultados de los proyectos GEODINOS
y BOLVAR, y de estudios previos de refraccin
profunda, se ha cartografiado el espesor cortical en
el Norte de Venezuela (Figura 22). En general, el
espesor cortical en el Norte de Venezuela se reduce
de aproximadamente 40 km al Sur del Sistema
Montaoso del Caribe a menos de 35 km en la
lnea de costa, mientras que excede los 35 km en
la Cordillera de la Costa (Gudez et al. 2003). Un
pronunciado adelgazamiento cortical se observa
en la Cuenca de Falcn a los 70W (ver tambin
Rodrguez & Sousa, 2003). Por otro lado, se
encontr una raz cortical profunda (ms de 50 km
de espesor cortical) al este de 64W en la Cuenca
Oriental.
El anlisis detallado de los perfiles que cruzan el
lmite de placas Caribe Sudamericana (Guedez,
2007; Christeson et al. 2008; Clark et al. 2008;
Magnani et al. 2009; Bezada et al. 2010), realizado
en el marco del proyecto BOLVAR, permiti
definir la transicin entre la corteza ocenica
en la Cuenca de Venezuela hacia la corteza del
continente sudamericano. Destacan estructuras de
alta velocidad asociadas con el sistema de fallas
transcurrentes (Magnani et al. 2009), as como
saltos importantes en el espesor cortical al Norte
de la costa venezolana (Magnani et al. 2009; Clark
et al. 2008). En el caso del perfil ms oriental
(64O), el salto propuesto por Clark et al. (2008)
es considerablemente mayor que el propuesto
por Schmitz et al. (2008), pero ambos modelos
coinciden en el valor del espesor cortical al Norte
del sistema de fallas transcurrentes, donde no se
observa un aumento significativo en la zona del

arco de islas extinto (Figura 16). En el perfil ms


occidental a 70O, la disminucin del espesor
cortical al Norte de la falla de Oca-Ancn ha sido
interpretado como un adelgazamiento cortical
asociado con la formacin de la cuenca de Falcn
(Bezada et al. 2007), pero podra re-interpretarse,
considerando la disminucin relacionada con el
sistema de fallas de Oca-Ancn, y el aumento hacia
el Norte relacionado con el arco extinto de las islas
de las Antillas de Sotavento.
Por otra parte, Quinteros et al. (2009) utilizaron
telesismos registrados en 20 estaciones de la
red sismolgica nacional y 25 estaciones del
proyecto BOLVAR, analizando ms de 75
eventos telessmicos, con distancias epicentrales
que variaron entre 27 y 90, y magnitudes (Mw)
mayores a 5. La corteza tiene su mayor espesor en
la regin de la Cuenca de Maturn, al oriente de
Venezuela, con valores de hasta 56 km de espesor,
mientras que los menores valores de espesores
corticales se observan al Norte de Venezuela en
direccin hacia el Caribe, con valores de hasta 18
km, aproximadamente, caractersticos de corteza
ocenica. En la regin insular se obtuvieron valores
de profundidad que varan entre 18 y 29 km.
En la serrana de Falcn el promedio del espesor
cortical obtenido es de 36 km, observando una
disminucin en la cuenca de Falcn (28 km en la
estacin JACV). Llama la atencin el resultado
de la estacin TERV, ubicada al extremo norte
de los Andes, con una profundidad de 27 km,
considerablemente menor de lo esperado.
La Cordillera de la Costa en el centro-norte
muestra espesores variables entre 25 km (TURV)
y 44 km (BIRV) con diferencias importantes en
el anlisis de las estructuras de la corteza inferior,
donde se observan reflexiones de gran ngulo,
que son interpretadas por Magnani et al. (2009)
como reflexiones del tope de la corteza inferior,
mientras Schmitz et al. (2008) las interpretan como
reflexiones del Moho. Consecuentemente, resultan
diferencias importantes en el espesor cortical al Sur
del sistema de fallas de San Sebastin, con espesores
mayores a 40 km derivados por Magnani et al.
(2009) y espesores entre 34 y 37 km segn Schmitz
et al. (2008). Los espesores derivados del anlisis
de funciones receptoras en la estacin TURV por

Captulo I - Caractersticas Geofsicas de la Corteza a lo Largo del Lmite de Placas

Niu et al. (2007) con 25 km y Quinteros et al.


(2009) con 24 km, son consistentes entre s, pero
ms de 10 km menores que los espesores obtenidos
por los estudios de ssmica activa. Del anlisis de la
estacin ms prxima hacia el Este, BIRV, ubicada
en el Este de la Cordillera de la Costa, se obtuvo
un espesor de 40 y 45 km, respectivamente, lo que
estara en el orden de espesores obtenidos en la
ssmica (Schmitz et al. 2008). El valor del espesor
para TURV de 25 km podra interpretarse como el
tope de la corteza inferior propuesto por Magnani
et al. (2009). Su anlogo tendra un segundo valor
identificado en los registros de la estacin BIRV
con 30 km de espesor (Amaz & Rojas, 2009), que
podra ser igualmente interpretado como tope de la
corteza inferior. Si se interpreta la discontinuidad
profunda como Moho, debe preguntarse sobre
la naturaleza de la zona suprayacente, que est
caracterizada por velocidades ssmicas corticales
(6,5 6,9 km/s).
Una posibilidad de explicar la existencia de
material de caractersticas corticales en la corteza
inferior de la Cordillera de la Costa, sin que todas
las discontinuidades se desarrollen de manera
completa, podra ser la existencia de manto
serpentinizado en la posicin de la actual corteza
inferior. Una situacin parecida se presenta en
la corteza inferior del antearco de los Andes
Centrales, donde gran parte del material de la
corteza inferior, con velocidades ssmicas entre 6,9
y 7,7 km/s, es interpretado como material de origen
mantlico (Schmitz et al. 1999). La corteza inferior
en la Cordillera de la Costa muestra valores de
velocidades ssmicas de 6,5-6,9 km/s, ligeramente
menores que los indicados para el antearco en los
Andes Centrales. La relacin Vp/Vs, obtenida para
la estacin sismolgica TURV, es de 1,75 (Amaz
& Rojas, 2009), por lo que esta relacin podra
justificar la presencia de serpentinita en la corteza
inferior; relicto geolgico de la fase de generacin
de la Cordillera de la Costa y las napas de Villa de
Cura (Sorensen et al. 2005).
El afloramiento de rocas con alta relacin
P/T (eclogitas y otras rocas con alto grado de
metamorfismo) en este cinturn montaoso sugiere
que stas pudiesen ser asociadas con la tectnica
relacionada con el sistema de fallas de San
Sebastin, que transporta material de alta velocidad

ssmica (Vp > 7 km/s) a niveles de corteza superior


(Magnani et al. 2009). Segn el desarrollo tectnico
de la Cordillera de la Costa, se esperara un mayor
espesor cortical para la parte occidental que para la
parte oriental. Con los datos corticales disponibles
a la fecha no se puede confirmar dicha situacin.
La complejidad tectnica descrita puede ser
responsable de la variacin del espesor cortical
obtenido en los diferentes estudios, ya que cada
mtodo puede ser susceptible a otra configuracin
en la corteza. Comparaciones entre estudios
ssmicos y de funciones receptoras en los Andes
Centrales (ANCORP Working Group, 1999; 2003;
Schmitz et al. 1999; Yuan et al. 2000; 2002; Sick
et al. 2006; Heit et al. 2008; Woelbern et al. 2009)
muestran variaciones en el orden de 10-15 km para
la regin de la Cordillera Oriental con una tectnica
compleja.
Para la Serrana del Interior en el oriente se observa
una diferencia entre los espesores corticales de
la parte central y la oriental. La primera present
un promedio de profundidad de Moho de 31
km, mientras que la seccin oriental alcanz un
valor promedio de 36,5 km, con la excepcin de
la estacin PCRV cuyo valor de profundidad es
de 24,5 km. Diferentes interpretaciones de las
reflexiones del Moho debajo de la Serrana del
Interior implican diferencias en el espesor cortical
al Sur del sistema de fallas de El Pilar (Clark et
al. 2008; Schmitz et al. 2008), pero coinciden los
espesores propuestos para la parte marina, al Norte
del sistema de fallas. En la Cuenca Oriental, los
valores varan entre 32 km en la sub-cuenca de
Gurico y un mximo de 56 km en la sub-cuenca
de Maturn. El Escudo de Guayana posee valores
uniformes para la discontinuidad de Moho que se
promedian en 41 km.
Las diferencias existentes entre los resultados
de funciones receptoras de algunas estaciones
sismolgicas, as como entre los modelos ssmicos,
podran relacionarse con la naturaleza de la corteza
inferior en la zona, posiblemente un producto
de una duplicacin de corteza inferior debido a
la subduccin de la placa Suramericana bajo la
placa Caribe. Para explicar las diferencias entre
los estudios de funciones receptoras, algunos
aspectos que se deben considerar son los perfiles de
velocidad usados en la conversin a profundidad de

19

20

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

las funciones de receptor y la relacin Vp/Vs, as


como el efecto de una interfaz inclinada en lugar
de horizontal, sobre el anlisis de las funciones
receptoras.
AGRADECIMIENTOS
Para la realizacin de las mediciones ssmicas de
gran ngulo en los aos 1998 2004 agradecemos
la participacin de un gran nmero de voluntarios,
personal de campo de varias instituciones y
universidades de Venezuela (UCV, USB, ULA) y
EEUU (Rice University, Indiana University, UCLA
San Diego, entre otras), los equipos del IRISPASSCAL Instrument Center, las tripulaciones a
bordo de los buques de investigacin R/V Maurice
Ewing y R/V Seward Johnson II, as como personal
de CVG Ferrominera Orinoco CA. Las discusiones
con F. Urbani, S. Grande, F. Audemard, N.
Orihuela, C. Len y A. Singer sobre la naturaleza
de la corteza en la Cordillera de la Costa han
contribuido al anlisis crtico de la informacin
existente en esta regin. Dems miembros del
grupo de trabajo de ssmica activa del proyecto
BOLVAR son: C.A. Zelt, D. Sawyer, (U Rice); V.
Rocabado, J. Snchez (FUNVISIS); G. Christeson,
(UTIG); M.B. Magnani (U Memphis); P. Mann (U
Houston), A. Escalona (U Stavanger). Este trabajo
es una contribucin a los proyectos G-2002000478,
PDVSA-INTEVEP FUNVISIS - 04-141 y NSF
Continental Dynamics Program EAR 0003572 y
EAR 0607801.
REFERENCIAS
Amaz, R. & Rojas, K. (2009). Estimacin del espesor
de corteza desde la Cuenca de Venezuela hasta el
Escudo de Guayana, a travs del anlisis de funciones
receptoras. Trabajo especial de grado. Indito.
Universidad Central de Venezuela, Caracas, 167 pp.
Ammon, C.J., Randall, G.E., Zandt, G., (1990). On the
No uniqueness of Receiver Function Inversions. J.
Geophys. Res., 95, 15303-15318.
Ammon, C. J. (1991). The isolation of receiver effects
from teleseismic P waveforms. Bull. Seismol. Soc.
Am., 81, 2504-2510.
ANCORP WORKING GROUP (Oncken, O., Lschen,
E., Mechie, J., Sobolev, S., Schulze, A., Gaedicke,
C., Grunewald, S., Bribach, J., Asch, G., Giese, P.,
Wigger, P., Schmitz, M., Lth, S., Scheuber, E.,
Haberland, C., Rietbrock, A., Gtze, H.-J., Brasse,

H., Patzwahl, R., Chong, G., Wilke, H.-G., Gonzlez,


G., Jensen, A., Araneda, M., Vieytes, H., Behn, G.,
Martnez, E., Amador, J., Ricaldi, E., Chumacero,
H., Luterstein, R. (1999). Seismic reflection
image revealing offset of Andean subduction-zone
earthquake locations into oceanic mantle. Nature,
397, 341-344.
ANCORP-WORKING GROUP. (2003). Seismic
imaging of a convergent continental margin and
plateau in the central Andes (Andean Continental
Research Project 1996, ANCORP 96). J. Geophys.
Res., 108, doi: 10.1029/2002JB001771.
Audemard, F.A. (1995). La Cuenca Terciaria de Falcn.
Venezuela Noroccidental: Sntesis Estratigrfica,
Gnesis e Inversin Tectnica, IX Congreso
Latinoamericano de Geologa, Caracas, Venezuela
(en CD).
Audemard, F.A. & Giraldo, C. (1997). Desplazamientos
dextrales a lo largo de la frontera meridional de la
Placa Caribe, Venezuela Septentrional. VIII Congreso
Geolgico Venezolano, Tomo I, Maracaibo, pp. 101
108.
Audemard, F.A., Machette, M.N., Cox, J.W., Dart,
R.L., Haller, K.M. (2000). Map and database of
Quaternary faults in Venezuela and its offshore
regions. US Geological Survey, Open-file report 00018, 72 pp. + mapa.
Av Lallemant, H.G. & Sisson, V.B. (2005). Exhumation
of eclogites and blueschist in northern Venezuela:
Constraints from kinematic analysis of deformation
structures. In: H.G. Av Lallemant, V.B. Sisson
(Eds.), Caribbean-South America plate interactions,
Venezuela. GSA Special Paper 394, pp. 193-206.
Ball, M., Harrison, C., Supko, P. (1969). Atlantic
opening and the origin of the Caribbean. Nature, 223,
167168.
Baquero, M., Acosta, J., Kassabji, E., Zamora, J.,
Sousa, J.C., Rodrguez, J., Grobas, J., Melo, L.,
Schneider, F. (2009). Polyphase development of the
Falcn Basin in northwestern Venezuela: Implications
for oil generation. En: K.H. James, M. A. Lorente &
J.L. Pindell (eds). The Origin and Evolution of the
Caribbean Plate. Geological Society, London, Special
Publications, 328, pp. 587-612.
Beck, C. (1986). Caribbean colliding. Andean drifting
and the MesozoicCenozoic geodynamic evolution of
the Caribbean, VI Congreso Geolgico Venezolano,
Caracas, Venezuela, 10, pp. 163182.
Bellizzia, A. (1986). Sistema Montaoso del Caribe, una
cordillera alctona en la parte norte de Amrica del
Sur. VI Congreso Geolgico Venezolano, Caracas,
Venezuela, 10, pp. 66576836.
Bezada, M.J., Miller, M., Niu, F., Pavlis, G., Zelt,
C., Schmitz, M., Rendn, H., Levander, A., y EL

Captulo I - Caractersticas Geofsicas de la Corteza a lo Largo del Lmite de Placas

BOLVAR WORKING GROUP. (2007). Estructura


ssmica de la corteza y manto superior en el lmite
de placas Caribe-Suramrica: Un enfoque integrado.
IX Congreso Geolgico Venezolano, Caracas, 12 pp.
Bezada, M.J., Schmitz, M., Jcome, M.I., Rodrguez, J.,
Audemard, F.A., Izarra, C. And The Bolvar Active
Seismic Working Group. (2008). Crustal structure in
the Falcn Basin area, northwestern Venezuela, from
seismic and gravimetric evidence. J. Geodyn., Vol.
45, 191-200, doi:10.1016/j.jog.2007.11.002.
Bezada M. J., Magnani M., Zelt C. A., Schmitz, M.,
Levander A., 2010. The CaribbeanSouth American
plate boundary at 65W: Results from wide angle
seismic data. J. Geophys. Res., Vol. 115, B08402,
doi:10.1029/ 2009JB007070.
Biju-Duval, B., Mascle, A., Montadert, L., Wanneson,
J. (1978). Seismic investigations in Colombia,
Venezuela and Grenada Basin, and on the Barbados
Ridge for future IPOD drilling. Geol. Mijnb. 57,
105116.
Boesi, T. & Goddard, D. (1991). A new geologic model
related to the distribution of hydrocarbon source rocks
in the Falcn Basin, Northwestern Venezuela. In:
Biddle, K.T. (Ed.), Active Margin Basins. American
Asociation of Petroleum Geologists Memoir, 52,
Tulsa, USA, pp. 303319.
Bosch, M. (1997). P wave velocity tomography of
the Venezuelan region from local arrival times. J.
Geophys. Res., Vol. 102, 5455-5472.
Bosch, M. & Rodrguez, I. (1992). North Venezuelan
collisional crustal block: the boundary between the
Caribbean and South American plates. J. South Am.
Earth Sci., Vol. 6, 133-143.
Burke, K., Cooper, C., Dewey, L.F., Mann, P., Pindell,
J. (1984). Caribbean tectonics and relative plate
motions. In: Bonini, W.E., Hargraves, R.B., Shagan,
R. (Eds.), The CaribbeanSouth American Plate
Boundary and Regional Tectonics. Geol. Soc.Am.,
Boulder Colorado, pp. 3161. M.162.
Case, J.E., Macdonald, W., Fox, P.J. (1990). Caribbean
crustal provinces, seismic and gravity. In: Dengo, G.,
Case, J.E. (Eds.), The Caribbean Region. Geol. Soc.
Am., Boulder Colorado, pp. 1536, H.
Castejn, B., Marquz, C., Urbez, M. (1986). Modelo
de corteza en la Costa Oriental del Lago de Maracaibo.
Trabajo especial de grado. Indito. Universidad
Central de Venezuela, Caracas, 208 pp.
Christeson, G.L., Mann, P., Escalona, A., Aitken,
T. (2008). Crustal structure of the Caribbeannortheastern South America arc-continent collision
zone, J. Geophys. Res., Vol. 113, B08104, doi:
10.1029/2007JB005373.
Clark, S.A. (2007). Characterizing the Caribbean-South
American Plate Boundary at 64W. Ph. D. Thesis.

Unpublished. Rice University, Houston, Texas, 81 pp.


Clark, S.A., Zelt, C.A., Levander, A., Magnani,
M.B. (2008). Characterizing the CaribbeanSouth
American plate boundary at 64 W using wideangle seismic data, J. Geophys. Res., 113, B07401,
doi:10.1029/ 2007JB005329.
Dewey, J.F. & Pindell, J.L. (1986). Neogene block
tectonics of eastern Turkey and northern South
America; continental applications of the finite
difference method; discussion and reply. Tectonics,
5(4), 703705.
Diebold, J.B., Stoffa, P.L., Buhl, P., Truchan, M.
(1981). Venezuela basin crustal structure. J. Geophys.
Res. 86, 79017923.
Diebold, J.B., Driscoll, N., Ew-9501 Science Team.
(1999). New insights on the formation of the
Caribbean basalt province revealed by multichanel
seismic images of volcanic structures in the
Venezuelan Basin. In: Mann, P. (Ed.), Caribbean
Basins. Sedimentary basins of the world, Amsterdam,
pp. 561589.
Donnelly, T.W. (1985). Mesozoic and Cenozoic plate
evolution of the Caribbean region. In: Stehli, F.G.,
Webb, S.D. (Eds.), The Great American Biotic
Interchange, pp. 89121.
Donnelly, T.W. (1994). The Caribbean Sea Floor.
In: Donovan, K., Jackson, T.A. (Eds.), Caribbean
Geology, and Introduction, pp. 4164.
Donnelly, T.W., Beets, D., Carr, M., Jackson, T.,
Klaver, G., Lewis, J., Maury, R., Schellenkens,
H., Smith, A., Wadge, G., Westercamp, D. (1990).
History and tectonic setting of Caribbean magmatism.
In: Dengo, G., Case, J.E. (Eds.), The Caribbean
Region. Geol. Soc. Amer., Boulder, Colorado, pp.
339374, H.
Edgar, N.T., Ewing, J.I., Hennion, J. (1971). Seismic
refraction and reflection in the Caribbean Sea. Am.
Assoc. Petrol. Geol. Bull. 55, 833870.
Erlich, R.N. & Barrett, S.F. (1990). Cenozoic plate
tectonic history of the northern VenezuelaTrinidad
area. Tectonics 9, 161184.
Feo-Codecido, G., Smith, F.D., Aboud, N., Di Giacomo,
E. (1984). Basement and Paleozoic rocks of the
Venezuela Llanos Basin. In: Bonini, W.E., Hargraves,
R.B., Shagan, R. (Eds.), The CaribbeanSouth
American Plate Boundary and Regional Tectonics.
Geol. Soc. Amer., Boulder, Colorado, pp. 175187.
M.162.
Gajardo, E., Nicolle, J.L., Castejn, B. Marquz, C.,
Urbez, M. (1986). Modelo de corteza en la Costa
Oriental del Lago de Maracaibo. III Congr. Venez. de
Geofsica, Caracas, pp. 102-111.
Giunta, G., Marroni, M., Padoa, E., Pandolfi, L.
(2003). Geological constraints for the geodynamic

21

22

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

evolution of the southern margin of the Caribbean


plate. In: Bartolini, C., Buffer, R.T., Blickwede,
J. (Eds.), The Circum-Gulf of Mexico and the
Caribbean: Hydrocarbon habitats, basin formation,
and plate tectonics. Amer. Ass. Petrol. Geol., M 79,
pp. 104125.
Gonzlez De Juana, C., Arozena, J., Picard, X. (1980).
En: Foninves (Ed.), Geologa de Venezuela y de sus
Cuencas Petrolferas, 2, pp. 95994. Caracas.
Grande, S. & Urbani, F. (2009). Presence of highgrade rocks in NW Venezuela of possible Grenvillian
affinity. En: K.H. James, M.A. Lorente & J.L.
Pindell (eds). The Origin and Evolution of the
Caribbean Plate. Geological Society, London, Special
Publications, 328, pp. 533-548.
Gudez, M.C. (2007). Crustal Structure across the
Caribbean-South American Plate Boundary at 70W
- Results from seismic refraction and reflection data.
MSc. Thesis, Unpublished. Rice University, Houston,
Texas.
Gudez, R., Schmitz, M., Cavada, J., Snchez, J. (2003).
Determinacin del espesor cortical y velocidades
ssmicas en el rea centro-norte de Venezuela. VII
Congreso Venezolano de Sismologa e Ingeniera
Ssmica, Barquisimeto, 10 pp.
Heit, B., Koulakov, I., Asch, G., Yuan, X, Kind, R.
(2008). More constraints to determine the seismic
structure beneath the Central Andes at 21S using
teleseismic tomography analysis. J. South Am. Earth
Sci., 25, 2236.
Jcome, M.I., Kusznir, N., Audemard, F., Flint, S.
(2003). The formation of the Maturn Foreland Basin,
Eastern Venezuela: thrust sheet loading or subduction
dynamic topography. Tectonics, Vol. 22, 117.
Jcome, M.I., Rondn, K., Schmitz, M., Izarra, C.,
Viera, E. (2008). Integrated Seismic, Flexural and
Gravimetric modelling of the Coastal Cordillera
Thrust Belt and the Gurico Basin: North-Central
Region, Venezuela, Tectonophysics, 459, 27-37,
doi:10.1016/j.tecto.2008.03.008.
Krger, F., Scherbaum, F., Rosa, J.W.C., Kind, R.,
Zetsche, F., Hhne, J. (2002) Crustal and upper mantle
structure in the Amazon region (Brazil) determined
with broadband mobile stations. J. Geophys. Res.,
107 (B10), 2265, doi: 10.1029/2001JB000598.
Langston, C. (1979). Structure under Mount Rainier,
Washington, inferred from teleseismic body waves. J.
Geophys. Res., Vol. 84, 4749-4762.
Levander, A., Schmitz, M., Av Lallemant, H.G.,
Zelt, C.A., Sawyer, D.S., Magnani, M.B., Mann,
P., Christeson, G., Wright, J., Pavlis, D., Pindell,
J. (2006). Evolution of the Southern Caribbean Plate
Boundary. EOS, 87 (9), pp. 97, 100.
Lilliu, A. (1990). Geophysical interpretation of Maturn

Foreland, Northeastern Venezuela, MSc. Thesis,


Houston University, Houston, Texas, USA. 124 pp.
Macellari, C.E. (1995). Cenozoic Sedimentation and
Tectonics of the Southwestern Caribbean pull-apart
Basins of South America. Memoir, 62. American
Association of Petroleum Geologist, pp. 757780.
Magnani, M.B., Zelt, C.A., Levander, A., Schmitz, M.
(2009). Crustal structure of the South American
Caribbean plate boundary at 67W from controlled
source seismic data. J. Geophys. Res., 114, B02312,
doi:10.1029/2008JB005817.
Malfait, B. & Dinkelman, M. (1972). Circum-Caribbean
tectonic and igneous activity and the evolution of the
Caribbean plate. Geol. Soc. Am. Bull., 83(2), 251
272. doi:10.1130/0016-7606.
Meschede, M. (1998). The impossible Galapagos
connection: geometric constrains for a near-American
origin of the Caribbean Plate. Geol. Rundsch. 87,
200205.
Meschede, M. & Frisch, W. (1998). A plate-tectonic
model for the Mesozoic and Early Cenozoic history
of the Caribbean Plate. Tectonophysics 296, 269291.
Miller, M., Levander, A., Niu, F., Li, A. (2009).
Upper mantle structure beneath the CaribbeanSouth American plate boundary from surface
wave tomography, J. Geophys. Res., 114, B01312,
doi:10.1029/ 2007JB005507.
Molnar, P. & Sykes, L. (1969). Tectonics of the
Caribbean and Middle America regions from focal
mechanisms and seismicity. Geol. Soc. Am. Bull.,
Vol. 80, 6391684.
Muessig, K. (1978). The Central Falcn Igneous Suite,
Venezuela: Alkaline Basalt intrusions of Oligocene
Miocene Age. Geol. Mijnb. 52, 261266.
Muessig, K. (1984). Structure and Cenozoic tectonics of
the Falcn Basin, Venezuela and adjacent areas. Geol.
Soc. Amer. Mem. 162, 217230.
Navarro, E., Ostos, M., Yoris, F. (1988). Revisin y
definicin de unidades litoestratigrficas y sntesis
de un modelado tectnico para la evolucin de la
parte Norte-Central de Venezuela durante el Jursico
Medio-Palegeno. Acta Cient. Venez. 39, 427436.
Niu, F. & James, D.E. (2002). Fine structure of the
lowermost crust beneath the Kaapvaal craton and
its implications for crustal formation and evolution.
Earth Planet. Sci. Lett., 200, 121130.
Niu, F., Bravo, T., Pavlis, G., Vernon, F., Rendn, H.,
Bezada, M., Levander, A. (2007). Receiver function
study of the crustal structure of the southeastern
Caribbean plate boundary and Venezuela. J. Geophys.
Res., Vol. 112, B11308.
Officer, C.B., Ewing, J.I., Hennion, J.F., Harkrider,
D.G., Miller, D.E. (1959). Geophysical investigations
in the Eastern Caribbean; summary of 1955 and 1956

Captulo I - Caractersticas Geofsicas de la Corteza a lo Largo del Lmite de Placas

cruises. In: Ahrens, L.H., Press, F., Rankama, K.,


Runcorn, S.K. (Eds.), Physics and Chemistry of the
Earth, 3, pp. 17109.
Orihuela, N. & Ruiz, F. (1990). Modelaje gravimtrico
de un perfil comprendido entre los poblados de
Altagracia de Orituco, Edo. Gurico y Caraballeda,
Dtto. Federal, Venezuela. V Congreso Venezolano de
Geofsica, Caracas, pp. 466-473.
Orihuela, N., Garca, A., Tabare, T. (2011). Mapa de
gravedad y anomala gravimtrica de Venezuela
derivado de datos satelitales. Rev. Fac. Ing. UCV, 26
(1), 51-58.
Orihuela, N., Garca, A., Arnaz, M. (2012). Magnetic
anomalies in the Eastern Caribbean. Int. J. Earth. Sci.
(Geol. Rundsch.), DOI 10.1007/s00531-012-0828-6.
Ostos, M. (1990). Tectonic evolution of the south-central
Caribbean based on geochemical and structural data.
Ph.D. Thesis: Houston, Rice University, 411 pp.
Prez, O.J. & Aggarwal, Y.P. (1981). Presentday tectonics of the southeastern Caribbean and
northeastern Venezuela. J. Geophys. Res. 86, 10791
10804.
Picard, X. & Goddard, D. (1975). Geomorfologa y
sedimentacin en la costa entre Cabo Codera y Puerto
Cabello. Asoc. Ven. Geol. Min. y Petr., Bol. Inform.,
18 (1), 39-106.
Pindell, J.L. & Dewey, J.F. (1982). Permo-Triassic
reconstruction of western Pangea and the evolution
of the Gulf of Mexico/Caribbean region. Tectonics 1,
179211.
Pindell, J.L. & Barrett, S.F. (1990). Geological
evolution of the Caribbean region; a plate tectonic
perspective. In: Dengo, G., Case, J.E. (Eds.), The
Geology of North America: The Caribbean Region.
Boulder, vol. H. Geological Society of America, pp.
405432.
Pindell, J. & Kennan, L. (2007). Cenozoic kinematics
and dynamics of oblique collision between two
convergent plate margins: the Caribbean-South
America collision in eastern Venezuela, Trinidad and
Barbados. In: The Paleogene of the Gulf of Mexico
and Caribban basins: processes, events, and petroleum
systems. GCSSEPM Foundation, 96 pp.
Pindell, J., Kennan, L., Stanek, K., Maresch, W.,
Draper, G. (2006). Foundations of Gulf of Mexico
and Caribbean evolution: eight controversies
resolved. Geologica Acta, 4 (12), 303341.
Quinteros, C., Piero, L., Rendn, H., Schmitz, M.,
Niu, F. (2008). Estimacin del espesor de corteza en
el noroccidente de Venezuela, a partir del anlisis de
funciones receptoras. XIV Congreso Venezolano de
Geofsica, Caracas, 8 p.
Quinteros, C., Rojas, K., Amaz, R., Piero, L., Schmitz,
M., Rendn, H. (2009). Anlisis de funciones

receptoras para la estimacin del espesor de corteza en


Venezuela. IX Congreso Venezolano de Sismologa e
Ingeniera Ssmica, Caracas, 12 pp.
Rodrguez, J. & Sousa, J.C. (2003). Estudio geolgicoestructural y geofsico de la seccin cabo San RomnBarquisimeto, estados Falcn y Lara. Undergraduate
thesis, Universidad Central de Venezuela.
Rossi, T. (1985). Contribution a letude geologique
de la frontiere Sud-Est de la plaque Caraibes Etude
geologique de la Serrana, La Serrana del Interior
Oriental (Venezuela) sur le transect Cariaco-Maturin,
Syntheses Paleogeographique et Geodynamique,
Ph.D. Thesis, Universite de Bretagne Occidentale,
France. 340 pp.
Roure, F., Carnevali, J.O., Gou, Y., Subieta, T. (1994).
Geometry and kinematics of the North Monagas thrust
belt (Venezuela). Mar. Petrol. Geol. 11, 347362.
Russo, R.M. & Speed, R.C. (1994). Spectral analysis
of gravity anomalies and the architecture of tectonic
wedging, NE Venezuela and Trinidad. Tectonics, Vol.
13, 613622.
Snchez, J., Gtze, H.-J., Schmitz, M. (2010). A 3-D
lithospheric model of the Caribbean-South American
plate boundary. Internat. J. Earth Sci. (Geol.
Rundsch.), 100, 7, 1697-1712.
Schmitz, M, (2011). Evaluacin del espesor cortical en
Venezuela mediante mtodos sismolgicos activos y
pasivos .Trabajo de ascenso, Facultad de Ingenieria,
UCV, 71 pp.
Schmitz, M., Lessel, K., Giese, P., Wigger, P., Araneda,
M., Bribach, J., Graeber, F., Grunewald, S.,
Haberland, C., Lth, S., Rwer, P., Ryberg, T.,
Schulze, A. (1999). The crustal structure beneath the
Central Andean forearc and magmatic arc as derived
from seismic studies the PISCO 94 experiment in
northern Chile (21 - 23 S). J. South Am. Earth Sci.,
12 (3), 237-260.
Schmitz, M., Chalbaud, D., Castillo, J., Izarra, C.
(2002). The Crustal Structure of the Guayana Shield,
Venezuela, from seismic refraction and gravity data.
Tectonophysics, 345 (1-4), 103-118.
Schmitz, M., Martins, A., Izarra, C., Jcome, M.I.,
Snchez, J., Rocabado, V. (2005). The major features
of the crustal structure in north-eastern Venezuela
from deep wide-angle seismic observations and
gravity modelling. Tectonophysics, Vol. 399, 109124; doi:10.1016/j.tecto.2004.12.018.
Schmitz, M., Bezada, M., vila, J., Vieira, E.,
Ynez, M., Levander, A., Zelt, C.A., Magnani,
M.B., Jcome, M.I. and THE BOLVAR ACTIVE
SEISMIC WORKING GROUP (2008). Crustal
thickness variations in Venezuela from deep seismic
observations. Tectonophysics, 459, 14-26, doi:
10.1016/j.tecto.2007.11.072.

23

24

Michael Schmitz, Alan Levander, Fenglin Niu,


Maximiliano J. Bezada, Claudia Quinteros, Colin A. Zelt, Jess vila y el grupo de Trabajo de Ssmica Activa del Proyecto Bolvar

Schubert, C. (1984). Basin formation along the Bocono


MoronEl Pilar fault system, Venezuela. J. Geophys.
Res. 89, 57115718.
Sick, C., Yoon, M., Rauch, K., Buske, S., Lth, S.,
Araneda, M., Bataille, K., Chong, G., Giese, P.,
Krawczyk, C., Mechie, J., Meyer, H., Oncken, O.,
Reichert, C., Schmitz, M., Shapiro, S., Stiller,
M., Wigger, P. (2006). Seismic images of accretive
and erosive subduction processes from the Chilean
margin. In: O. Oncken, G. Chong, G. Franz, P. Giese,
H.J. Gtze, V. Ramos, M. Strecker, and P. Wigger
(Eds.), Frontiers in Earth Sciences, Vol. 1, The Andes
Active Subduction Orogeny, Springer, Berlin,
Heidelberg, pp. 147-169.
Sorensen, S.S., Sisson, V.B., Ave Lallemant, H.G.
(2005). Geochemical evidence for possible trench
provenance and fluid-rock histories, Cordillera
de la Costa eclogite belt, Venezuela. In: H.G. Av
Lallemant, V.B. Sisson (Eds.), Caribbean-South
America plate interactions, Venezuela. GSA Special
Paper, No. 394, pp. 173-192.
Sousa, J.C., Rodrguez, J., Giraldo, C., Rodrguez, I.,
Audemard, F., Alezones, R. (2005). An integrated
geological-geophysical profile across northwestern
Venezuela. 6th ISAG, Barcelona, Spain, Resmenes
ampliados, IRD, 4 pp.
Sykes, L.R., Mccann, W.R., Kafka, A.L. (1982). Motion
of Caribbean Plate during last 7 million years and
implications for earlier Cenozoic movements. J.
Geophys. Res., 87, 1065610676.
Unger, L.M., Sisson, V.B., Ave Lallemant, H.G. (2005).
Geochemical evidence for island-arc origin of the
Villa de Cura blueschist belt, Venezuela. In: H.G.
Av Lallemant, V.B. Sisson (Eds.), Caribbean-South
America plate interactions, Venezuela. GSA Special
Paper, No. 394, pp. 223-249.
Van Der Hilst, R. (1990). Tomography with P, PP, and
pP delay-time data and the three-dimensional mantle
structure below the Caribbean region. Faculteit
Aardwetenschappen der Rijksuniversiteit Utrecht,
Utrecht.

Van der Hilst, R., Mann, P. (1994). Tectonic implications


of tomographic images of subducted lithosphere
beneath northwestern South America. Geology 22,
451454.
Vandecar, J.C., Russo, R.M., James, D.E., Ambeh,
W.B., Franke, M. (2003). Aseismic continuation
of the Lesser Antilles slab beneath continental
South America. J. Geophys. Res., 108(B1), 2043,
doi:10.1029/ 2001JB000884.
Weber, J.C., Dixon, T.H., Demets, C., A., Ambeh, W.B.,
Jansma, P., Mattioli, G., Saleh, J., Sella, G., Bilham,
R., Prez, O. (2001). GPS estimate of relative motion
between the Caribbean and South American plates,
and geologic implications for Trinidad and Venezuela.
Geology 29, 7578.
Wilson, J. (1966). Are the structures of the Caribbean and
Scotia arcs analogous to ice rafting? Earth Planet. Sci.
Lett., 1, 335338, doi: 10.1016/0012-821x(66)900197.
Woelbern, I., Heit, B., Yuan, X., Asch, G., Kind, R.
(2009). Receiver function images from the Moho and
the slab below the Altiplano and Puna plateaus in the
Central Andes. Geophys. J. Int., 177 (1), 296-308.
Yuan, X., Sobolev, S.V., Kind, R., Oncken, O. And
Andes Seismology Group (2000). New constraints
on subduction and collision processes in the Central
Andes from comprehensive observations of P to S
converted seismic phases. Nature, 408, 958961.
Yuan, X., Sobolev, S.V., Kind, R. (2002). Moho
topography in the central Andes and its geodynamic
implications. Earth Plan. Sci. Lett., 199, 389402.
Zelt, C. & Smith, R.B. (1992). Seismic traveltime
inversion for 2-D crustal velocity structure. Geophys.
J. Int., 108, 16-34.
Zandt, G. & Ammon, C.J. (1995). Continental crust
composition constrained by measurement of crustal
Poissons ratio. Nature, 374, 152154.
Zhu, L. & Kanamori, H. (2000). Moho depth variation
in southern California from teleseismic receiver
functions. J. Geophys. Res., 105, 29692980.

Captulo II

25

ACTIVE BLOCK TECTONICS IN AND AROUND THE CARIBBEAN:


A REVIEW
Franck A. Audemard M.

Abstract. The knowledge of the Caribbean and its


plate boundaries has largely improved along the last
35 years despite getting progressively much more
complex. The borders of the Caribbean plate are
actual plate boundary zones PBZ-, in which many
tectonic blocks of different size, composition, origin
and geometry, are amalgamated. These blocks
somehow surround the Caribbean large igneous
province or oceanic plateau. The identification of
tectonic blocks along the southern Caribbean border
happened first, boosted by the fact that the plate
boundary was much less conspicuous compared
to the others. This was favored to a great extent
by the boundary cutting across continental areas.
Instead, the northern boundary of the Caribbean
plate became a natural laboratory for GPS Geodesy
because of the apparent simpler geometry of the
PBZ. Surprisingly, GPS networks have not solved
all targeted kinematic issues because most of the
networks are on small islands, sitting within the
deformation zone. Several of these islands in fact
exist because the contractional deformation made
them crop out, such as Jamaica and Hispaniola.
This problem is also common to the CaribbeanAtlantic PBZ. Stable reference points inside the
Caribbean Sea, such as San Andrs, Providencia
and Aves islands, are needed to resolve the motion
between North America and Caribbean plates.
From recent GPS results, the inner Caribbean plate
appears as a single, almost rigid body, at least under
the current 2-3 mm/a resolution of the approach.
The Hess Escarpment, which exhibits a nonnegligible seismic activity along its southwestern
submarine termination close to Costa Rica, may
be slipping in that order. Nevertheless, in the frame
of the Neogene Caribbean geodynamic evolution,
this submarine feature, which cuts the Caribbean
ocean floor into two large pieces, juxtaposes two
different Caribbean regions at naked eye. Finally,
strain partitioning at different scale is common to
all Caribbean plate boundaries. Block indentationextrusion and induced subductions are also
common.
Keywords: Block Tectonics, Escape, Indentation,
Induced
Subduction,
Buoyancy,
Oblique
Convergence, Hess, Caribbean

Resumen In-Extenso. La comprensin del Caribe


y sus bordes de placa en los ltimos 35 aos ha
definitivamente mejorado, a pesar de hacerse
progresivamente mucho ms complejo. De hecho,
las fronteras de la placa Caribe son zonas de
borde de placa ZBP-, en las cuales bloques
tectnicos de diverso tamao, composicin, origen
y geometra estn amalgamados (Figuras 1 y 2).
Estos bloques de cierta forma rodean la meseta
ocenica del Caribe. El reconocimiento de estos
bloques tectnicos ocurri primero a lo largo de
su frontera sur, como consecuencia que el lmite
de placa era menos sobresaliente. Ello se vio
ampliamente favorecido por estar la frontera en
reas continentales (Figuras 1 y 2). Por el contrario,
el borde septentrional de la placa Caribe se convirti
en un laboratorio natural de geodesia satelital
por GPS, como consecuencia de su aparente
simplicidad estructural. Sin embargo, las redes
GPS no resolvieron las incgnitas cinemticas
planteadas, dado que las redes estn instaladas
en pequeas islas, las cuales estn inmersas
dentro de la propia zona de deformacin (Figuras
1 y 2). Varias de estas islas de hecho existen por
los procesos contraccionales que las han elevado
fuera del mar, tal como Jamaica y Espaola. Igual
situacin se presenta a lo largo del borde de
placa entre las placas Caribe y Atlntica. Puntos
de referencia estable dentro del mar Caribe, tales
como las islas de San Andrs, Providencia y Aves,
son necesarios para resolver el movimiento relativo
entre las placas Norteamericana y Caribe.
El proceso de particin de las deformaciones de
distintas escalas es comn a las cuatro fronteras
de la placa Caribe (Figuras 3 y 4). En Amrica
Central, una franja costera, limitada por la trinchera
mesoamericana y el arco volcnico activo, escapa
hacia el NO, aprovechando el debilitamiento de
la corteza continental de la placa Caribe por el
volcanismo activo a nivel de la Amrica Central.
Igual situacin es reportada en la porcin norte de
las Antillas Menores, donde el antearco de este
sector de las Antillas se desplaza hacia el norte
relativo al resto del arco volcnico. En el lmite
de placas Norteamrica-Caribe, la franja ms
septentrional de la isla Espaola, limitada entre
las fallas Espaola Norte y Septentrional, al norte
y sur respectivamente, se desplaza hacia el oeste

26

Franck A. Audemard M.

con respecto al resto de Espaola. En el margen


sur de la placa Caribe, el bloque de Bonaire, asi
como el bloque que contiene las napas Caribe
sobrecorridas en el norte de Venezuela (aflorantes
en Cordillera de la Costa y serranas del Interior
central y oriental), hacen lo propio (Figuras 5 y 9),
aunque a tasas de movimiento ms lento que en los
otros bordes de placa.
Por otra parte, la indentacin y expulsin de
bloques, as como subducciones inducidas, son
procesos igualmente frecuentes en los bordes de
la placa Caribe. La indentacin por altos o relieves
submarinos de alta flotabilidad (ej., Carnegie y
Cocos), como motor de expulsiones de bloques
tectnicos, ha sido invocado para el bloque
Norandino y la franja costera de Amrica Central
que se extiende entre Costa Rica y Guatemala,
respectivamente (Figuras 2 a 5). En otros casos,
tales particiones de deformaciones han sido
atribuidas a la covergencia oblicua entre la placa
subductante y la sobrecorrida, como para el caso
del antearco de las Antillas Menores del norte y el
bloque septentrional de Espaola (Figura 8). No
obstante, el mejor ejemplo regional de indentacinexpulsin lo conforma la colisin y posterior
suturacin del bloque del Choc (originalmente
perteneciente al Arco de Panam) contra la
fachada occidental suramericana, que induce la
expulsin de una gran porcin de Suramrica, que
se extiende entre el golfo de Guayaquil o graben
de Jambel, en Ecuador, y las Antillas Holandesas
de Sotavento, al norte de Venezuela, incluyendo
toda Colombia montaosa (Figuras 1, 5 y 9). Este
proceso es relativamente joven, probablemente
iniciado en el Mioceno superior, con el cierre
parcial del canal del Caribe, pero hecho efectivo
ya en el Plioceno (5-3 Ma), cuando ocurre el
paroxismo tectnico ms reciente de la Cordillera
Oriental de Colombia y de los Andes de Mrida en
Venezuela, al igual que la transcurrencia dextral del
sistema que se extiende entre las fallas de Dolores
(Ecuador) y Bocon (Venezuela). Parte del retardo
en el acoplamiento efectivo del bloque de Choc
contra Suramrica, adems de la oblicuidad entre
los bloques en confrontacin, puede deberse a
la baja rigidez del indentor (Arco de Panam), el
cual sufre fuertes deformaciones internas (flexin
oroclina y fallamiento sinestral de orientacin
NO-SE; Figura 6). El choque efectivo del bloque
Choc conlleva a la expulsin de los bloques
Norandino, Maracaibo y Bonaire hacia el NNE. La
subduccin que conlleva a esta colisin est hoy
da parcialmente fosilizada entre el bloque Choc

y Suramrica, en asociacin al sistema de fallas de


Romeral. Esta colisin tiene expresin superficial
hasta la latitud de 4 N en Colombia, y se expresa
en superficie por el alineamiento ENE-OSO de las
fallas de Garrapatas, Ro Verde, Ibagu y el cambio
de estilo estructural del piedemonte llanero de la
Cordillera Oriental de Colombia de transcurrencia
dextral al sur al de transpresin dextral al norte,
aproximadamente a la latitud de Santa F de
Bogot. En profundidad, tal cambio estructural al
nivel de la losa de subduccin de afinidad caribea
(y fuertemente suturada contra el bloque Choc al
sur del golfo de Urab), coincide con el desgarre
de Caldas, ubicado algo ms al norte (5,6 N) que
su expresin superficial que llega hasta la latitud
de 4 N. Esta losa de subduccin ha sido muy bien
definida recientemente con base en anomalas de
velocidades de ondas P, la cual se hunde hacia
el ESE, bajo el bloque triangular de Maracaibo
y los Andes de Mrida, hasta profundidades de
700 km, pero no alcanza extenderse ms al norte
de la falla de Oca-Ancn (Figuras 7, 10 y 11), la
cual debe funcionar como su desgarre norte, para
separarla de la subduccin inducida de las Antillas
de Sotavento, a la cual se asocia el cinturn de
deformacin surcaribeo. A partir de vectores
GPS, el bloque Norandino senso stricto puede
ser subdividido en hasta 13 bloques tectnicos
menores (Figura 12), tal como a fines del Proyecto
GEORED del Servicio Geolgico Colombiano; pero
a fines de modelacin de deformacin rgida se ha
fragmentado de manera conservadora en slo 3
subbloques (Figura 13).
Por su parte, el interior de la placa Caribe parece
ser una nica unidad rgida, al menos hasta la
resolucin de los resultados GPS del orden de 2-3
mm/a. El escarpe de Hess, el cual presenta una
actividad ssmica no despreciable en su extremidad
suroeste (Figuras 14 y 15), parece moverse en ese
orden de velocidad. No obstante, en trminos de
la evolucin geodinmica negena del Caribe, este
rasgo mayor submarino que corta en dos el piso
ocenico Caribe en direccin NE-SO, yuxtapone
dos entidades caribeas muy dispares a simple
vista. Proponemos que este accidente puede
haber jugado un rol muy importante en la migracin
relativa hacia el este de la parte meridional caribea,
la contentiva del LIP o meseta ocenica, en el
Mioceno medio y el Mioceno superior. Actualmente,
se estara iniciando una reactivacin moderna,
igualmente con movimiento predominantemente
sinestral con componente subordinada normal;
pero esta vez ligada a la subduccin del alto de

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

Cocos de alta flotabilidad.


Palabras clave: Tectnica de Bloques, Escape,
Indentacin, Subduccin Inducida, Flotabilidad,
Convergencia Oblicua, Hess, Caribe.

INTRODUCTION
The original concept of a Caribbean plate CAsliding eastward with respect to the Americas in
a simplistic but very creative drawer-like manner
(Hess, 1962), has much evolved in the last 50
years. This model postulated early on that the very
small CA plate was sort of squeezed among larger
surrounding plates, and escaping like an orange pip
to the east relative to the others. Said in another
way, four rigid lithospheres in the Caribbean and
Middle-America regions could be defined from
the recorded earthquakes, active calc-alkaline
volcanoes and spreading ridges: The Caribbean
CA-, Cocos CO-, Northern America NA- and
Southern America SA- plates (Molnar & Sykes,
1969; Figure 1). This geometry required that the
northern and southern CA plate boundaries were
of the roughly east-west-trending transform type,
whereas the borders on the west and east were
type-B subduction zones (Figure 1). From carefully
observing a digital topographic-bathymetric map
of the Caribbean region, three of the four CA plate
boundaries are easily recognizable and traceable, in
a rough manner. However, this is not the case for
the southern border. The topographic expression
of the Bocon fault BF- in Venezuela is in no
way comparable to that of the Motagua-Polochic
Fault System MPFS- in Guatemala. Regarding
this issue, Kafka & Weidner (1979), among many
others, found difficult to analyze northwestern SA
using the concept of plate tectonics. Nevertheless,
Molnar & Sykes (1969) already sketched rather well
the major tectonic features to be taken into account
in that SW corner. In the late 70s and early 80s,
for instance, Schubert (1979; 1980; 1982a; 1982b;
1984) postulated that the CA-SA plate boundary
in Venezuela ran along the transform Bocon, San
Sebastin and El Pilar faults. Very early on, this led to
the fact that several tectonic blocks (or microplates)
had to be defined or identified along this southern
CA plate boundary to trace the possible connection
between the southern tips of the CO-CA subduction
on the west and the Atlantic-CA subduction on the

east. As a matter of fact, as Mann & Burke (1984)


indicated, the concept of Plate Boundary Zones
(PBZ) had to be applied to those two transform
boundaries because of the diffuse and spread
Neogene deformation over across-strike widths of
hundreds of kilometers, which was first highlighted
by Burke et al. (1980), Kafka & Weidner (1981)
and Sykes et al. (1982). This concept was taken
from that of Dewey & Sengr (1979), who adapted
it from the original application developed by
Luyendyk & MacDonald (1976) for deformation
zones at mid-ocean ridges. As predicted by Mann
& Burke (1984), only the gathering and integration
of more seismic and geologic studies, in which
satellite geodesy could not be considered because
not yet born, had led to a better knowledge of the
current Caribbean plate boundaries. This paper
intends to discuss how the concept of PBZ along
the Caribbean borders has been generalized not
only to the transform boundaries but has been
enriched by geo-scientific studies over the years.
Very particularly, the definition and identification
of tectonic blocks have been boosted in recent years
by unexpected GPS-derived vectors that have
led to the individualization of some of such blocks.
This has been very true to the four corners of the
CA plate where very complex tectonic settings are
imaged at the transfer zone or connection between
transform boundaries and typical subduction
zones; yet not well understood at every of these
four corners. To simplify the description of the four
CA plate borders, this paper will present each of
the borders separately in a clockwise way, starting
from the western border corresponding to the COCA subduction.
WESTERN CARIBBEAN PLATE BOUNDARY
Most of Central America is located in the
northwestern corner of the CA Plate. The western
CA plate boundary is essentially a type-B
subduction zone running along the Pacific coast
of Central America (Molnar & Sykes, 1969), very
well imaged in the Pacific ocean bathymetry by
the Middle-America trench, which extends farther
northwestward offshore along the Pacific coast of
Mexico into the Gulf of California (Figures 1 and
2). To the southeast, the Middle-America trench
and its related NE-dipping subduction extends

27

28

Franck A. Audemard M.

to a transform fault, the N-S trending Panam


Fracture zone, which puts in contact the CO and
Nazca NZ- oceanic plates to the west and east of
the transform respectively. The CO Plate subducts
under the CA plate at a convergence rate of 70-80
mm/a, and at an azimuth of about 20-22 (DeMets
et al. 1990; DeMets, 2001). Recently, this slip rate
has been bracketed at 78 1 mm/a, following the
construction of a GPS station at Cocos Island, being
the only current permanent station sitting on the
CO plate (Protti et al. 2012). The subducted slab
plunges at a fairly steep and average angle of about
45 (Bevis & Isacks, 1984; Burbach et al. 1984),
although, for instance, Protti et al. (1994), as well
as DeShon et al. (2006), discuss that the subducting
slab may locally exhibit considerable changes in
shape (dip and orientation) along southern Costa
Rica, in association to the Cocos ridge (Figure 2),
to which they bring a wealth of seismologic data as
supporting evidence.

This convergent plate boundary drives different


tectonic behaviors in the overriding (CA) plate
north and south of the Cocos ridge. Its effects can
be even followed to the northeast as far as into the
Caribbean plate interior, affecting the Caribbean
oceanic floor, as shall be discussed later in this paper.
The major change in the overriding plate, as could
be expected, does not coincide with the change in
plate at the Panam Fracture zone. Therefore, the
subduction of the Cocos Ridge along with the CO
plate is a major geodynamic process that defines
the current configuration not only under the Nicoya
Peninsula but also of this plate boundary, as well
as the tectonic processes occurring within the
overriding CA plate at larger scale.

Figure 1. Major tectonic features of the Caribbean region (modified from Stephan, 1982). Abbreviations (same as in text): BB
Bonaire block; BF Bocon fault; CAVA Central America volcanic arc; CB Choc block; CCRDB Central Costa Rica deformed belt;
EPGFZ Enriquillo-Plantain Garden fault zone; LAS Leeward Antilles subduction; MP Mona Passage; MPFS Motagua-Polochic fault
system; NAB North Andes Block; NHDB North Hispaniola deformed belt; NLAF Northern Lesser Antilles forearc; NPDB North
Panam deformed belt; RFS Romeral Fault System; SCDB Southern Caribbean deformed belt; SMBF Santa Marta-Bucaramanga
fault; TMB Triangular Maracaibo block.
Figura 1. Rasgos tectnicos mayores de la region caribea (modificado de Stephan, 1982). Abreviaturas (iguales al del texto):
BB Bloque de Bonaire; BF Falla de Bocon; CAVA Arco volcnico de Amrica Central; CB Bloque Choc; CCRDB Cinturn de
deformacin de Costa Rica Central; EPGFZ Zona de falla de Enriquillo-Plantain Garden; LAS Subduccin de las Antillas de
sotavento; MP Paso de la Mona; MPFS Sistema de falllas de Motagua-Polochic; NAB Bloque Norandino; NHDB Cinturn de
deformacin de Espaola Norte; NLAF Antearco de las Antillas Menores del Norte; NPDB Cinturn de deformacin Nor-Panameo;
RFS Sistema de Fallas de Romeral; SCDB Cinturn de deformacin Sur-Caribeo; SMBF Falla de Santa Marta-Bucaramanga; TMB
Bloque Triangular de Maracaibo.

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

Figure 2. Major tectonic features of the Caribbean region on relief map (from Lpez, 2010). Submarine features are well
expressed.
Figura 2. Rasgos tectnicos mayores de la region caribea en mapa de relieve (segn Lpez, 2010). Los rasgos submarinos se
expresan muy bien.

Northwestern Caribbean corner north of Cocos


Ridge
The northwestern Caribbean corner is wedgeshaped between the roughly NW-SE trending COCA subduction trench and the ENE-WSW-trending
transform fault, which cuts across Central America
through Guatemala (Figure 1). Here, the NA Plate
juxtaposes against the northern edge of the CA
Plate along a left-lateral transform fault system,
known offshore as the Swan Fault Zone that bounds
on the south the Cayman spreading center (Case &
Holcombe, 1980), and extends westward onshore
Central America as the Motagua-Polochic Fault
System MPFS- (Molnar & Sykes, 1969; Malfait
& Dinkelman, 1972; Figure 1). Relative NA-CA
plate motion is about 2 cm/a (Sykes et al. 1982;
Dixon et al. 1998; DeMets, 2001).

plate (Figure 3). This supports an average dip


of the subduction slab of about 45 along this
subducting slab stretch. The volcanic arc extends
from the MPFS to central Costa Rica, onshore of
where the Middle America trench looses its surface
expression, down south to the Central Costa Rica
deformed belt (CCRDB) as defined by Marshall et
al. (2000). Volcanoes are closely spaced, 12-30 km
apart, with elevations ranging from 100 m to more
than 4000 m (Carr, 1984). In general, the volcanic
front is 10-15 km wide (Carr & Stoiber, 1990).

Corner extension at acute fault junction


The volcanic arc associated to the CO plate
subduction comprises 75 basaltic to dacitic
volcanoes with documented Holocene activity, 31
of which have been active in historic times (Simkin
et al. 1981; Carr & Stoiber, 1990). They lie along a
line closely paralleling the Middle America trench,
some 150 km inside the overriding Caribbean

Figure 3. Strain partitioning along the Middle America trench.


Cartoon showing trench parallel sliver escaping towards the
NW, between CAVA and CO-CA subduction (from Alvarado et
al. 2011).
Figura 3. Particin de deformaciones a lo largo de la trinchera
mesoamericana. Bloque esquemtico mostrando el escape hacia
el noroeste del bloque limitado por el CAVA y la subduccin de
Cocos bajo el Caribe (segn Alvarado et al. 2011).

29

30

Franck A. Audemard M.

Just south of the MPFS and east of the volcanic


arc, in the mass wedge defined by the two former
features, lies a system of several N-S oriented
grabens (Figure 4; Dengo, 1968; Dengo &
Bohnenberg, 1969; Weyl, 1980; Mann et al. 1990;
Gordon & Muehlberger, 1994; Guzmn-Speziale
et al. 2005; Lyon-Caen et al. 2006). These grabens
are seismically active. In fact, Guzmn-Speziale
et al. (2005) have derived an ongoing east-west
trending extension from focal mechanism solutions
resulting in N-S trending normal faulting, which in
turn bounds the abovementioned grabens.

Figure 4. Corner effect between the MPFS and CAVA (from


Lyon-Caen et al. 2006), where east-directed escape of the
wedge-shaped corner induces E-W extension.
Figura 4. Efecto esquina entre el MPFS y CAVA (segn LyonCaen et al. 2006), donde se aprecia que el escape hacia
el este de la esquina definida por el MPFS y CAVA induce
extensin con fallas normales submeridianas.

Guzmn-Speziale (2001) had previously calculated


a maximum rate of opening (stretching) of 8
mm/a. This extension accounts for the CA-NA slip
reduction in Guatemala from 20 to 12 mm/a near
longitude 269.5 E, calculated by Lyon-Caen et al.
(2006). On the other hand, Guzmn-Speziale et al.
(2005), also using published CMT solutions, show
that a right-lateral strike-slip system parallels the
CO plate subduction and runs along the volcanic
arc axis extending from Mxico to the CCRDB.
This behavior may be explained by the fact that
silica and feldspar-rich rocks of continents and
island arcs deform more easily at low temperatures
than oceanic basalts do, as proposed by McKenzie
(1972). In this case, being rocks hot at volcanic
centers, regardless of their mineralogical
composition, they definitely constitute a preferential
weakness plane in the overriding plate.

Parallel-to-trench escaping sliver


Several authors (Fitch, 1972; Harlow & White,
1985; Guzmn-Speziale, 1995a; DeMets, 2001; La
Femina et al. 2002; Lyon-Caen et al. 2006; Alvarado
et al. 2011) have proposed that earthquakes along
the Central America Volcanic Arc CAVA- are
due to oblique subduction of the CO Plate. This
implies that slip partitioning takes place, in which
the normal-to-subduction component of CO-CA
convergence is taken by pure subduction along the
Cocos slab itself, while parallel-to-trench rightlateral component is taken by a forearc sliver,
named Central American coastal microplate by
Lyon-Caen et al. (2006). This upper plate block is
bounded by the CAVA and Middle America trench
on the north and south, respectively (La Femina et
al. 2002; Lyon-Caen et al. 2006; Alvarado et al.
2011; Figures 3 and 4). This microplate is moving
right-laterally with respect to stable CA at 10 mm/a
(Lyon-Caen et al. 2006). However, still in 2002,
Guzmn-Speziale & Gmez (2002) pointed out
that this model presented several problems, such as
very small along-arc components of relative plate
motion, seismogenic fault planes perpendicular to
the volcanic arc for some of the earthquakes, and
buttressing of the supposedly detached forearc at
its northwestern end. This issue has essentially
been resolved by satellite geodesy (Lundgren
et al. 1999; La Femina et al. 2002; Lyon-Caen
et al. 2006; Alvarado et al. 2011). The driving
mechanism of such parallel-to-trench motion can
be imputed to stress transfer applied by the buoyant
Cocos Ridge during its subduction along with the
CO plate under the CA plate, which acts as a sort of
a rigid indenter into the overriding plate (Jacob et
al. 1991; Montero, 1994a y b; Kolarsky et al. 1995;
Silver et al. 1995; Suarez et al. 1995; Tajima &
Kikuchi, 1995; Trenkamp et al. 2002; La Femina et
al. 2009). Shallow subduction of the Cocos Ridge
beneath the CA plate results in six major tectonic
effects (Kolarsky et al. 1995). These effects are: a) a
volcanic gap in the Costa Rican volcanic arc chain,
b) a shallowing of the dip of the subducted CO plate
beneath Costa Rica, c) forearc indentation of the
Pacific margin of Costa Rica (between Nicoya and
Osa peninsulas), d) structural inversion of forearc
(Terraba) and backarc (Limon) basins, e) arching
of on- and offshore acoustic basement in a direction
parallel to plate convergence between Costa Rica

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

and the CO plate, and f) a radial stress pattern


around the underthrust area of the Cocos Ridge as
inferred from earthquake and geologic indicators.
This last effect has been recently confirmed from
GPS-derived slip vectors (La Femina et al. 2002;
Lyon-Caen et al. 2006; La Femina et al. 2009;
Alvarado et al. 2011).
NA-CA-CO triple junction
NA-CA-CO triple junction has been classically
been defined as the intersection between the MPFS
and the Middle America trench in the Gulf of
Tehuantepec, offshore southeastern Mexico (White
& Harlow, 1993). Conversely, most workers
(Muehlberger & Ritchie, 1975; Plafker, 1976;
Burkart, 1978; Snchez-Barreda, 1981; Burkart,
1983; Guzmn-Speziale et al. 1989; GuzmnSpeziale, 2001) agree that this triple junction is not
a trench-trench-fault triple point in the classical
view of McKenzie & Morgan (1969). Instead, it is
a very broad zone of deformation. In that particular
respect, north of the MPFS in the NA plate, the
deformation zone includes the strike-slip faults
of Southeastern Mxico tectonic province (after
Guzmn-Speziale et al. 1989; Guzmn-Speziale,
2010) as well as the Reverse-Faults tectonic
province (Guzmn-Speziale & Meneses-Rocha,
2000; Guzmn-Speziale, 2010). This implies that
the left-lateral slip of the MPFS on the northern
side of its western fault tip is partly taken by both
shortening and lateral escape inside the NA plate
(Figure 1). To the south of the MPFS in the CA
plate, Lyon-Caen et al. (2006) show that the triple
junction is also complex and the deformation is
distributed over a wedge-shaped, 400 km-wide
area (Figure 4). This kinematic model is entirely
consistent with that proposed by Plafker (1976).
Besides, the extension expressed by N-S trending
grabens at the acute corner between the MPFS
and MAVA can be imputed to the escape effect
imposed by the strike-slip motion of both bounding
faults: left-lateral along the MPFS and right-lateral
on MAVA (Figure 4). In that respect, DeMets et al.
(2010) indicate that distributed east-west extension
of at least 5 1 mm/a occurs in areas of western
Central America. The result is consistent with
geologic and seismic observations of east-west
extension across central and western Honduras
and Guatemala (Manton, 1987; Guzman-Speziale,

2001; Lyon-Caen et al. 2006).


Central America south of Cocos Ridge
Panam microplate
The tectonic regime of the southern Caribbean
region in southern Central America is dominated by
the interaction of four major plates (CA, SA, CO,
and NZ; Figure 1). An added level of complexity is
introduced by the Isthmus of Panam, which acts
as a separate Panam or Choc-Panam microplate
(Adamek et al. 1988; Kellogg & Vega, 1995;
Lundgren et al. 1999; Taboada et al. 2000; Trenkamp
et al. 2002). The Panam block is considered a
broad zone of deformation (Pennington, 1981)
with diffuse boundaries. A suture zone between the
Choc and the North Andes Block NAB- defines
the eastern boundary. The Uramita fault zone in
the north and the Istmina fault zone in the south
(Taboada et al. 2000; Suter et al. 2008) mark an
island arc-continental collision of Neogene age
(Figure 5; Keigwin, 1978; Lonsdale & Klitgord,
1978; Keigwin, 1982; Keller et al. 1989). Offshore
southern Panam, the boundary between the
Panam microplate and the NZ plate is delimited
by the Southern Panam fault zone or Southern
Panam deformed belt, a diffuse left-lateral fault
zone that accommodates the eastward motion of
the NZ plate, and north-south convergence (Jordan,
1975; Hey, 1977; Adamek et al. 1988; Silver et al.
1990; Kolarsky & Mann, 1995; Mann & Kolarsky,
1995; Westbrook et al. 1995). The western
boundary is a broad zone of distributed left-lateral
shear in central Costa Rica (Montero-Pohly &
Dewey, 1982; Fan et al. 1991; Protti & Schwartz,
1994; Montero-Pohly, 1994a y b; Montero et al.
1998; Lundgren et al. 1999; Marshall et al. 2000;
Montero-Pohly, 2001), between the CA and the
Panam microplate, named the Central Costa Rica
Deformed Belt CCRDB. This belt seems to act as
a crustal (lithospheric?) left-lateral lateral ramp,
accommodating shortening and left-lateral slip
simultaneously but not necessarily on the same
tectonic features.
North Panam deformed belt NPDBThe north Panam deformed belt NPDBdefines the northern boundary of the Panam

31

32

Franck A. Audemard M.

microplate (Silver et al. 1990; (Figures 1 and 5).


Convergence at the NPDB is a consequence of: a)
slow southwestward convergence of the CA plate
with the Panam microplate (Kellogg & Vega,
1995; Trenkamp et al. 2002); b) ductile buckling
deformation (oroclinal bending) generated by
eastward motion of Panam against the North
Andes microplate NAB- (Wadge & Burke, 1983;
Silver et al. 1990; Montes et al, 2012); c) escaping
deformation along NW-SE trending left-lateral

strike-slip faults from the collision with the SA


plate (Wadge & Burke, 1983; Mann & Burke, 1984;
Mann & Corrigan, 1990), and d) the NE-directed
stresses (backarc thrusting) transferred from the
subduction of the buoyant, aseismic Cocos Ridge
beneath southern Costa Rica, between Nicoya and
Osa peninsulas (Jacob et al. 1991; Kolarsky et al.
1995; Silver et al. 1995; Surez et al. 1995; Tajima
& Kikuchi, 1995; Trenkamp et al. 2002).

Figure 5. Structural map of the northwestern corner of South America (from Taboada et al. 2000), showing the collision and suturing
of the Choc block (CB, corresponding to the southern part of the original Panam Arc) against SA. Major faults in Colombia and
Venezuela are shown.
Figura 5. Mapa estructural de la esquina noroeste de Suramrica (segn Taboada et al. 2000), donde se muestra la colisin
y suturacin del bloque del Choc (CB, correspondiente al sector sur del Arco de Panam original) contra Amrica del Sur. Se
identifican las fallas principales en territorio colombiano y venezolano.

The convergent nature of the NPDB has long


been recognized (Bowin, 1976; Pennington, 1981;
Wolters, 1986; Adamek et al. 1988; Silver et al.
1990; 1995). GPS measured convergence rates
between the Panam microplate and the CA plate
are to the southwest at 7 mm/a (Trenkamp et al.
2002). Substantial evidence for active convergence
along this boundary includes the 1991 (Ms 7.5)
Valle de la Estrella earthquake in eastern Costa Rica

(Jacob et al. 1991; Soulas, 1991; Plafker & Ward,


1992; Goes et al. 1993; Protti & Schwartz, 1994;
Surez et al. 1995), active folding in the NPDB
(Silver et al. 1990; 1995), a possible subduction
zone boundary marking the shallow subduction
of the CA plate beneath the Panam microplate
(Bowin, 1976; Wolters, 1986), and intermediatedepth earthquakes delineating a slab (Adamek et
al. 1988; Camacho et al. 2010).

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

True subduction or pseudo-subduction?


The controversy centers on the existence (or lack)
of a Wadati-Benioff zone depicting the subducted
slab of the CA plate beneath the Panam microplate.
Although Adamek et al. (1988) show a cross section
of relocated hypocenters that images a subducting
slab, they reject the NPDB as a subduction zone
due to the lack of volcanism and absence of events
below 70 km. Given the slow convergence rate
and the recent initiation of subduction (probably
< 10 Ma old) along the NPDB, in addition to the
fact that the CA plate is more rigid than a typical
oceanic crust because of being a thickened crust
(Reflector B; Large Igneous Province LIP- or
oceanic plateau; Edgar et al. 1971; Donnelly, 1973;
Case, 1975; Houtz & Ludwig, 1977; Biju-Duval
et al. 1978; Burke et al. 1978; Sinton et al. 1998;
Hauff et al. 2000), the slab has not yet sunk enough
to produce dehydration and subsequent volcanism.
Furthermore, it may never produce any volcanism
since it will keep as a shallow subduction while the
subducting CA slab corresponds to a thickened
more rigid and buoyant- oceanic crust. The same
situation and ongoing controversy is posed for the
Southern Caribbean deformed belt SCDB- along
northwestern Venezuela.
Several previous studies (Adamek et al. 1988; Silver
et al. 1990) concluded that the oblique convergence
between the CA plate and the Panam block had
not formed a Wadati-Benioff zone but merely
demonstrated an amagmatic overthrusting of the
Panam microplate onto the CA plate. Although
the NPDB in central Panam is an area of moderate
seismicity, previous studies using teleseismic data
from earthquakes in the NPDB were limited by the
dearth of events of magnitude >5.0. Camacho et al.
(2010), using data from local and regional networks,
present credible evidence of a Wadati-Benioff zone
rather than simple underthrusting for the Caribbean
margin of Panam. Earthquakes locate below the
Moho depth of 28 km, within the mantle, and to
depths of 80 km. Although Silver et al. (1990)
point out that the subduction of the CA plate under
Panam is a nonself-sustaining subduction zone in
the sense described by McKenzie (1977), there is
evidence of a subducted slab to a depth of at least
80 km.

To exacerbate this discussion, Stephan (1982) and


Stephan et al. (1986) proposed to call these festoons
of the NPDB and SCDB, as well as the one of
Hispaniola-Puerto Rico (corresponding to the Los
Muertos trench) pseudo-subductions, arguing that
they exhibit little intermediate seismicity to none,
and absence of volcanism. These authors proposed
that the festoons were just produced by plate-scale
bending, across the CA plate in the E-W direction in
their view, and subsequent overriding over the CA
plate. It is worth mentioning that this proposal is
far ahead in time of inception of the flat subduction
concept. Rather, we consider them as incipient
induced subductions (Audemard, 1993; 1998;
2009), thus justifying their youth and shallow depth
into the mantle, as supported by the absence of
generalized intermediate seismicity and the lack of
volcanism. This second aspect can be well explained
by the flat slab geometry by itself. So could shallow
seismicity if the subduction is very young. In our
view, these two alleged subductions are induced
from other major geodynamic processes, rather
than driven by mantle convection. In the case of
the NPDB, the Panam microplate is overriding
the CA plate because of the northward diachronic
and progressive collisional and later suturing
processes of the southern part of the original
Panam block (known as the Choc-Baud block,
in Colombia) against the NW corner of SA and
subsequent progressive internal (oroclinal-type?)
bending of the non-rigid Panam microplate. The
bending, as alleged by Stephan et al. (1996), in our
view, should account for a significant part of the
shortening taken in by the NPDB, which has not
stopped yet while the NPDB and SCDB festoons
do not collide. In the case of the SCDB, the festoon
is produced by the NNE-directed escape of the
North Andean Block (NAB; in the sense of Ego
et al. 1995), which adds to the N-S shortening
affecting the CA plate, due to the weak (~4 mm/a)
SSE-directed convergence between the Americas
since the Oligocene (Ladd, 1976; Sykes et al. 1982;
Burke et al. 1984). Audemard (2009) proposes that
the Leeward Antilles Subduction LAS- is rather
young and is different from the subduction that
led to collision of the Panam Arc against western
South America along the Choc Block and San
Jacinto terranes.

33

34

Franck A. Audemard M.

Panam microplate rigidity


As to the internal deformation of the Panam
microplate, Montes et al. (2012)s reconstruction
requires that the Panam microplate used to be
broken down on smaller blocks that rotated around
vertical axes. They also conclude that the current
S shape of the Isthmus may have been achieved
by oroclinal bending where discrete faults separated
relatively rigid blocks and helped their rotation.
Left-lateral offset of the Campanian to Eocene
belt between 38 and 28 Ma, opening of the Canal
Basin at ~25 Ma (Farris et al. 2011), and initiation
of the NPDB (Silver et al. 1990) at 15 Ma (Montes
et al. 2012), as the Panam-Choc Block was first
thrust to the NW onto the CA plate (Kellogg &
Vega, 1995; Camacho et al. 2010) are all results of
oroclinal bending of the arc. From their restoration,
Montes et al. (2012) agree with Farris et al. (2011),
that the gap between the former Panam-Choc
block and continental SA narrows at about 25 Ma
and effectively disappears at about 15 Ma. This
would coincide with the onset of shortening along
the NPDB estimated by Montes et al. (2012). This
would also place the age of effective collision of the
Baud-Choc block against the Pacific continental
edge of the SA plate not earlier than late Miocene.
The final suturing of both blocks is actually still
underway at the Uraba gulf (Caribbean Colombian
coast), implying that effective stress transfer of the
collision from the Baud-Choc block to the future
NAB may well have happened in the Pliocene, as
suggested by Audemard (1993; 1998; 2009) and
supported by a wealth of geologic data collected
in Venezuela (Audemard, 1993; Audemard &
Audemard, 2002; Audemard, 2003; 2009). Such a
delay of several millions years between first stages
of collision and effective suturing, responsible for
the NNE-directed escape initiation of the NAB, can
be accounted for by: a) acute angle of attack between
the former Panam Arc and the Pacific coast of
South America, with a first impact zone tending to
the southern tip of the arc; b) Oroclinal bending of
the Panam Arc (mainly preserved in the current
Panam microplate) attests to the low rigidity of
the arc, which should surely retard the effective
stress transfer of the Panam Arc collision against
continental South America while the indenter is
internally deforming itself; and c) N-S diachronism
of the effective suturing of the Choc-Baud block

against northwestern South America, and freeing


itself from the Panam Arc. Certain refinements in
the chronology of the escape of the North Andes
Block at regional scale can be proposed. The onset
has occurred at some time between the beginning
of Choc accretion (Middle Miocene; Duque-Caro,
1990), the effects of this in the Eastern Cordillera
of Colombia (10.5 Ma, Cooper et al. 1995), similar
to the Mrida Andes (Audemard & Audemard,
2002), and development of the Middle America
land bridge, showing coupling of Choc with South
America. First Caribbean-Pacific planktonic faunal
divergence occurred by 6.2 Ma (Keller et al. 1989),
mammal exchange occurred by 3.3 Ma (Gingerich,
1985), planktonic faunas show Caribbean-Pacific
separation by 3.1 Ma (Keigwin, 1978) and final
gateway closure occurred by 1.8 Ma (Keller et al.
1989).
From the above discussion, it remains clear that the
Panam microplate is not a rigid block. Rather, it
amalgamates a set of NW-SE elongated slivers at
least in Panam (Figure 6); split apart by equally
trending left-lateral active faults (Silver et al. 1990;
Rockwell et al. 2010). Furthermore, the microplate
progressively bends through time, becoming a
more tightly bent orocline, with an apex pointing
roughly to north. Many of the major faults of
Panam have long been recognized in the geology
and geomorphology (Jones, 1950; Woodring,
1957; Stewart et al. 1980), although they were not
considered Holocene (Cowan et al. 1998; Cowan,
1999; Schweig et al. 1999; Petersen et al. 2005).
Among these, the Pedro Miguel, Limn, and
related faults comprise a zone that extends from the
southern flank of the Sierra Maestra in north-central
Panam southward for at least 40 km, crossing the
Panam Canal between the Miraflores and Pedro
Miguel Locks, and extending southward offshore
into the Gulf of Panam (Figure 6; Rockwell et al.
2010).
Timing of North Andes Block escape
Further constraint on the North Andes Block
NAB- escape is provided by the dextral-slip onset
of the Dolores-Pallatanga-Algeciras-GuaicramoBocon fault system. This brittle system is
responsible for allowing the free NNE-directed
escape of the NAB (Case et al. 1971; Dewey, 1972;

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

(2010) and Bezada et al. (2010) show that no cold


discontinuity goes beyond 100 km deep north of
the Oca-Ancn Fault system, while a very large
anomaly is south of this fault (Figure 7), which
deepens in ESE direction instead, and to which the
seismic Bucaramanga nest seems related.

Figure 6. Structural Location map of central Panam, showing


the Pedro Miguel (PMF), Limn (LF), Rio Gatn (RGF), Miraflores
(MF), and Azuero-Sona (A-SF) faults (from Rockwell et al. 2010).
This significant brittle deformation in NW-SE trending slivers,
attests to the low rigidity of the Panam microplate.
Figura 6. Mapa de ubicacin estructural de Panam central,
mostrando las fallas Pedro Miguel (PMF), Limn (LF), Rio Gatn
(RGF), Miraflores (MF), y Azuero-Sona (A-SF) (segn Rockwell
et al. 2010). La importante deformacin frgil en bloques
elongados NO-SE, atestiguan la baja rigidez de la microplaca
de Panam.

Pennington, 1981; Stephan, 1982; Audemard,


1993, 1998; Freymueller et al. 1993; Ego et al.
1996). Opening of the pull-apart Jambel Graben
in the Gulf of Guayaquil, Ecuador, is related to this
northward escape of northwestern South America
(Audemard, 1993, 1998; Figure 1). This author
attributes basin formation to localized transtension
in a horsetail splay structure at the SW termination
of that very long fault system, nearing the Nazca
trench at the NZ-SA-NAB triple junction. Basin
fill began in the Late Miocene (Bentez, 1986).
If SCDB is a result of the NE escape of the
North Andes Block as proposed by Audemard
(1993; 1998; 2009), this major feature may shed
additional light on the escape onset. Indications
that SCDB is relatively young come from: (1)
seismic profiles and bathymetry (Silver et al. 1975;
Talwani et al. 1977; Mascle et al. 1979; Kellogg
& Bonini, 1982; Ruiz et al. 2000); (2) PliocenePleistocene deformation of the accretionary prism
west of Santa Marta (Ruiz et al. 2000); (3) lack of
significant sedimentation in Los Roques canyon
in the last few tens of millions years; (4) paucity
of intermediate thrust earthquakes, up to 200
km deep, under the Maracaibo Basin (Orihuela
& Cuevas, 1992; Malav & Surez, 1995); and
(5) Recent tomographic images from Bezada

Figure 7. P-wave velocity anomalies in the shallowest 130 km


of the model (from Bezada et al. 2010). The blue areas are
considered colder than the mantle and the red ones hotter.
North of the Oca-Ancn fault OAF, the colder areas do not
show beyond 130 km in depth. Then, there is no evidence of a
deep cold oceanic slab in association with the SCDB.
Figura 7. Anomalas de velocidad de la onda P en los
primeros 130 km de profundidad del modelo (de Bezada
et al. 2010). Las reas azules se consideran ms fras que el
manto circundante y las rojas ms calientes. No hay reas fras
al norte de la falla de Oca-Ancn OAF- a profundidades
superiores a los 130 km. Entonces, no hay evidencia de una losa
de subduccin fra profunda asociada al SCDB.

35

36

Franck A. Audemard M.

Forces driving NAB escape


Other driving mechanisms than the Panam-Choc
collision against the SA Pacific face have been
invoked or put forward to induce escape of the
NAB. Pennington (1981) and Gutscher et al. (1999)
proposed that the arrival of the aseismic Carnegie
Ridge at the Ecuador-Colombian trench initiated
the escape, very similar to Cocos Ridge under
Costa Rica (Figure 2), while Kellogg & Mohriak
(2001) proposed that the rapid oblique subduction
of the NZ plate and the Carnegie Ridge might be
driving together the northeastward escape. Elastic
modeling of observed horizontal displacements
in the Ecuador forearc is consistent with partial
locking in the subduction zone and partial transfer of
motion to the overriding SA plate (Trenkamp et al.
2002; White et al. 2003). Gutscher et al. (1999) and
Bourdon et al. (2003) noted an apparent shallowing
in the subduction of the NZ plate in NW Ecuador
based on the distribution of hypocenters obtained
from a local seismic network as well as the nearly
complete absence of intermediate depth seismicity
between 2.5 N and 1 S. They hypothesized that
this shallow subduction zone was the subducted
extension of the Carnegie Ridge. Pennington
(1981), Gutscher et al. (1999), Kellogg & Mohriak
(2001), and Trenkamp et al. (2002) proposed that
the subduction of the thick buoyant crust of the
Carnegie Ridge resulted in increased coupling with
the overriding South American plate. Scholz &
Small (1997) proposed that even the subduction of
a large seamount would increase the normal stress
across the subduction interface, thereby increasing
seismic coupling. The inferred continuation of the
Paleo-Carnegie Ridge beyond the trench (Gutscher
et al. 1999) has a NE orientation and is compatible
with the displacement direction of the North Andes.
GPS measurements of ENE rapid subduction (58
2 mm/a) of the NZ plate and Carnegie Ridge
under South America at the Ecuador trench are also
consistent with the northeastward displacement of
the North Andes (Trenkamp et al. 2002).
Egbue & Kellogg (2010) propose that stable SA
(Guyana shield) is acting mechanically as a rigid
buttress for the margin-normal component of NZ-SA
convergence, while the margin-parallel component
of the NZ-SA America convergence is driving the
North Andes northeastward toward the relatively

free CA-North Andes boundary. For extrusion


tectonics to occur, the block being impinged on
cannot be bilaterally confined (Tapponnier et al.
1982). The escape direction is always toward the
free boundary, which in the northern Andes is
towards the Caribbean, as previously indicated for
the NAB by Audemard (2003) after applying the
same concepts.
As a matter of fact, we are still inclined to support
the Baud-Choc block collision against NW SA
as the main driving cause for the NAB escape.
However, the buoyant Carnegie ridge subduction,
as well as the convergence obliquity between the
NZ plate and the SA freeboard, may substantially
add to it. From the time of initiation of the Carnegie
ridge collision, discussed just after this, it would
seem that this collision comes into the escape story
somehow late, at the Plio-Pleistocene boundary.
In the same way, NZ-SA convergence obliquity
cannot be the driving force of the NAB escape
because the latter would be a process as old as the
own NZ plate subduction.
Gutscher et al. (1999), based on examination of
the basement uplift signal along a trench-parallel
transect, and adakite volcanism along the Ecuador
arc, proposed that the ridge reached the trench and
has been colliding with South America for at least
2 Ma and most likely for the last 8 Ma. Rather,
Spikings et al. (2001) from plate convergence
rate calculations, suggest the Carnegie Ridge
collided with the Ecuador Trench at ~15 Ma.
Subsequently, Garrison & Davidson (2003) among
others, documented that adakites in the Andes can
be explained by the state of equilibrium that exist
between the mantle wedge derived arc magma
and the thickened garnet-bearing continental crust
and therefore do not require melting of subducted
oceanic slabs. Pedoja (2003) and Cantalamessa
& Di Celma (2004), based on analyses of marine
terrace uplift, independently postulated that ridge
subduction began at the Pliocene-Pleistocene
boundary (1.8 Ma). Long before, Lonsdale &
Klitgord (1978) proposed that the Carnegie
Ridge arrived at the trench about 1 Ma, based
on interpretation of magnetic anomalies and
bathymetry of the CO and NZ plates.

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

NA-CA PLATE BOUNDARY


This PBZ is the one that best images how the
definition of faults, and later identification of
discrete blocks, have evolved through time,
recognizing step by step the very large complexity
of the PBZ. Due to the fact that this PBZ slaloms
through islands of diverse size and varied geology,
satellite geodesy was sought and put in practice
rather early, compared to the rest of the Caribbean
region, to bring insights in the active tectonics of
this NA-CA plate boundary. Nevertheless, the
different block entities are still being (re-)defined
with the help of GPS networks while the others in
larger continental areas or blocks are roughly better
known at present because major active bounding
tectonic features are better and/or more exposed.
As clearly expressed by Dixon et al. (1991), over
the past two decades, a fundamental objective of
neotectonic research in the Caribbean region has
been to determine the present motion of the CA
plate using Global Positioning System (GPS)
technology.
Prior to the use of GPS for measuring present plate
motions, CA-NA plate velocities were estimated
from conventional marine geophysical and
seismologic observations. The predictions of the
latter estimates varied widely, ranging from 11 6
mm/a of sinistral strike-slip motion (Jordan, 1975;
Stein et al. 1988; DeMets et al. 1990; 1994) to 37
10 mm/a (95% uncertainty) of oblique convergence
(Sykes et al. 1982) along much of the CA-NA plate
boundary. The wide range of predicted motions
resulted from disagreements about which, if any,
data constituted reliable measures of CA plate
motion, including whether earthquake slip vectors
from the Middle America and Lesser Antilles
trenches are systematically biased by strain
partitioning (Sykes et al. 1982; Stein et al. 1988;
DeMets, 1993; 2001; Deng & Sykes, 1995) and
whether magnetic anomalies from the Cayman
spreading center record the full CA-NA rate (Sykes
et al. 1982; Rosencrantz & Mann, 1991). The first
unambiguous geodetic determination of present
day CA plate motion was reported by Dixon et al.
(1998) from GPS measurements made at three sites
during the early to mid-1990s. Relative to sites
on the NA plate, all three stations moved 18-20
mm/a, ~80% faster than predicted by the NUVEL-

1A model (DeMets et al. 1994), widely used at


the time. Subsequent geodetic measurements at
additional sites in the eastern Caribbean confirmed
this result (MacMillan & Ma, 1999; DeMets et al.
2000; DeMets, 2001; Sella et al. 2002) and further
demonstrated that CA-SA plate motion significantly
exceeds that predicted by NUVEL-1A (Weber et al.
2001; Sella et al. 2002). DeMets et al. (2010) stress
that it is thus now well established that the CA
plate moves significantly faster than predicted by
NUVEL-1A. On top of this, these authors express,
and largely discuss, that all previous geodetic
models of Caribbean plate motion have had two
significant, though unavoidable, drawbacks related
to their underlying geodetic data.
Structure of the NA-CA PBZ
In terms of plate boundary structural complexity,
the understanding of the NA-CA plate boundary
evolved from a simple fault system (Molnar &
Sykes, 1969; Jordan, 1975; Ladd, 1976; Plafker,
1976; Pennington, 1981) to the complex PBZ as
recently defined, for instance, by Benford et al.
(2012 a and b; Figure 8), but was already rather well
depicted by the review of Mann & Burke (1984),
which was slightly improved in that of Mann et al.
(1990). This NA-CA plate boundary zone consists
of a 100-250-km-wide seismogenic zone of mainly
left-lateral strike-slip deformation extending over
2000 km along the northern edge of the Caribbean
Sea. The dominant structural element in the central
plate boundary zone is the Cayman trough (Figures
1 and 2), a submarine pull-apart basin formed by
at least 1100 km of oceanic spreading at the MidCayman spreading center, a 100-km-long jog
between left-lateral faults of the plate boundary
(CAYTROUGH, 1979). The spreading center
has been active since the Middle Eocene and is
currently spreading at a rate of about 15 mm/yr
(Rosencrantz et al. 1988). In the western Cayman
trough, the plate boundary appears to be a single,
active strike-slip fault exhibiting locally complex
restraining bends (Mann et al. 1991a). To the east
of the Cayman trough in Jamaica, Hispaniola, and
Puerto Rico, the plate boundary is especially wide
with a seismogenic zone up to 250 km wide at the
longitude of the island of Hispaniola. In addition,
the NA-CA PBZ is essentially left-laterally
transforming west of Central Hispaniola (west of

37

38

Franck A. Audemard M.

the Beata ridge longitude), while it functions as


an oblique subduction from eastern Hispaniola
eastward up to its connection to the Lesser Antilles
subduction, near the Virgin Islands (Dolan et al.

1998; Mann et al. 2002). Ten Brink et al. (2009)


define this PBZ along Hispaniola and Puerto Rico
as a bivergent crustal wedge.
A PBZ of two-parallel fault zones

Figure 8. Tectonic blocks in the CA-NA PBZ defined by Benford et al. (2012b) from GPS data.
Figura 8. Bloques tectnicos definidos en la zona de frontera de placas CA-NA por Benford et al. (2012b) a partir de datos GPS.

Discrete strike-slip fault zones, which are throughgoing or mappable for hundreds of kilometers, have
been recognized along this PBZ. They are linked
by either transpressional (push-up or restraining
bend) segments or transtensional (pull-apart basin)
segments. In this respect, the NA-CA transform
boundary somewhat resembles California (Allen,
1981). In the NA-CA PBZ, these faults are (Mann
and Burke, 1984; their Figure 4) from west to east:
(1) the Chixoy-Polochic Fault Zone of Guatemala
and southern Mexico (Burkart, 1978, 1983, Erdlac
& Anderson, 1982); (2) the Motagua Fault Zone of
Guatemala (Plafker, 1976); (3) the Swan Fault Zone
of the Cayman Trough (Case & Holcombe, 1980);
(4) the Oriente Fault Zone of the Cayman Trough
(Case & Holcombe, 1980); (5) the Duanvale Fault
Zone of Jamaica (Horsfield, 1974; Mann & Burke,
1980); (6) the Enriquillo-Plantain Garden fault
zone EPGFZ- of Jamaica and Hispaniola (Mann
et al. 1983); (7) the Septentrional fault zone of
the Dominican Republic (Bowin, 1975; Mann
et al. 1984); and (8) the Puerto Rico trench fault
zone (Perfit et al. 1980; Case & Holcombe, 1980).
These Neogene faults are arranged into two distinct
but rather parallel left-lateral strike-slip zones. A
northern strike-slip zone strikes westward from the
Puerto Rico trench and passes through: (1) Northern
Hispaniola (Septentrional fault zone); (2) along the
southern margin of Cuba (Oriente Fault Zone) and
(3) into a zone of sea-floor spreading (Mid-Cayman
spreading center) at the center of the Cayman
trough. On the other hand, the southern strike slip
zone extends westward from Central Hispaniola

(Enriquillo fault zone) and passes through: (a)


the southern peninsula of Haiti (Dcrochement
senestre sud Hatien of Calmus, 1983, and Bizon
et al. 1985; also known as Plantain Garden fault
zone, which extends westward into Jamaica across
the Jamaica passage); (b) Jamaica as a series
of reverse and thrust faults, which conforms a
transpressional bend. The main left-lateral strikeslip fault through this bend, as revealed by GPS
vectors, runs east-west across Jamaica axially or in
the southern half of the island (DeMets & WigginsGrandison, 2007) or offshore south of the island
(Benford et al. 2012b); (c) the southern boundary
of the Cayman trough (The Walton and Swan fault
systems, east and west of the spreading center,
respectively) and (d) the onshore prolongation of
the Swan Fault system across onshore Guatemala
(Motagua-Polochic fault system MPFS-). This
second southern left-lateral strike-slip fault system
connects to the east to the N-dipping Los Muertos
trench that runs south offshore of Hispaniola and
Puerto Rico islands. This in turn transfers slip to
the Anegada Passage fault system. In an implicit
manner, the first block ever defined in this NACA PBZ corresponds to the one bounded by
the two previously described largely left-lateral
transcurrent fault system. This block thus extends,
in a west-east direction, from the Cayman sea-floor
spreading center to the junction of the Anegada
Passage fault system with the Puerto Rico trench.

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

Microplates within the PBZ

Hispaniola-Puerto Rico and Gnave microplates

In spite of the active tectonics of the NA-CA PBZ


being well identified and characterized as early as
the 1980s, the fragmentation of the PBZ in discrete
smaller blocks has happened much later with the
inception of the GPS technology and progressive
installation of numerous networks of rather small
aerial coverage in the various islands by different
groups (Calais et al. 2002 in Hispaniola, Jansma &
Mattioli, 2005 in the Lesser Antilles, and DeMets
& Wiggins-Grandison, 2007 in Jamaica). Slip
vectors of local/regional coherence in tendency
have definitely helped define the boundaries of the
different tectonic blocks conforming this PBZ.

Consequently, the Hispaniola-Puerto Rico and


Gnave microplates, in that chronologic order, are
the two first tectonic entities to be defined within
the NA-CA PBZ. The two microplates together
conform the entire block extending from the
Cayman spreading center to the Virgin Islands,
previously described within the NA-CA PBZ. They
are contiguous on the east-west direction, being the
Gnave microplate the one sitting to the west. The
limit between the two microplates has been subject
of live debate. It may lie west of Hispaniola, within
central Hispaniola or may be diffuse (Manaker et
al. 2008; Calais et al. 2010; Benford et al. 2012a).
As a matter of fact, while Heubeck et al. (1991),
Mann et al. (1991b), Rosencrantz & Mann (1991)
and Mann et al. (1995) place the eastern boundary
of the Gnave microplate in the western flank of
the Central Range of Hispaniola, Pubellier et al.
(1991; 2000) place this transpressional boundary in
the trans-Haitian faults, within the Haitian Foldand-Thrust belt, and particularly indicate that the
slip along the Los Muertos trench is transferred
to the 120 trending Chane des MatheuxMontagnes du Trou dEau fault. The latter lies
in a more southwestern position with respect to
the Central range of Hispaniola. Pubellier et al.
(2000) claim that the limit proposed by previous
authors used to be the active boundary but this
microplate boundary has migrated in SW direction
and taken the present configuration since the
Pliocene instead. Shortening through the Neogene
in the eastern part of the Gnave microplate was
thus accommodated slowly by ramp anticilines
propagating and migrating southwestward.
Benford et al. (2012a) test several elastic block
models for various Gnave microplate geometries,
from a Gnave microplate extending between the
Cayman spreading center and the Mona passage,
which was rejected at a high confidence level, to
a likely diffuse boundary within or offshore from
western Hispaniola (western coast of Haiti), trying
to fit to the eastern boundary proposed by Pubellier
et al. (2000).

Several groups of workers have proposed that the


fault bounded region extending eastward from the
Cayman sea-floor spreading center to the transfer
zone of the Puerto Rico oblique subductionLesser Antilles frontal subduction, may reflect
the existence and active movement of intervening
microplate(s) within the NA-CA PBZ. For example,
Byrne et al. (1985), Mauffret & Jany (1990), and
Masson & Scanlon (1991) proposed the existence
of an active fault-bounded Hispaniola-Puerto Rico
microplate on the basis of active, submarine fault
boundaries present on both the north and south
sides of the Virgin Islands, Puerto Rico, and eastern
Hispaniola. Heubeck et al. (1991), Mann et al.
(1991b), Rosencrantz & Mann (1991) and Mann
et al. (1995) have all emphasized the presence of
an adjacent but discrete microplate to the west of
the Hispaniola-Puerto Rico microplate, stretching
westward from central Hispaniola to the Cayman
spreading center. Rosencrantz & Mann (1991)
have named this fault-bounded area, the Gnave
microplate after Gnave Island, one of the larger
landmasses entirely within the proposed microplate
boundaries in western Hispaniola. The 190,000km2, rectangular-shaped microplate is bounded to
the west by the Mid-Cayman spreading center, to the
north and south by the two continuous and discrete
northern and southern left-lateral strike-slip fault
zones earlier described, that extend from the MidCayman spreading center to the about longitude 71
W in central Hispaniola (around the meridian of
Beata ridge), and to the east by a complex thrust
zone in central Hispaniola (Heubeck & Mann,
1991b; Mann et al. 1991c).

The southern limit of the Gnave microplate runs


along the Walton and Plantain Garden-Enriquillo
Fault Zones, which are connected by a very large
transpressional bend, resulting in the uplift and

39

40

Franck A. Audemard M.

growth of the island of Jamaica (Mann et al. 1985;


2007). From a dense GPS network in Jamaica,
DeMets & Wiggins-Grandison (2007) determine
that the kinematic and seismic data thus indicate
that deformation on the island is dominated by
left-lateral shear along largely E-W-trending strikeslip faults. In addition, relative to the CA Plate,
the Jamaican GPS velocities calculated by these
authors exhibit a nearly monotonic increase in
site velocities from south to north along a transect
orthogonal to the islands major east-west faults.
However, velocities that they calculated along
a transect orthogonal to the islands numerous
NNW-striking faults, also increase monotonically
from the WSW to ENE, being supportive of an
interpretation of those faults as restraining bends.
In other words, Jamaica definitely sits in the
boundary between the Caribbean and the Gnave
microplate. On top of that, it is very likely that
the 8 1 mm/a estimated by DeMets & WigginsGrandison (2007), represents a minimum value for
the rate of Gnave-CA Plate motion. Besides, they
also estimate a firm upper limit of 13 1 mm/a
for Gnave-Caribbean Plate motion assuming
that additional elastic or permanent deformation
occurs north of Jamaica beyond the reach of the
existing GPS network. We add that part of that
unmeasured deformation may also lie south of
Jamaica, since most restraining bends tend to be
bivergent.
Break-up of
microplate

the

Hispaniola-Puerto

Rico

In turn, the Hispaniola-Puerto Rico microplate


appears otherwise to the rigid tectonic block
proposed by Byrne et al. (1985), Mauffret & Jany
(1990), and Masson & Scanlon (1991). The latter
authors propose for the tectonic setting of Puerto
Rico, major strike-slip movement on nearly eastwest lines in the vicinity of the Puerto Rico trench
coupled to a small counterclockwise rotation
of a Puerto Rico block within the broader plate
boundary zone. This simple model is attractive
to these authors because it predicts the tectonic
regime south of Puerto Rico (opening of the
Anegada passage and overthrusting of this block
onto the CA plate at the Los Muertos trench), and
provides an explanation for a possible component
of extension across the Puerto Rico Trench west

of 65.5W. This microplate has been broken


down into 3 smaller blocks, where Hispaniola and
Puerto Rico islands sit each in a different block
Hispaniola block and Puerto Rico block-, and are
separated by the rift system of the Mona Passage
(Figure 8). The extension of the Mona Passage,
responsible for the tsunamigenic 1918 Mayagez
earthquake, results from a differential ENE motion
relative to NA plate, where Hispaniola appears
lagging behind Puerto Rico (Calais et al. 2002).
Jansma et al. (2000) have shown that GPS data
from Puerto Rico and Hispaniola are consistent
with about 5 mm/a of east-west extension in the
Mona Passage, the marine strait between the two
islands. The same GPS study also revealed that
the Puerto Rico-northern Virgin Islands microplate
proposed by Masson & Scanlon (1991) on the basis
of marine geophysical data is moving at a rate no
faster than about 1 mm/a relative to the stable CA
plate. In that respect, Mann et al. (2002) state that
GPS velocities in the Puerto Rico, Virgin Islands
and Lesser Antilles sites show that these areas are
moving as part of the stable CA plate, at least at
the 2-3 mm/a level. Mann et al. (2002) extend this
conclusion to the entire Lesser Antilles volcanic arc
and forearc area, at least in the area of Barbados
Island. In the same way, the behavior of the Puerto
Rico-Virgin Islands and Lesser Antilles as a single
block indicates very slow (<1.5 mm/a) or no
motion on the eastern part of the Los Muertos fault
south of Puerto Rico and the Anegada Passage fault
between Puerto Rico and St. Croix. These results
do not support tectonic models involving present
day rotation of the Puerto Rico-Virgin Islands
block about a nearby vertical axis and/or eastward
tectonic escape of a Puerto Rico-Virgin Islands
microplate (Jansma et al. 2000).
As to the driving mechanism of the opening of
the Mona Passage, Mann et al. (2002) invoke that
northeastern Hispaniola during oblique subduction
has collided against the Bahamas platform while
Puerto Rico moved more freely, allowing even
some 25 of counterclockwise rotation of Puerto
Rico. Then, complex rifting in the Mona Passage
(Grindlay et al. 1997; van Gestel et al. 1998) reflects
both rifting and rotation as the uncollided area to
the east (Puerto Rico and Virgin Islands) rotates in a
counterclockwise direction. Paleomagnetic studies
of the Neogene carbonate platform in Puerto Rico

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

confirm 25 of counterclockwise rotation of the


island in late Miocene-Pliocene time (Reid et al.
1991). In other words, Mann et al. (2002) express
that this collided or impeded area of the CA
plate to the west of the Mona Passage undergoes
widespread shortening while the uncollided area
to the east of the Mona Passage in Puerto Rico
and the Virgin Islands is characterized mainly by
counterclockwise rotation about a hinge point in
the Mona Passage, broad arching of the Puerto
Rico-Virgin Islands arch, normal faulting related
to separation from the collided area, and strike-slip
faulting.
North Hispaniola (Septentrional) microplate
In much the same way as Alvarado et al. (2011)
have proposed a parallel-to-trench sliver for the
Nicaragua-El Salvador forearc (see discussion
above), a microplate has been defined in the northern
Hispaniola region between the North Hispaniola
fault (to which is associated the submarine
North Hispaniola Deformed Belt NHDB-) and
Septentrional fault (Dolan et al. 1998; Dolan &
Bowman, 2004). The best-fit elastic strain model
by Mann et al. (2002) indicates strike-slip motion
of 12.8 mm/a and 9 mm/a on those two faults,
respectively. To the east, the North Hispaniola fault
prolongs into the Puerto Rico trench, which shows
clear evidence for low-angle thrust faulting. In
turn, the Septentrional fault in Hispaniola appears
to merge with active strike-slip faults been mapped
on the inner wall of the Puerto Rico trench, the
most prominent of these being the Bunce and
Bowin faults (ten Brink et al. 2004; Grindlay et
al. 2005b). Both, the North Hispaniola and Puerto
Rico trench faults, mark the subduction of Atlantic
lithosphere beneath Hispaniola and Puerto Rico,
respectively. Dolan et al. (1998) express that highly
oblique subduction of Atlantic oceanic lithosphere
along the deformation front north of Hispaniola
and Puerto Rico is partitioned onto two major
structures: (1) the oblique left-lateral thrust fault
along the south-to-SSW-dipping interface along the
top of the underthrust Atlantic slab; and (2) the leftlateral Septentrional-northern Puerto Rico slope
strike-slip fault system. This strain partitioning
indicates that the part of Hispaniola north of the
Septentrional fault constitutes a small microplate
caught up within the plate boundary that is moving

eastward relative to the NA plate at a much slower


rate than the bulk of Hispaniola to the south (Figure
8). These authors propose that the Septentrional
fault probably merges into the southwest-dipping
decollement along the top of the subducted slab at
a depth of ~15 to 30 km.
In Hispaniola, Manaker et al. (2008) confirm from
GPS data that CA-NA oblique convergence is
partitioned between plate boundary parallel motion
on the Septentrional and Enriquillo faults in the
overriding plate and plate-boundary normal motion
at the plate interface on the Northern Hispaniola
fault. East of the Anegada Passage, conversely,
the CA-NA plate motion is accommodated by unpartitioned oblique slip on the faults bounding the
Puerto Rico block to the north (Puerto Rico oblique
subduction) and to the south (Los Muertos thrust).
An increase of convergence obliquity, which is the
case when moving westward from Puerto Rico into
Hispaniola, is enough justification for inducing
strain partitioning, as proposed by Fitch (1972)
and McCaffrey (1992). Interestingly, the transition
coincides with the subduction of the buoyant
Bahamas platform under Hispaniola, whereas
normal oceanic lithosphere descends obliquely
beneath Puerto Rico. We believe that an increase
of convergence obliquity, due to the change of
orientation of the North Hispaniola and Puerto
Rico trench at the Mona Passage, must add to a
higher plate coupling introduced by the collision of
the Bahamas platform against northern Hispaniola.
This is supported by the following statement from
Dolan et al. (1998): Pronounced along-strike
changes in structural style along the HispaniolaPuerto Rico slope record an east-to-west transition
from highly oblique, predominantly left-lateral
underthrusting of Atlantic oceanic lithosphere
(NA plate) along the 085 margin northwest of
Puerto Rico, to oblique left-lateral contractional
deformation along the 110 plate boundary off
north-central Hispaniola.
Four or five microplates within the NA-CA PBZ?
So far, four microplates (or blocks; Figure 8)
have been defined from geological, seismological
and geophysical marine data within the NA-CA
PBZ, whose existence and definition have been
progressively confirmed by numerous GPS studies

41

42

Franck A. Audemard M.

carried out over the last 25 years. From west to


east, the different entities that interact with the
neighboring CA and NA plates are: the Gnave,
Hispaniola, Septentrional and Puerto Rico-Virgin
Islands blocks. As the GPS networks in the region
have flourished and got denser, the modeling of
the ever-increasing GPS data has become more
refined. The most recently developed models are
those of Manaker et al. (2008) and Benford et al.
(2012a and b). As a matter of fact, Benford et al
(2012a and b) were in need of even proposing a
fifth block from GPS data results, inside the CA
plate; incorporation that they have named the
Nicaragua Rise block model. It would appear that,
as suggested earlier by us in this paper, a significant
portion of the bivergence of the Jamaica restraining
bend lies offshore south of the island. A suggestive
submarine promontory on the bathymetric chart
southwest of Jamaica is much appealing to make
this transpressional bend larger and wider than the
actual portion above sea level. This would rule out
the need of creating a fifth entity in this PBZ.
ATLANTIC-CA PLATE BOUNDARY
The Lesser Antilles Arc, comprising a score of
islands, is 850 km long with a radius of curvature
of 450 km. It stretches submeridionally from the
SA continental margin to the eastern termination
of the Greater Antilles (Puerto Rico and Virgin
Islands), from which is separated by the Anegada
Passage. The active island arc of the Lesser Antilles
marks the eastern boundary of the CA Plate, which
is underthrusted, in a westward-dipping subduction
zone, by the old oceanic crust of the western central
Atlantic Ocean. It is part of a wider island arc system
(the eastern Caribbean) including, to the west, a
back-arc basin (Grenada Basin) and a remnant
arc (Aves Swell or Ridge). Farther to the west, the
Venezuela Basin is one of the main oceanic basins,
with abnormally thick crust, of the CA Plate.
The Lesser Antilles trench is morphologically
continuous with the Puerto Rico trench to the
northwest, where subduction transitions from
roughly trench-normal (Lesser Antilles) to highly
oblique (Puerto Rico). The transition is marked in
the upper plate by the Anegada Passage faults.
Bouysse & Westercamp (1990) provide a thorough
review of the complex history of this island arc,

which has probably been active since the Early


Cretaceous. With the onset of the Early Eocene,
a volcanic front settled upon this Mesozoic arc
substratum. Its trace can be located from Grenada,
in the south, to Anguilla, in the north, and constitutes
the Older arc (Grenada-Grenadines-St. Vincent-St.
Lucia-Martinique-Marie Galante-Grande Terre
of Guadeloupe-Antigua-St. Bartholomew-St.
Martin-Anguilla). This volcanic line later ceased
its activity, and after several million years, the
volcanism resumed along the Recent arc, which
is still active today, and the locus of the Volcanic
Caribbees. Between Grenada and Martinique
the two arcs are imbricated; but to the north, they
diverge, and from Martinique northwards, they
are split apart by a corridor some 50 km wide, the
Kallinago depression (Bouysse, 1979). These two
northern branches have been called the Outer arc
to the east (from Marie Galante to Anguilla) and
the Inner arc to the west (i.e. the northern half of
the Volcanic Caribbees; Dominica-Les SaintesBasse Terre of Guadeloupe-Montserrat-RedondaNevi-St. Kitts-St. Eustatius-Saba). A 110 km long
submarine segment extends the Inner arc as far
as the Anegada Passage. It has been extinct since
the Late Pliocene and includes Luymes Bank and
Noroit Seamount (Bouysse et al. 1985). La Desirade
and Barbuda islands are located on the eastern rim
of the northern Lesser Antilles Mesozoic basement.
Together with the islands of the Outer arc, they
have been called the Limestone Caribbees.
To the west of the Lesser Antilles, the Grenada
Basin, some 150 km wide, separates the former
from the Aves Swell, a remnant arc extinct since
the onset of the Cenozoic era (Bouysse, 1988).
The surface contact between the Caribbean plate
and the subducted Atlantic crust is underlined by
a strong negative gravity anomaly (Bowin, 1976).
The distance between the present volcanic front
(Recent arc) and this anomaly is approximately
constant and of some 150 km, implying that is a
normal slab of intermediate dip (close to 40-45
W). Physiographi-cally, this contact corresponds to
a classical subduction trench (more than 6000 m
deep) only to the north of the latitude of Antigua,
and constitutes the southeastern extension of the
Puerto Rico trench. To the south, the trench is
progressively filled in by sediments and passes to
the Barbados accretionary wedge, which rises above

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

sea level to form Barbados Island. The accretionary


complex increases in width to the south (up to 300
km beyond the contact of the plates) and shows a
maximum thickness of 20 km (Westbrook, 1975;
1982). The Barbados complex is one of the best
examples of sedimentary accretionary wedges in
the world, and has been thoroughly studied (Moore
& Biju-Duval, 1984; Brown & Westbrook, 1987;
ODP leg 110 Shipboard Scientific Party, 1987).
It owes its importance to the huge longitudinal
sedimentary input coming from the Orinoco and
Amazon rivers.
A simple PBZ?
Feuillet et al. (2002) report active trench parallel
extension in the northern half of the Lesser Antilles
arc, north of about 15 N where the Tiburon Ridge
is being subducted, and argue that it results from
strain partitioning. Lpez et al. (2006) report a
systematic discrepancy between slip vectors of
thrust fault earthquakes at the Lesser Antilles
trench (LAT) and the predicted direction of NA-CA
convergence. They mention that a possibility has
been that the discrepancy resulted because neither
was well constrained. Estimating CA motion has
been challenging owing to the limited data along
the plates complex boundaries. Similarly, earlier
studies had few slip vectors because interplate thrust
events are infrequent. To address these difficulties,
Lpez et al. (2006) estimate a new CA-NA Euler
vector using available GPS data from sites in the
presumably stable interior of the Caribbean, and
compare the predicted velocities to a larger set of
slip vectors. The discrepancy persists, suggesting
the Northern Lesser Antilles Forearc (NLAF)
moves as a distinct entity from both the CA and NA.
For simplicity, Lpez et al. (2006) treat its motion
as a coherent block, but because GPS sites are not
within the NLAF, distributed deformation is also
possible. Although there is no geologic evidence
for the boundaries of the presumed NLAF block,
Lpez et al. (2006) conclude from the modeled
GPS data that the motions of Martinique, Barbados,
and Trinidad are similar to that of the Caribbean,
suggesting that none are on the NLAF block, and
the southern LAT is weakly coupled. Manaker et
al. (2008) found that the discrepancies between
earthquake slip vectors and GPS velocities were
small and within the error limits of both data types.

Also, the boundaries of a possible Lesser Antilles


block are poorly defined and deformation could
result from diffuse deformation within the arc rather
than slip on a single fault zone, as shown by the
broad distribution of mostly trench-perpendicular
normal faulting within the arc (Feuillet et al.
2002). Manaker et al. (2008) therefore do not favor
a separate Lesser Antilles Block in their model
geometry, while recognizing that a denser and more
precise data set may require it.
Later on, Feuillet et al. (2010) bring the missing
geologic and geophysical supporting evidence.
From a new high-resolution marine data acquired
aboard R/V Le Surot, these authors map active
normal faults offshore Montserrat in greater
detail. The main faults of the Montserrat-Havers
fault zone have cumulative scarps up to 200 m
high, and offset sedimentary layers by hundreds
of meters. They are arranged in a righ-stepping,
en echelon, transtensional array, which confirms,
after Feuillet et al. (2010), that they accommodate
the lef-lateral component of motion resulting from
slip partitioning of oblique convergence along the
volcanic arc. These N110E-trending faults even
cut across Montserrats recent volcanic complex.
Consequently, the northern Lesser Antilles forearc,
in much the same way as the northern sliver of
Hispaniola, bounded by the North Hispaniola
and Septentrional faults, and the terrane defined
by the MAVA and the Middle America trench
in the NA-CA and CO-CA PBZs, respectively,
accommodates the trench-parallel component of
oblique convergence at those PBZs. Furthermore,
both, the Lesser Antilles forearc and the sliver in
Central America, take profit of the presence of
an active volcanic arc along one of its boundary,
which weakens the upper plate (the CA plate in
both cases).
SA-CA PLATE BOUNDARY
Hess & Maxwell (1953) proposed that the
boundary was a simple dextral wrench system.
Over 40 years of neotectonic analyses, integrating
surface geology, geomorphology, microtectonics,
seismotectonics and palaeoseismology, combined
with data from conventional geological studies
and on and offshore seismic reflection data, allow

43

44

Franck A. Audemard M.

a more precise view. The more we learn about the


boundary zone, the more complex it seems.
There is wide consensus that the CA plate is moving
east relative to SA and recent GPS results strongly
support this (Freymueller et al. 1993; Prez et
al. 2001a and b; Weber et al. 2001). However,
northern South America reflects interaction of the
CA, SA and NA plates and the Panam microplate.
Deformation along the boundary is driven by
oblique convergence (Silver et al. 1975; Prez
& Aggarwal, 1981; Speed, 1985; Lugo & Mann,
1992; Russo & Speed, 1992). Today this is more
intense in the west than in the east where wrenching
dominates.
Northern Venezuela lies in the interaction zone
between the SA and CA Plate, while western
Venezuela and northern Colombia comprise
a number of interplaying tectonic blocks or
microplates (Figures 1 and 9). The CA-SA Plate
boundary from Colombia to Trinidad is over a
100 km wide active transpression zone on and
offshore northern Venezuela (Audemard, 1993;
Singer & Audemard, 1997; Pindell et al. 1998;
Audemard et al. 2000; 2005; Ysaccis et al. 2000).

Significant positive relief is present along the


Coastal and Interior ranges of north-central and
northeastern Venezuela. Further west, the southern
Caribbean boundary broadens to as much as 600
km and includes several small tectonic blocks or
microplates (Figures 1 and 9). The Triangular
Maracaibo Block TMB- is bounded by the leftlateral Santa Marta-Bucaramanga fault SMBF- in
Colombia and the right-lateral Bocon fault BFin Venezuela. The dextral Oca-Ancn fault OAFseparates it in the north from the Bonaire Block
BB-. Both, the Maracaibo and Bonaire blocks,
are being extruded northward. Extrusion is driven
by the collision and suturing of the Choc Block
against the Pacific side of northern South America
(Duque-Caro, 1978; 1990; Audemard, 1993; 1998;
2009), confirmed by GPS plate motion studies
(Freymueller et al. 1993; Kellogg & Vega, 1995;
Kaniuth et al. 1999; Trenkamp et al. 2002). The
Maracaibo and Bonaire blocks override the CA
Plate at the Leeward Antilles subduction LAS.
North of the Netherlands Leeward Antilles, south
dipping, amagmatic, flat subduction has developed
in the last 5 Ma. The onset of this escape tectonics
is a key issue, which has already been discussed.

Figure 9. Tectonic blocks in the CA-SA PBZ defined by Audemard (2002) from geologic data and geodynamic criteria.
Abbreviations and equivalences with Figure 1: CDNP = NPDB Northern Panam deformed belt; FB = BF Bocon Fault; FSMB =
SMBF Santa Marta-Bucaramanaga Fault; FOA = OAF Oca-Ancn Fault; SAS = LAS Leeward Antilles Subduction; SFR =RFS Romeral
Fault System.
Figura 9. Bloques tectnicos definidos en la zona de frontera de placas Caribe-Amrica del sur por Audemard (2002) a partir de
datos geolgicos y criterios geodinmicos. Abreviaturas: CDNP = NPDB Cinturn de deformacin Nor-Panameo; FB = BF Falla
de Bocon; FOA = OAF Falla de Oca-Ancn; SFR =RFS Sistema de Fallas de Romeral; SAS = LAS Subduccin de las Antillas de
Sotavento; FSMB = SMBF Falla de Santa Marta-Bucaramanga.

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

Along the CA coast of Venezuela


Much of present day dextral slip along the southern
Caribbean boundary seems to be focused along the
major Bocon-San Sebastin-El Pilar-Los BajosWarm Spring fault system where geological and
geodetic data indicate a slip rate of 8-10 mm/a.
Most authors see this system as the plate boundary
(Hess & Maxwell, 1953; Schubert, 1979; Stephan
1985; Prez et al. 2001a and b). Others note that
deformation is distributed over a 100 km or more
wide zone and interpret orogenic float in the Andes
(Audemard, 1991; Jcome, 1994; Audemard
& Audemard, 2002), across the Falcn Basin
(Porras, 2000) and eastern Venezuela (Ysaccis,
1997; Ysaccis et al,. 2000). According to this
understanding, the zone is flanked by both A and
B-type subduction. Yet others recognize southeastdirected A-subduction or under-thrusting below the
Mrida Andes (Kellogg & Bonini, 1982; De Toni &
Kellogg, 1993; Colletta et al. 1996; 1997).
The PBZ exhibits strain partitioning from the
Mrida Andes in the west to the Interior range in
the east. In the Andes, partitioning occurs between
the dextral Bocon fault and thrust faults on both
mountain flanks (Audemard & Audemard, 2002).
The north-central Coastal Range also exhibits
strain partitioning. Dextral slip occurs in the core
along the San Sebastin, La Victoria faults and
minor synthetic Riedel shears while transverse
shortening is accommodated by frontal thrust faults
in the south (Guarumen basin; Audemard, 1999).
A sub-sea, mirror thrust fault system may exist to
the north and the easternmost portion of the LAS
zone must account for some shortening as well.
Farther east, strain partitioning involves NNWSSE-trending shortening over a 250 km wide zone
from north of La Blanquilla to the active thrust
front of the Interior Range. Slab detachment (Russo
and Speed, 1992; Russo, 1999; Bezada et al. 2010)
associated with an incipient A-subduction, results
in the largest onshore negative Bouguer anomaly
in the world, south of the Interior Range. Dextral
slip concurrently occurs along the main east-west
striking El Pilar fault and along the NW-SE striking
Los Bajos-El Soldado faults and minor parallel
and/or synthetic Riedel shear faults. Partitioning
might also occur on Trinidad.

Although transpression is the dominant process


at plate boundary scale, transtension also occurs
in two pull-apart basins, the Cariaco Trough
(Schubert, 1979; 1984; Escalona et al. 2011) and
the Gulf of Paria (Babb, 1997; Ysaccis, 1997) at
the ends of the El Pilar fault. The age of opening
of these basins constrains the timing of wrenching
onset between 15 and 10 Ma in eastern Venezuela.
Most of El Pilar fault strike-slip motion transfers
via the Los Bajos and El Soldado synthetic Riedel
shears to the Warm Springs fault of the Central
Range of Trinidad (Weber et al. 2001). However,
this shallow crustal deformation does not exclude a
deeper plate boundary, as indicated by instrumental
seismicity and as suggested by Prez & Aggarwal
(1981) and Sobiesiak et al. (2002; 2005).
Present day motion
Several authors have predicted that the CA Plate
moves east at about 20 mm/a relative to South
America (Jordan, 1975; Rosencrantz et al. 1988;
Stein et al. 1988; Calais et al. 1992; DeMets et al.
2010) and GPS studies carried out in the CA-SA
PBZ have confirmed this (Prez et al. 2001a and b;
Weber et al. 2001; Trenkamp et al. 2002; Jouanne
et al. 2011). In the east, dextral slip along the plate
boundary transfers to subduction in the Lesser
Antilles, although the detailed tectonic setting
remains still unclear.
GPS slip magnitudes remain rather constant from
west to east but orientations change. ESE oblique
convergence occurs between San Andrs Island,
east of Honduras, and stable SA, confirming
predicted N075W convergence (Jordan, 1975;
Minster & Jordan, 1978) and seismotectonic
models (Pennington, 1981; van der Hilst & Mann,
1994). Data from eastern Venezuela indicate almost
pure wrenching at the CA-SA PBZ [086 2 with
respect to the Central Range of Trinidad (Weber
et al. 2001) and 084 2E with respect to South
America (Prez et al. 2001a)].
Vector orientations indicate transtension, as
postulated by Robertson & Burke (1989), Algar
& Pindell (1993) and Pindell et al. (1998), and
inferred by Weber et al. (2001). Stress tensors of
Colmenares & Zoback (2003) indicate the same.
However, Choy et al. (1998) determined stress

45

46

Franck A. Audemard M.

tensors indicating NW-SE compressive wrenching.


Several authors came to this same result using
geological data (Beltrn & Giraldo, 1989;
Audemard, 1993; 2000; Audemard et al. 2000;
2005). Lately, Jouanne et al. (2011) derive slip
vectors that show pure east-west dextral motion
across eastern Venezuela with a minor component
of shortening; all in perfect agreement with the
geological tensors and focal mechanism solutions
compiled by Audemard et al. (2005).
GPS results published by Prez et al. (2001a, p. 70,
Figure 1) indicate compression across the boundary
south of the main dextral system (Audemard et al.
2005) where GPS rates are as high as 20-25% of
the main dextral rate (18-20 mm/a), the same order
of magnitude as those calculated from long-term
geological criteria (Audemard, et al. 2000). In this
area strain is partitioned between the major eastwest dextral faults and ENE-WSW trending thrusts
and folds of the Interior Range and the MargaritaBlanquilla platform.
Slip of 8-10 mm/a along the dextral Bocon-San
Sebastin-El Pilar-Warm Spring faults, from the
southern Mrida Andes to Trinidad, accounts for
up to 50% of the 20 mm/a dextral relative motion
at the CA-SA PBZ. A recent study (Jouanne et al.
2011) estimates that the other half of the entire
PBZ slip rate, across the El Pilar Fault region, may
take place as creep. Greatest slip along the Bocon
fault from Late Quaternary geological markers is
about 9-10 mm/a (Soulas, 1986; Audemard, 1997a;
Audemard et al. 1999; 2008). The Warm Spring
fault also accounts for half of the dextral motion
across Trinidad.
Prez et al. (2001a) indicate that strain is distributed
across a zone at least 110 km wide, with 68%
(almost 14 mm/a) of this occurring in a 30 kmwide fault zone involving the El Pilar fault and
sub-parallel faults to the north. Weber et al. (2001)
claim that the Warm Springs fault is the southern
plate boundary in Trinidad. However, there is still
8 mm/a to be accommodated between southwestern
most Trinidad and continental SA. This implies the
presence of another major fault south of Trinidad,
as proposed by many authors (Soulas 1986; Beltrn,
1993; 1994; Audemard et al. 2000; Audemard,
2002; Audemard & Audemard, 2002).

Recently, Prez et al. (2011), from new GPS


mea-surements, reveal that the 20 2mm/a of
eastward displacement of the CA plate relative to
SA in northwestern and north-central Venezuela is
partitioned into a dextral shear velocity of 12 2
mm/a along a locking depth of 14 4 km beneath
the surface expression of the Bocon Fault, with
convergence normal to its northeast strike at 12
to 16 mm/a. Of this convergence, about a third
concentrates in the Andean region close to the
Bocon fault and manifests itself geologically as
thrust slip on a narrow belt of thrust faults nearly
parallel to it, running ~25 km away along both
sides of the Bocon fault main trace.
North Andes Block
Northwestern SA is a broad convergent plate
boundary zone characterized by active seismicity,
a volcanic arc, subduction, and an on-going arccontinent collision. The North Andes Block as
originally defined by Pennington (1981); NABis bounded by the Colombian-Ecuador Trench
and the Panam Block to the west, the South
Caribbean Deformed Belt SCDB- to the north,
and the Bocon fault and East Andean fault zones
to the east (Figure 9). It is very likely that this
eastern boundary will tend to simplify itself in the
future, as suggested by the presence of Quaternary
active faulting cutting across the Colombian
and Venezuela Llanos, through the Meta fault in
Colombia, which controls a very long strech of
the river of same name, and Tala fault in eastern
Venezuela, this latter resulting from tectonic
inversion of the Espino graben southern bordering
fault (Figure 1). Cenozoic deformation in this broad
zone has been produced by the converging NZ, SA
and CA plates and the Panam microplate (Kellogg
& Bonini, 1982).
Geodetic measurements using GPS from the Central
and South American (CASA) GPS project show
that the NZ oceanic plate is rapidly converging
with stable SA (Freymueller et al. 1993; Trenkamp
et al. 2002). The convergence direction is slightly
oblique to the Colombia-Ecuador Trench. The
aseismic Carnegie Ridge, produced by the passage
of the NZ plate over the Galapagos hotspot, is
being subducted in the Ecuador-Colombia trench.
CASA measurements also suggest that a large part

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

of the northern Andes is escaping to the northeast


relative to stable SA at a rate of at least 6 2 mm/a.
Tapponnier et al. (1982), as already argued for by
Audemard & Audemard (2002) and Audemard
(2003), using plane rigid indentation experiments
on unilaterally confined blocks of plasticine, were
able to model intra-continental deformation and the
evolution of strike-slip faults due to the collision
of India and Asia. The faults that develop, allow
the escape of the detached block in the direction
of the free boundary. Proposed driving mechanisms
for the escape of the NAB already discussed
earlier, are: collision with the Panam Arc, rapid
oblique subduction of the NZ plate, and the
subduction of the aseismic Carnegie Ridge at the
Ecuador-Colombia Trench.
NAB: One or several blocks?
After regional Caribbean maps (Case & Holcombe,
1980; Stephan, 1982; Mann & Burke, 1984; Mann
et al. 1990; Mascle & Letouzey, 1990; Audemard,
2002; 2003) and geodynamic models (Stephan et
al. 1990; Audemard, 1993; 1998; 2009), it appears
that the NAB as defined by Pennington (1981), in
fact amalgamates several discrete blocks, which
have a certain degree of kinematic independence,
as shown by the GPS results of Freymueller et
al. (1993) and Trenkamp et al. (2002). Audemard
(2002), partly based on Freymuellers results,
proposes a block individualisation (Figure 9).
As a matter of fact, the plate boundary in western
Venezuela, Colombia and Ecuador is eventually up
to 1000 km wide and comprises a set of discrete
tectonic blocks, independently moving among
the surrounding larger plates (CA, SA, NZ and
Panam microplate), among which the Maracaibo
Block stands out for its triangular shape with one
of its apexes pointing due south. The Triangular
Maracaibo Block TMB- is bounded just by rather
rectilinear strike-slip (SS) faults: the left-lateral
SS Santa Marta-Bucaramanga Fault SMBFS- in
Colombia and right-lateral SS Bocon Fault BFin Venezuela and separated on the north from the
Bonaire Block BB- by the right-lateral SS OcaAncn Fault OAF-. In turn, the BB is bounded
on the north by the Curacao Ridge; renamed
more recently as the SCDB. The Leeward Antilles
Subduction LAS-, responsible for the SCDB, may

well be an inherited structure of the Cretaceous


active margin of the SA continent. Escalona &
Mann (2011) instead consider this feature may
have functioned as a back-thrust of the Caribbean
nappes during their emplacement onto the SA
margin, spanning the Eocene-Early Miocene and
overriding then the back arc (towards the CA plate
sea floor). Both TMB ans BB are being extruded
NNE and are overriding the CA plate north of the
Leeward Antilles Islands, where a young (mainly
in the last 5 Ma), south-dipping, amagmatic, flat
subduction (LAS) has been forming in recent times.
Extrusion of these blocks is to be related to the
collision of the Panam Arc against the Pacific side
of northern SA and its later suturing (Audemard
1993; 1998; 2009), although we do not completely
rule out that the buoyant Carnegie Ridge may have
partly contributed to the escape. Neither is ruled
out that the LAS may have accommodated part of
the Americas convergence since the Oligocene.
Besides the TMB and BB, Audemard (2002) also
splits apart the Choc Block from the Panam
microplate, both originally constituting the larger
Panam Arc; before the progressive clockwise
docking of the southern part of the Panam Arc
(new Choc Block CB-) against the northwestern
SA margin. Consequently, the new definition of the
NAB refers to the rest of the original block defined
by Pennington (1981) after individualizing TMB,
BB and CB. But, is this NAB a single entity?
How many subductions under the NAB?
Since the NAB, as defined in this contribution,
does not extend northward to the SCDB anymore,
except for a short portion between the Baud suture
(Uraba Gulf) and the SMBFS (city of Santa Marta),
this excludes the LAS, which shall be discussed
apart later.
The NAB is for a very significant portion defined
to the west by the Ecuador-Colombian Trench,
under which plunges the oceanic NZ plate. In spite
of changing from flat-slab to normal subduction,
depending on whether the NZ plate brings along
a ridge (i.e., Carnegie, Malpelo ridges) or not, the
Ecuador-Colombian Trench is rather continuous
up to Panam. However, the NAB is limited on
the west by the Choco-Baud suture to the north

47

48

Franck A. Audemard M.

of the ENE trending Garrapatas fault, which is the


southern tip of the CB (Figures 5 and 9). Audemard
(2002) indicates that the effect of the Choc-Baud
suturing and indentation can be seen as far inland
in Colombia as the Eastern Cordillera of Colombia.
North of a line at 4 N latitude, comprising the rightlateral Garrapatas, Ro Verde and Ibagu faults,
which are the southern edge of the indentation at
crustal levels, the following major changes are
reported by Audemard (2002): (1) The Eastern
Cordillera is significantly much wider to the
north, as well as it exhibits higher relief; (2) the
Llanos foothills of the Eastern Cordillera are more
transpressional to the north and more wrenching to
the south; (3) the Cauca Valley narrows significantly
north of the Garrapatas Fault, squeezed between
the Western and Central cordilleras, and the Cauca
River has substantial difficulties to keep flowing
north because the valley floor is being raised;
(4) The axis of the Central Cordillera is shifted
dextrally by the dextral Ibagu fault; and (5) The
Cauca Romeral fault system (RFS) appears to
change sense of slip at that latitude. The fault trend
is arcuate with convexity to the east and tends to
run parallel to the Ecuador-Colombian Trench. This
major tectonic accident seems to accommodate the
parallel-to-trench strike-slip component of strain
partitioning, due to NZ-SA convergence obliquity
(Audemard, 2002). To the north of latitude 4 N,
the trench trends NNW-SSE. So do SS faults. This
leads to left-lateral SS along the Cauca-Romeral
Fault System, instead of right-lateral to the south of
latitude 4 N. On top of that, the left-lateral sense of
slip to the north is required to allow the NAB escape
towards the NNE between the Cauca-Romeral fault
system on the west and the EC Foothills System on
the east, although not as fast as the TMB, which is
not directly under the influence of the CB and is

closer to the northern free boundary (to the SCDB).


To some extent, the NAB moves slower to the north
because it is more compressed and refrained by the
CB indentation.
Although the Ecuador-Colombian Trench appears
continuous along trend, the subduction slabs appear
torn at the latitude of 5.6 N, as suggested by the
eastward jump of the intermediate depth seismicity
(Pennington, 1981; Taboada et al. 2000; Ojeda &
Havskov, 2001; Meja & Meyer, 2004; Corredor,
2003; Corts & Angelier, 2005; Vargas & Mann,
2013). An east-west tear cutting the subduction
slab has been proposed by Vargas & Mann (2013;
Figure 10). They have named it the Caldas tear
based on the location in the Caldas department of
Colombia. The nearly 240 km long Caldas tear is
a narrow, east-west trending boundary between
two subducted slabs of different dip and nature.
The northern zone is the down-dip extension of
the Panam Arc, has a shallower dip, and is not
associated with active arc volcanism. It corresponds
to a thickened oceanic crust, with the typical
Caribbean oceanic plateau affinity. The southern
zone has a steeper dip and is associated with an
active volcanic front. This volcanic arc relates to
the NZ slab. It is a normal subduction where the
slab is conveying a normal oceanic crust. Earlier,
using the distribution of earthquakes deeper than 80
km, Ojeda & Havskov (2001) proposed that such
discontinuity represented a boundary between two
subducted slabs with differing dips and strikes: the
northern subduction zone, called the Bucaramanga
subduction zone, has a shallower dip (27) and
more northeasterly strike, and the southern, called
the Cauca subduction zone, has a steeper dip (3540) and a more northerly strike.

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

Figure 10. Block diagram showing seismic surface estimated by interpolation and filtering of local earthquakes (h 10 km; from
Vargas and Mann, 2013). Blue lines, shore line. Bold black lines, limits of the convergent margins. Bold red lines, the southern border
of the Panama indenter that includes the Sandra Ridge and the Caldas tear. Triangles, red (active) and green (inactive) volcanoes.
Orange dashed lines, wireframe model suggested for indicating the subduction geometry of the Caribbean plate
Figura 10. Bloque 3D mostrando superficie ssmica estimada a partir de la interpolacin y filtrado de sismos locales (h 10 km;
segn Vargas and Mann, 2013). Lnea azul, lnea de costa. Lnea gruesa continua, fronteras de mrgenes convergentes. Lnea
gruesa roja, el borde meridional del indentor de Panam, que incluye el alto de Sandra Ridge y el desgarre Caldas. Tringulos:
rojo (volcanes activos) y verde (volcanes inactivos). Lnea naranja segmentada, lnea de contorno para definir geometra de la losa
de subduccin caribea

Although not matching perfectly, it would appear


that all major changes at crustal levels mentioned
at the latitude of 4N by Audemard (2002), with
the rare exception of the volcanic arc, which is
the surface expression of partial melting at around
100-120 km in depth and then tightly linked to the
slab geometry at depth, are reflecting the same
major geodynamic process as the Caldas tear: The
collision and indentation of the southern tip of the
Panam Arc (CB) against northwestern SA. This
process necessarily implied a backward jump of
subduction from the east to the west board of the
Baud-Choc Block sometime in the Pliocene.
This implies in turn that at the latitude of the CB,
there are two subduction slabs sinking in the mantle
an eastern flat slab of CA affinity dying out as
CB-NAB suturing goes on, and a younger western
slab taking the relay in activity-, while there is
only one to the south of the southern tip of the CB:
the normal NZ subducted slab. Also keep in mind
there is a third slab in a more northerly position,
offshore north of Venezuela. This subduction (or
underthrsuting), induced by NNE escape of the BB
and TMB, is to be presented next.

Leward Antilles Subduction LAS-: True or


induced subduction?
Earlier in this paper, we have introduced the
discussion about the nature of this subduction,
which displays a similar condition to the ones of
the NPDB and the Hispaniola-Puerto Rico (Los
Muertos) trench.
Many authors have postulated that subduction is
occurring on the northern boundary of BB from the
Uraba Gulf in the west to the Los Roques Canyon in
the east. This subduction trends roughly east-west,
has CA oceanic plateau affinity, dips south and is
very young in age. Several evidences of geological,
geophysical and seismological nature have been
presented to support the existence of LAS: (1)
Seismic profiling and bathymetric surveying across
the Curacao Ridge image a typical accretionary
wedge and its forearc basin (Silver et al. 1975;
Talwani et al. 1977; Mascle et al. 1979; Kellogg
& Bonini, 1982; Mascle et al. 1986); (2) An
infrequent seismicity of intermediate depth below
50 km deep has been recorded, evidencing mainly
southeast dipping thrusting (Malav & Surez,

49

50

Franck A. Audemard M.

1995; Audemard et al. 2005; Palma et al. 2010). The


spatial distribution of these hypocenters allowed
to delineating a south-dipping slab (Orihuela &
Cuevas, 1992); and c) A cold slab, 450 km long
and 16 southeast dipping, has been recognized
from seismic tomography (van der Hilst, 1990; van
der Hilst & Mann, 1993; 1994), plunging beneath
northwestern Venezuela and northern Colombia.
Kellogg & Bonini (1982), just from seismicity, had
proposed a similar length (400 km).
But very recently, new findings have seen light.
Bezada (2010) and Bezada et al. (2010), by
analyzing teleseismic traveltimes with frequencydependent kernels, produced a 3D P-wave
velocity perturbation model. The model depicts
the subduction of a section of the CA plate under
SA with an east-southeast direction. The CA
subducting slab penetrates, after these authors, the
mantle transition zone, affecting the topography
of the 410-km and 660-km discontinuities (Figure
11). However, the imaged cold slab considerably
differs in geometry and lateral extent from the van
der Hilsts proposal. In this new model, the slab at
any depth either at 95 km, 375 km deep or even
deeper (Bezada, 2010)-, never appears north of the
vertical projection of the Oca-Ancn fault system.
In other words, it only exists beneath the Maracaibo
block (TMB) at depth. Besides, it rather dips ESE,
instead of SSE as early proposed.
From Audemards geodynamic evolutionary models
(Audemard, 1993; 1998; 2009), it seems clear that
LAS keeps no relation with the subduction that led
to collision-suturing of the Panam Arc (the trailing
edge of the CA plate) against western SA along the
Choc Block and San Jacinto terranes. This old
Caribbean subduction, fully functioning prior to
the Panam Arc-SA collision, had to trend more
sub-meridionally along the western coast of SA.
This major plate boundary was rather straight and
was as old as Cretaceous in certain models (Pindell
& Dewey, 1982; Beck, 1985; Pindell et al. 1988;
Stephan et al. 1990; Taboada et al. 2000). This
old subduction still trends subparallel to both the
Ecuador-Colombian Trench and the eastern border
of the Choc block. Instead, Audemard (1993;
1998; 2009) proposes that LAS must come much
later into the regional evolution, sometime around
5-3 Ma ago, as a direct consequence of the NNE-

directed BB-TMB escape. This two-subduction


model is reutilized by Taboada et al. (2000), who
show a 3-D model with two subduction slabs in
this region, both composed of abnormally thick
CA oceanic lithosphere, at different depths in
asthenosphere. As advanced by Audemard (1993;
1998; 2009), this suggests that the slabs are of
different age, being LAS at an incipient stage.
Besides, this author proposes that LAS roots at
the northern tip of the old Caribbean subduction
(ending somewhere near the city of Santa Marta;
at a clear v-shaped break in the LAS trench; see
Mascle & Letouzey (1990) map and Figures 1, 2, 5
and 9) and extends progressively eastward around
the northern border of the BB. By assuming that a
30 S-dipping slab has descended at 400-450 km in
depth below northwestern SA (Kellogg & Bonini
1982), Audemard (2009) estimated a convergence
not larger than 100 km (using 30 dip from Kellogg
& Bonini, 1982) or 170 km (using 17 dip from
van der Hilst & Mann, 1994). These estimates are
certainly upper bounds, since they were imputed to
the LAS, whose slab does not sink that deep. As
a matter of fact, the study of Bezada (2010) and
Bezada et al. (2010) can confidently illuminate this
younger subduction only to a depth of about 100
km and only present north of the OAF (Figure 7).
Summarizing, this newly imaged geometry by
Bezada (2010) and Bezada et al. (2010) of this
slab (Figure 11) seems to reflect the shape at
depth of the subduction that led to the collision
of the southern end of the Panam Arc against
NW SA, still sinking into the asthenosphere. The
surface trace of this extinct subduction must be
squeezed onshore at present day between the CB
and the NAB, corresponding to the Baud-Choc
suture zone. Its northward extension should not go
beyond the city of Santa Marta and die out against
the OAF; which implies that this major feature at
depth has acted as a tear between both subductions,
as well as the boundary between the BB and TMB
at shallower structural levels. As suggested by
Audemard (1993; 1998; 2009), this long lived
subduction is in no way the LAS. It is much older.
Instead, LAS is an induced subduction due to
escape tectonics occurring since the Pliocene. An
additional argument favoring this is the LAS extent,
from near the city of Santa Marta on the west to Los
Roques Canyon on the east. These two localities

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

are intercepted by the northward projections of


the SMBF and the BF, respectively. However, this
interpretation does not rule out a certain prior role
in accommodating either convergence between the
Americas since the Oligocene or shortening during
the Caribbean nappe emplacement as a backthrust
during between the Eocene and the Middle
Miocene. In some way, we allow to inheritance of
this LAS up to a certain extent.
Breakdown of the NAB

Figure 11. Block diagram showing shape of the Caribbean


subducted slab under the Triangular Maracaibo block (modified
by Maximiliano Bezada from Bezada, 2010 and Bezada et al.
2010). View from the SE. Contour lines of the slab are at 270
(red) and 480 (blue) km depth
Figura 11. Bloque 3D mostrando la geometra de la losa
Caribe subducida bajo el bloque Triangular de Maracaibo
(modificado por Maximiliano Bezada a partir de Bezada,
2010 y Bezada et al. 2010). Vista desde el SE. Se muestra
lneas de contorno de la losa a 270 (roja) y 480 (azul) km de
profundidad, para una mejor ubicacin relativa bajo Venezuela

The original NAB proposed by Pennington (1981)


already contains the Baud-Choc CB-, Bonaire
BB-, Maracaibo TMB- blocks, and the new
NAB. The new NAB is defined by the EcuadorColombian Trench and the eastern limit of the CB
on the west and a score of faults on the east, which
comprises the Dolores, Pallatanga, Algeciras and
Guaicramo faults (Figures 5, 9).
For the development of the Servicio Geolgico
Colombianos GEORED Project, Mora-Pez
(2011) has discretized 20 tectonic blocks in
Colombian territory surrounded by the larger

NZ, CA, SA and Panam Microplate, of which


2 are contained in the BB, 4 in the TMB and 1
corresponds to the CB (Figure 12). This implies
that NAB on its own amalgamates 13 minor
tectonic blocks in Colombia, following this block
segmentation. This block individualization for the
GEORED Project has been based on 3 assumptions
(GEORED, 2009; Lpez, 2010; Mora-Pez, 2011):
(1) A tectonic sub-block represents a homogeneous
crustal entity limited by main active faults, and
it is thought to possess a characteristic pattern of
internal deformation; (2) For some fault segments,
current kinematics, rupture dimensions and
terminations remains unclear; moreover, detailed
palaeoseismological and neotectonic studies are yet
scarce. Due to this, it has been necessary to make
some general assumptions about fault extension,
geometry and sense of slip from geophysical
interpretation, geological maps and reports; and (3)
Some regions at plate interface are expected to be
more heterogeneous and to have higher sub-block
segmentation. But, owing to the fact that some of
those places have not been studied in depth and are
difficult to access, the polygons of the sub-blocks
are only approximately outlined.
More recently, for modeling the velocity vectors
of the Panam Microplate and neighboring larger
plates (CO, NZ, SA and CA), a simplification
of this model has been generated (Figure 13),
adjusting the block geometries to comply with the
vectors estimated by Trenkamp et al. (2002).
INTERNAL CA PLATE DEFORMATION:
SINGLE OR SEVERAL BLOCKS?
Although the interior of the Caribbean plate is
relatively aseismic in comparison to its boundaries,
several indications of active deformation well
within the plate interior include: (1) occasional
intraplate earthquakes (Molnar & Sykes, 1969;
Kafka & Weidner, 1979), with magnitudes as
large as M 5.1; (2) intraplate N-S trending normal
faulting affecting young sediments in the Nicaragua
Rise seen on reflection profiles and displayed on
the Case & Holcombe (1980) map (Holcombe et al.
1990); and (3) strike-slip mechanisms for intraplate
earthquakes (Molnar & Sykes, 1969; Kafka &
Weidner, 1979). In addition to this, DeMets et al.
(2010) retake the issue once raised by many authors

51

52

Franck A. Audemard M.

and use it as motivation: What is the role of the


Beata Ridge in the internal CA plate deformation?
Driscoll & Diebold (1998) pose the problem: One
or two plates? but referring to the CA plate, with
respect to the Beata Ridge role. We herein pretend

to bring to live this discussion again on another


major internal CA plate feature instead: The Hess
Escarpment; without leaving aside the already
existing one.

Figure 12. Tectonic blocks in NW South America defined by Lpez (2010) in the frame of the Colombian GEORED Project. See
discussion in text. Block labelling: 1. Alta Guajira; 2. Baja Guajira; 3. Santa Marta-Cuenca del Cesar-Ranchera; 4. Serrana de
Perij; 5. Cuenca de Maracaibo; 6. Sin-San Jacinto; 7. Valle Inferior del Magdalena; 8. Macizo de Santander; 9. Pamplona; 10.
Choc; 11. Cordillera Central-Septentrional; 12. Serrana de San Lucas-Bajo Magdalena; 13. Valle Medio del Magdalena; 14.
Cordillera Oriental-Flanco Oeste; 15. Cordillera Oriental-Flanco Este; 16. Cauca-Nario; 17. Medio Cauca; 18.Valle del CaucaPata; 19. Valle Superior del Magdalena; 20. Macizo de Garzn-Serrana de La Macarena
Figura 12. Bloques tectnicos en el noroeste de Amrica del sur definidos por Lpez (2010), en el marco del proyecto colombiano
GEORED. Referirse a discusin en el texto

Figure 13. Tectonic blocks in the CA-NZ-SA PBZ and Panam microplate, defined for GPS data modelling, as a simplification of
block individualization of figure 12 and in agreement with figure 9. Basemap after Trenkamp et al. (2002)
Figura 13. Bloques tectnicos definidos en la zona de frontera de placas Caribe-Amrica del sur-Nazca y microplaca de Panam,
a fin de modelaje de data GPS, como una simplificacin de la definicin de bloques de la figura 12 y en corcondancia con la
figura 9. Mapa base de Trenkamp et al. (2002)

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

Hess Escarpment
The Hess Escarpment forms a prominent
bathymetric break between the Nicaragua Rise to
the north and the Colombian Basin to the south
(Figures 1 and 2). It is poorly studied and is known
mainly from bathymetric surveys and local seismic
profiling. Despite the fact that the Hess Escarpment
is not associated with any seismicity (issue that
shall be argued against later in this paper), Mann
& Burke (1984) predict that it may have a rightlateral oblique-slip component on the basis of:
(1) its linear definition of the northern margin of
an area of late Neogene deposition occupying the
floor of the Colombian Basin; and (2) the stepping
of offsets at 81 and 74W (offset at 81 W steps
left and appears to be a restraining bend, and offset
at 74W steps right and appears to be a pull-apart).
The Hess Escarpment appears to terminate to the
northeast along the western edge of the Beata
Ridge, while its western end appears to terminate
in Central America near an area of Quaternary
alkaline volcanism (Wadge & Wooden, 1982)
and near the southern end of the back-arc basin
(Nicaragua Depression) behind the Middle America
frontal arc. Continuous seismic profiling from the
Venezuelan Basin across the Beata Ridge and Hess
Escarpment to the Nicaragua Rise indicates that the
prominent seismic reflectors A (early Paleogene)
and B (late Cretaceous) are continuous across
the Hess Escarpment (Moore & Fahlquist, 1976).
Mann & Burke (1984) coincide in saying that the
Hess Escarpment irrespective of any earlier history,
appears to be a major late Neogene structural
boundary in the Caribbean interior, although the
amount of present-day displacement appears small.
Compartmentalization by the Hess Escarpment
In that sense, at very large scale, two conspicuously
distinguishable uneven halves are noticeable at
first sight in the Caribbean region (Figures 1 and
2). These two large units are split apart by the
Hess Escarpment, which runs ENE-WSW in the
CA plate interior. The northern unit shows simpler
plate boundaries, whereas the rest is dominated by
block tectonics around the entire periphery, except
for the rather simple and normal Atlantic oceanic
subduction beneath the eastern CA plate. In fact,
the two PBZs of the northern unit correspond to:

(1) on the north, the transform fault system that


bounds the Cayman sea-floor spreading center; a
major transtensional relay between two subparallel
left-lateral strike-slip faults, with generation of
oceanic floor; and (2) on the southwest, the COCA plate boundary, where normal ocean-floor
subduction takes place. The only complication here
is the presence of a trench-parallel sliver moving
NW and parallel to the trench, which denotes that
the PBZ is undergoing strain partitioning. On the
contrary, the southern CA unit exhibits 3 areas
where the continental or arc masses are overriding
the thickened CA plate, namely the NPDB, SCDB
and Muertos trench. These correspond to the
3 festoons described by Stephan et al. (1986):
Panam, Venezuela and Hispaniola-Puerto Rico
psedosubductions, respectively. Different mechanisms have been evoked to explain the presence of
these induced subductions, as discussed earlier in
this paper.
Effects of the Hess Escarpment on the CA PBZs
In the PBZs themselves, the presence of the Hess
Escarpment is also influential and noticeable. For
instance, at its SW projection across the Middle
America land bridge, north and south of its
alignment, the size and width of Middle America
drastically changes (Figures 1 and 2). North of the
alignment, Middle America is large and appears
robust (corresponding to Nicaragua, El Salvador
and Guatemala) while to the south is in fact no
more than a narrow land bridge along Costa Rica
and Panam. Geologically speaking, the juxtaposed
blocks are also different. Baumgartner et al. (2008)
propose that the Chortis (Mesquito)-Chorotega
boundary is in the southwestward projection of the
Hess Escarpment into central Costa Rica. Linkimer
et al. (2010) express that this boundary does not
conflict with geology and Vp/Vs estimates. In
general, Vp/Vs is higher on the Chorotega side
(south of the projection), although more Vp/Vs
data will be needed to determine if its location is
visible from Vp/Vs changes. These authors also
suggest that this boundary continues on the western
margin of the Nicoya Gulf, separating the Mesquito
and Nicoya terranes and following the boundary
suggested by Hauff et al. (2000), who consider
the Nicoya Peninsula as part of the Chorotega
Block and both the Tortugal and Santa Elena

53

54

Franck A. Audemard M.

areas as part of the Chortis Block (i.e., Mesquito


Terrane). Nor does the northeastern projection
differ. This eventual northeastward prolongation
roughly runs where the Mona Passage meets the
North Hispaniola-Puerto Rico trench. In other
words, where the North Hispaniola Fault bends
25 counterclockwise to continue eastward along
the 75N trending Puerto Rico oblique subduction.
It coincides with the transition from a fully
partitioned (Hispaniola) to non-partitioned (Puerto
Rico) plate boundary. On top of that, it overlaps
with where the surface expression of a buoyant
Bahamas platform disappears. As discussed earlier,
the incorporation of the buoyant Bahamas platform
into the subduction zone results in the tectonic
pinning of Hispaniola, whereas Puerto Rico moves
eastwards at the full Caribbean plate velocity. This
in turn requires extension between Hispaniola and
Puerto Rico, consistent with GPS data (Jansma et
al. 2000) and offshore observation of active normal
faults in the Mona Passage (Grindlay et al. 1997;
van Gestel et al. 1998). At scale of the entire NACA PBZ, this intersection also coincides with:
(1) The double bend that this PBZ zone follows
around northern Hispaniola; (2) The major change
from pure transform on the west to (partitioned
or not) oblique subduction on the east; and (3)
Change in orientation of the Mesozoic Caribbean
Arc (MCA in the sense of Bouysse, 1988) from
its original NW-SE trend, before being bent 25
counterclockwise. Is it a coincidence that such a
conspicuous and long feature tends to end where
these major PBZ changes take place? If assumed
that the Hess Escarpment in the Neogene allowed
a significant eastward migration of the MCA, after
colliding Cuba and Northern Hispaniola with the
buoyant Bahamas platform, it would imply that
this transform boundary used to slip left-laterally;
conversely to Mann & Burke (1984)s proposal
(right-lateral slip of motion based on potential
trasntensional and transpressional geome-tries).
This would also account for the late MiocenePliocene rotation undergone by the Puerto RicoVirgin Islands block, described by Reid et al. (1991)
and the very slow oblique slip rate (or no relative
motion) of the Los Muertos trench south of Puerto
Rico, while the MCA kept migrating eastward and
bending counterclockwise. It somehow came in help
of the Cayman spreading center and its associated
transform boundary faults, in accommodating left-

lateral motion. If this assumption happens to be


right, the left-lateral strike-slip rate of the Hess
Escarpment fault system must have been similar to
the opening rate proposed for the Mona passage of
about 2-3 mm/a, and less than 5 mm/a, by Jansma
et al. (2000) and Mann et al. (2002). More recently,
model slip rates on the Mona Passage fault derived
by Manaker et al. (2008), show SE-NW extension
at 5.7 4.3 mm/a trending ~N65 E, in good
agreement with previously reported extension rate
of 5 3 mm/a between Hispaniola and Puerto Rico
from GPS studies (Jansma et al. 2000; Jansma &
Mattioli, 2005). This slip rate should be much the
same as the one along the Los Muertos trench. This
geometry would also support that the Puerto Rico
block is attached to the Aves ridge and the stable
CA plate, as proposed by Mann et al. (2002), thus
implying that most of the Los Muertos trench and
the Anegada passage features are inactive or below
the current GPS threshold of 2-3 mm/a. These
statements are confirmed by recent results from
Manaker et al. (2008) as well. On one hand, slip
on the Anegada Passage fault is 3 3mm/a directed
~N43W, implying NW-SE extension across this
region. However, relative GPS velocities across the
Anegada Passage are close to zero, suggesting that
the actual slip rate may be closer to the lower end of
the estimated model rate. On the other hand, these
authors estimate that model oblique slip rate on the
Los Muertos Thrust decreases to 1.7 1.7mm/a
east of the Mona Passage (south of the Puerto Rico
block). Then, we could summarize that the most
recent phase of eastward migration of the MCA
may have essentially occurred through rotation of
the northern and southern hinge zones of the Lesser
Antilles subduction, being counterclockwise near
the Virgin Islands on the north and clockwise in
eastern Venezuela and Trinidad, while the CA
plate was shortened in the north-south direction
by the convergence between the Americas set on
in the Oligocene and still ongoing, thus requiring a
small amount of wrenching on both north and south
PBZs, This is in agreement with the amount of late
Cenozoic strike-slip motion in the southern PBZ
estimated by Audemard & Giraldo (1997).
Beata Ridge and Aruba Gap
The Beata Ridge is marked by a triangular-shaped
uplifted area at the place where the Caribbean

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

is narrowest, between the Guajira Peninsula of


Colombia and Hispaniola. It forms the boundary
between the Colombian Basin and the Venezuelan
Basin. Fox & Heezen (1975) suggested that the
structure of the Beata Ridge consists of a steep
fault scarp bounding its western edge and a series
of fault blocks, which step down to the floor of the
Venezuelan Basin. Using seismic profiles, Ladd et
al. (1981) were unable to confirm whether the east
flank of the Beata Ridge is a series of fault blocks
or it consists of parallel volcanic ridges.
North-south seismic profiling in the Aruba Gap
between the southern edge of the Beata Ridge and
the Curacao Trench has revealed the presence of
three fault profiles showing flower structures,
which Hopkins (1973) has interpreted as east-west
trending strike-slip faults. This interpretation is
consistent with strike-slip focal mechanisms for
this general area from Molnar & Sykes (1969).
From an age control established by Hopkins
(1973) for the sedimentary sequence on the Beata
Ridge, there is no record of any sediment younger
than Late Pliocene, thus constraining the time of
deformation.
A variety of geologic and seismic observations have
been cited as evidence for deformation at the Beata
Ridge, possibly driven by slow convergence of
the SA and NA plates across the Caribbean region
(Dixon & Mao, 1997; Mller et al. 1999). DeMets
et al. (2010) list the following: (1) Heubeck & Mann
(1991) suggest that the Caribbean plate consists of
rigid subplates east and west of the Beata Ridge
coinciding with the Venezuelan and Colombian
basins; (2) Consistent with this interpretation, Leroy
& Mauffret (1996) interpret apparently reactivated
reverse faults imaged in marine seismic profiles
that cross the eastern flank of the Beata Ridge as
evidence for contraction across the Beata Ridge and
hence deformation within the CA plate. Mauffret
& Leroy (1999) further interpret compressional
features along the Beata Ridge as evidence for
NE-SW shortening between independently moving
microplates flanking the Beata Ridge and estimate
that the convergence rate across the Beata Ridge
has averaged 9.0 1.5 mm/a for the past 23 Ma.
After Driscoll & Diebold (1998), the timing and
cause for the formation of the Beata Ridge remains
controversial (Holcombe et al. 1990; Mauffret &

Leroy, 1997). For example, recent studies have


proposed that a renewed phase of deformation
occurred in the Miocene and that deformation
is still ongoing in the region of the Beata Ridge
(Mauffret & Leroy, 1997). In the Mauffret & Leroy
(1997) model, the Colombian plate is currently
overthrusting the Venezuelan plate with the
thrust front being located along the eastern flank
of the Beata Ridge. Such a two-plate kinematic
model for the Caribbean was first proposed by
Dewey & Pindell (1985), and requires differential
motion between the Venezuelan and Colombian
basins. This plate kinematic model predicts that
deformation across Beata Ridge accommodated the
differential motion between the eastern and western
Caribbean plates. On the basis of seismic reflection
and dredge data, Fox et al. (1970) proposed that
Beata Ridge was uplifted after the Eocene (i.e.,
post reflector A). An alternative hypothesis, the
single-plate model for the Caribbean, purports that
the majority of the deformation observed across
the Beata Ridge and Venezuelan Basin occurred
early in the history of the Caribbean (i.e., Late
Cretaceous) prior to large sediment accumulation.
Minor fault reactivation along the eastern Beata
Ridge and Venezuelan Basin is inferred to having
been caused by the different styles of deformation
in response to the north-south compres-sional
stress (Holcombe et al. 1990; Diebold et al.
1995). Driscoll & Diebold (1998) seem to bring
conclusive arguments to this debate. They conclude
that the sediment thickness and stratal geometry of
the overlying sedimentary successions across the
Venezuelan Basin and Beata Ridge suggest that
the majority of the observed deformation in this
region occurred soon after the emplacement of the
volcanics, and is best explained by the one-plate
model. Minor fault reactivation in the Neogene
along the eastern flank of the Beata Ridge is
associated with an accommodation zone (i.e., tear
fault) that records a change in the deformation style
from subduction of the CA plate along the Los
Muertos Trough to obduction of the CA plate onto
the Bahamas platform along Hispaniola. We feel
that the resumed tectonic activity across the Beata
Ridge in the Neogene may be linked to left-lateral
motion along the Hess Escarpment. More plausible
is this if the Neogene deformation happens to be
compressional.

55

56

Franck A. Audemard M.

GPS tests for CA plate internal deformation


DeMets et al. (2010)s paper is largely motivated by
the need of testing whether the CA plate undergoes
internal deformation at present day. Simple, but
rigorous numerical experiments with the GPS site
velocities carried out by these authors indicate
that any east-to-west deformation across the Beata
Ridge and Lower Nicaraguan Rise is unlikely to
exceed 2 mm/a, and within the uncertainties is
zero. The kinematic evidence for insignificant eastto-west deformation agrees with results reported
by Driscoll & Diebold (1998), who conclude that
marine seismic data from the Beata Ridge do not
require the occurrence of significant contraction
across this structure since the Miocene. If such
contraction has occurred, as suggested by Mauffret
& Leroy (1999), results from DeMets et al. (2010)
suggest that it has now ceased.
Mechanisms of CA plate internal deformation
Several mechanisms have been invoked to explain
the internal deformations within the CA plate:
(1) Convergence between the Americas since
the Oligocene across the CA plate (Ladd, 1976;
Sykes et al. 1982; Burke et al. 1984; Pindell et
al. 1988), inducing the overriding of neighboring
continental masses of either continental or oceanic
affinity or N-S shortening at CA plate scale; (2)
Plate-scale bending, across the CA plate in the
east-west direction, and subsequent overriding
over the CA plate (Stephan 1982; Stephan et al.
1986), responsible for the generation of several
festoons (NPDB, SCDB, and Hispaniola-Puerto
Rico); (3) Tectonic collision that deforms the
overriding CA plate, such as those of the Cocos
Ridge or the Panam Arc; and (4) Tectonic escape
of continental blocks, such as the NNE-directed
escape of the North Andes Block NAB- (Case et
al. 1971; Dewey, 1972; Pennington, 1981; Stephan,
1982; Audemard, 1993; 1998; Freymueller et al.
1993; Ego et al. 1996); which in turn is to be also
related to tectonic collision of either the Carnegie
Ridge, the southern tip of the Panam Arc or both
jointly. Very little of the internal CA deformation
has been attributed to major features lying within
the plate interior except for the Beata Ridge; above
discussed.

In addition, we propose a different mechanism


of internal deformation, involving the Hess
Escarpment. We believe that two plate boundary
conditions in two different regions, in connection to
both projections of the Hess Escarpment at PBZs,
helped its development as well as its left-lateral
strike-slip motion, probably in the late Mioceneearly Pliocene, as bracketed by counterclockwise
rotations in Puerto Rico measured by Reid et al.
(1991): (1) on the northeastern projection of the
Hess Escarpment, the collision and pinning of
Northern Hispaniola against the buoyant Bahamas
platform; and (2) on the southwestern end, the
high coupling at CO-CA plate interface and
simultaneous collision-incipient indentation of
the buoyant Cocos Ridge. The buoyant Bahamas
platform on one end acts as a buttress while the
buoyant Cocos Ridge on the other tip acts much
as an indenter. These two processes effectively
added to each other to engine the relative faster
escape of the southern unit of the CA plate with
respect to the northern one. In addition, under the
convergence between the Americas, which has
induced N-S shortening across the shortest axis
of the CA plate, the ENE-WSW orientation of the
Hess Escarpment, in comparison to the east-west
orientation of the current NA-CA PBZ, facilitates
even more the migration and escape of the
southern half of the CA plate, due to its wedge
shape enlarging to the east, in the direction of the
escape. The 25 counter-clockwise rotation of
the 500 km long Puerto Rico Block alone helps to
accommodate some 200 km of N-S shortening in
the Eastern Caribbean, east of the Mona Passage.
In turn, this escape could also account for the
decompression of the northern half of the CA plate,
reflected in the opening of N-S trending grabens in
the Nicaragua Rise, and even onshore Guatemala
and Honduras, resulting in the N-S flattening
of the Nicaragua Rise and contiguous continental
Central America region.
An alternate driving mechanism could be advanced
for the geometry and variable along-trend kinematics
of the Los Muertos trench. Although it would keep
contributing to accommodate the N-S shortening
introduced by the convergence of the Americas,
as proposed by many authors, it would incorporate
it in a different manner. Under the assumption
that the Hess Escarpment once slipped left-

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

laterally, while the Puerto Rico Block was rotating


counterclockwise and the Puerto Rico subduction
along this block was progressively evolving
from frontal to oblique subduction with time, the
southern edge of the Puerto Rico Block had to be
overthrusted onto the CA plate to accommodate part
of the compression across the arc. This had to lead
to the generation of the Los Muertos trench. Since
the NE projection of the Hess Escarpment intersects
the western tip of the Los Muertos subduction and
forearc, we propose that the induced Los Muertos
subduction (or overthrusting) can be regarded as
a transpressional fault-tip thrust fault. In fact, it
is a S-to-SSW-verging fold-and-thrust belt. The
arguments favoring this alternate interpretation are:
(1) the geometric angular relationship of the Los
Muertos trench with the Hess Escarpment (~130);
(2) the widening of the Los Muertos sedimentary
wedge towards the west (Granja et al. 2006); (3)
its more pronounced arcuate shape to the west; and
(4) the slip rate decrease from west to east. Most
researchers agree on this. Motion on the low-angle
Muertos thrust fault decreases from the west tip
eastward to longitude 66 W where little or no active
underthrusting is occurring (Masson & Scanlon,
1991). After Manaker et al. (2008), model oblique
slip rates on the Los Muertos Thrust average 5
2 mm/a, decreasing from 7.3 1.0 mm/a west of
the Mona Passage to 1.7 1.7 mm/a east of the
Mona Passage. Mann et al. (2002) estimate 6 to
1 mm/yr of thrust motion across the Los Muertos
fault, also decreasing from west to east. Assuming
that the Hess Escarpment has slipped left-laterally
for some 10 Ma (since Late Miocene and coeval
with the Puerto Rico Block rotation) at an average
slip rate between 2 to 5 mm/a (in the same range
as the slip on the Los Muertos trench or the Mona
Passage extension), a minimum shortening of 20 to
50 km has been accommodated on the Los Muertos
fold-and-thrust belt (sedimentary wedge) if fully
coupled, related to transpression at the northeastern
tip of the Hess Escarpment. In such a case, the Los
Muertos trench would be linked to a pseudo- or
induced subduction.
It must be kept in mind that this relative escape
along the Hess Escarpment may have been arrested
or substantially been slowed down more recently
by the collision and later suturing of the Choc
block against the western border of NW SA, once

attached to the trailing edge of the CA plate that


effectively happened sometime in the Pliocene.
In that sense, we can put forward that the Hess
Escarpment has not been inactive in the past; and
is not currently inactive either, as we shall discuss
next.
Seismicity along the Hess Escarpment
Published maps, such as USGS-NEIC and MIDAS
Consortium (1998) and Tarr et al. (2010), among
others, suggest that the southwestern submarine
expression of the Hess Escarpment, near the
Caribbean shore of Coast Rica, exhibit an associated
shallow seismicity, extending along its bathymetric
expression. This seismicity is very sparse but well
aligned along the trend of the Hess Escarpment. For
instance, the plotted events on the USGS-NEICMIDAS map are above Mw 4.2. To better image
this seismic activity, we show herein a map of the
region with the plotted seismic data shallower than
33 km deep, available from the Global Earthquake
Search of the USGS for the period January 1973
and June 2013 (Figure 14). The fewer (only 4)
but larger dots represent shallow earthquakes of
magnitude M 5.0 within the above mentioned
time window.
What are the implications of this seismic activity? It
is well expressed on all published maps, as well as
on our figure 14, that the seismicity along the Hess
Escarpment is restricted essentially to its southwestern termination. The rest of its extent appears
deprived of any seismicity. It could be argued that
this is biased by the low-density distribution of the
seismographic stations in the Caribbean Sea region
and particularly within the CA plate, but any event
of magnitude M 5.0 should be recorded regardless
of its hypocentral location inside the Caribbean.
Then, the current seismic distribution should
reflect the actual activity of the Hess Escarpment.
If so, why is the crustal seismicity restricted to
its southwestern termination then? As discussed
earlier, this major tectonic feature may have played
a fundamental role in the eastward migration of
the southern uneven half of the CA plate sometime
in the late Miocene-Pliocene, as suggested by the
25 counterclockwise rotation recorded in rocks
of that age in Puerto Rico, but its left-lateral
motion may have ceased, been arrested or slowed

57

58

Franck A. Audemard M.

Figure 14. Seismicity shallower than 33 km deep, available from the Global Earthquake Search of the USGS for the period
January 1973 and June 2013, plotted on GLOBE bathymetry-topography (GLOBE Task Team and others, 1999). Larger cricles
are epicenters of events M 5.0. Note the seismic activity along the southwestern termination of the Hess Escarpment, off the
Caribbean coast of Costa Rica
Figura 14. Mapa de sismicidad de eventos de profundidad inferior a 33 km, extraidos del Global Earthquake Search of the USGS
para el perodo enero 1973 - Junio 2013, representados en la topografa-batimetra de GLOBE (GLOBE Task Team and others,
1999). Los crculos ms grandes son de eventos de M 5.0. Obsrvese la actividad ssmica asociada a la terminacin suoreste del
escarpe de Hess, costa afuera de la costa caribea de Costa Rica

down afterwards, sometime in the Pliocene. As


suggested here, that motion break can be imputed
to the suturing of the Choc block, at the trailing
edge of the CA plate, at that time. The present day
seismicity of the Hess Escarpment close to Central
America has then to be considered as a renewal
or reactivation in seismic activity and a resume
in motion of the Hess Escarpment in the Late (?)
Quaternary, since four focal mechanism solutions
herein presented attest to the left-lateral and normal
components of slip along this major feature (Figure
15). It is very likely that this reactivation is a direct
consequence of the collision-indentation of the
buoyant Cocos Ridge against the Pacific board
of Central America. The main argument in favor
of this hypothesis is the nearness of this seismic
activity to the southwestern termination of this very
major tectonic feature, as well as the alignment of
the Hess Escarpment along the northern edge of
the paralleling Cocos Ridge. To some extent, the
normal left-lateral slip along the Hess Escarpment
is accompanying the overthrusting of the Panam
microplate onto the CA plate, just north of its
northwestern termination along the CCRDB, which

crosses Costa Rica and acts as a thrust left-lateral


lateral ramp for the microplate emplacement. Both
kinematics seem to result from the ongoing Cocos
Ridge indentation.
It is worth mentioning that the major and current
phase of dextral SS in northern Venezuela (CA-SA
PBZ), which would necessarily be the southern
boundary of this southern half of the CA plate, was
set on around 17 Ma (Middle Miocene; Audemard,
1993; 1998; 2009). This strike-slip rupture
propaga-tion was diachronic from west to east,
starting at 17 Ma with Oca-Ancn fault OAF- in
western Venezuela and ending with the generation
of the El Pilar fault at around 10 Ma in the Gulf of
Paria (Audemard, 1993; 1998; 2009). In addition,
the east-west trending OAF, which used to be part
of that southern CA-SA boundary, transferred its
dextral slip onto the Bocon fault sometime around
5-3 Ma (Audemard, 1993; 1998; 2009). Then, it
is worth exploring whether the left-lateral Hess
Escarpment and the right-lateral Oca-Ancn-San
Sebastin-El Pilar fault system together, on the
north and south respectively of the wedge-shaped

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

southern half of the CA plate, did drive and ease the


eastward migration and escape of that southern

CA plate, between the Middle Miocene and Early


Pliocene.

Figure 15. Focal mechanism solutions for earthquakes along the southwestern termination of the Hess Escarpment, off the Caribbean
coast of Costa Rica, extracted from the Global CMT catalog (Dziewonski et al. 1981 and Ekstrm et al. 2012). They attest to normal
and left-lateral components of slip along this tectonic feature
Figura 15. Soluciones de mecanismo focal de sismos asociados a la terminacin suoreste del escarpe de Hess, costa afuera de la
costa caribea de Costa Rica, extraidas del Global CMT catalog (Dziewonski et al. 1981 y Ekstrm et al. 2012). Estos mecanismos
atestiguan las componentes de movimiento normal y lateral izquierda de este accidente tectnico

CONCLUSIONS
With time and gathering of a wealth of geologic,
geophysical, seismological and geodetic data in the
last 3 decades, the knowledge of the Caribbean and
its plate boundaries have become more complex
and better understood. To speak of transform or
subduction boundaries in the case of the Caribbean
region appears very simplistic nowadays. These
borders are real plate boundary zones PBZ-, in
which many tectonic blocks of very diverse and
varied geometry and composition are taken in. In
the southern Caribbean, the lack of a conspicuous
boundary in comparison to other borders triggered
first regionally the study of onshore major tectonic
features and structures, whereas the northern
boundary of the Caribbean plate became a natural
laboratory for GPS studies, due to the limited
extent of the geology in a score of islands. In this
PBZ, GPS networks did not resolve all problems
in terms of kinematics because most of the

networks are on islands, sitting within the active


deformation zone. This problem is also common to
the Caribbean-Atlantic PBZ. The networks require
of stable reference points inside the Caribbean Sea,
such as San Andrs, Providencia and Aves islands.
Although current and recent GPS results point to
very little internal deformation in the Caribbean
plate, probably below 2-3 mm/a, the Hess
Escarpment has a non-neglegible seismic activity
along its southwestern submarine termination, close
to the Costa Rica Caribbean coast. Nevertheless, in
the frame of the Caribbean geodynamic evolution,
this major submarine feature cutting the Caribbean
ocean floor into two uneven halves needs to be
revisited and rethought. It splits the Caribbean
plate into two parts, which exhibit very different
geodynamic characteris-tics within the Caribbean
plate and along its borders.
All Caribbean plate borders contain more than one
tectonic block, and strain partitioning at different

59

60

Franck A. Audemard M.

scale is common to all Caribbean plate boundaries.


Trench-parallel slivers have been defined all over.
Between the active volcanic arc and the oceanic
trench, slivers have been proposed at the northern
Lesser Antilles forearc, as well as along the
Middle America trench north of Nicoya Peninsula
in Central America. But they also happen along
northern Venezuela (Bonaire block and Interior
ranges of Venezuela) and north of the Septentrional
fault in northern Hispaniola.
On top of that, indentation and indentation-extrusion
processes are more frequent than ever thought
in the PBZs. The best imaged of these processes
is the collision and suturing of the Panam block
against NW South America, but the high buoyancy
of different ocean-floor anomalies (Cocos and
Carnegie ridges, and Bahamas carbonate platform)
on oceanic plates have been held responsible
for strain partitioning within the PBZ. In some
occasions, convergence obliquity has been called
upon as the driving force, such as at the northern
Lesser Antilles forearc.
The refinement of tectonic block definition, as well
as the identification of tectonic blocks, has been
boosted by GPS Geodesy, and it still has plenty
to provide in that sense. In some cases, smallscale spatial consistency of GPS vectors has led
to the proposal of small tectonic blocks, such as
that of the Nicaragua Rise, which incorporates the
southern part of Jamaica.
Finally, blocks overriding the Caribbean plate have
already been defined at three places. They seem
not to be true subductions, but tectonic blocks, of
composition different to oceanic crust, that thrust
onto the Caribbean oceanic lithosphere, not driven
by mantle convection. Coincidentally, all three
festoons, interpreted as induced subductions, have
been identified south of Hess Escarpment.
ACKNOWLEDGEMENTS
To pretend to thank all who somehow helped
to my personal brainstorming on the Caribbean
geodyna-mics over the last 30 years would simply
be irrational. The list of references herein included
reflects to some extent who may have contributed to
the geo-fantasy that my brain may have painted

and described herein after that life-long time


span of live and prolonged discussions, endless
someone elses paper reading, numerous congress
and workshop attendances, shared beers, wines and
meals around a friendly table, and so forth. Some of
you, however, deserve my respect and admiration.
There are corner-people in the development of
the Caribbean geodynamics. To them, my most
humble respects!!!! My graduate students at School
of Geology, Mines and Geophysics of Universidad
Central de Venezuela (-UCV-, Caracas, Venezuela)
deserve these lines, because always challenging
me to go beyond and beyond. Particularly in the
preparation of this contribution, we are very
thankful to Marina Pea for redrawing Figure
1; also to Miguel Palma, from the Seismological
Department of the Venezuelan Foundation
for Seismological Research FUNVISIS-, for
generating Figures 14 and 15. These figures were
built with GMT (The Generic Mapping Tools),
created in 1988 by Paul Wessel and Walter H. F.
Smith, and we wish to thank them for the free use
of it. I particularly thank Maximiliano Bezada for
handing over and preparing Figure 11 especially
for this contribution, following a few requirements
of mine. My thanks also go to Carlos Reinoza for
preparing Figure 13. My heartiest and most sincere
words of thankfulness to my co-Invited Editor
of this GEODINOS compilation book, Michael
Schmitz from the Geophysics Dpt. of FUNVISIS,
for constantly nagging me and progressively
raising pressure on me, without which I would have
never assembled these ideas. This is a contribution
to projects FONACIT-ECOS Nord 200300090
and 200900818 and FONACIT 2001002492
(SEGMEN-TACIN SISMOGENTICA FALLA
DE BOCON), 2002000478 (GEODINOS),
2012002202
(GIAME)
and
2013000361
(TSUNAMI). I wish to dedicate this piece of work
to Jean-Francois Stephan ( 21 December 2013)
who largely contributed to the Venezuelan tectonics
and Caribbean Geodynamics over the years, as well
as to Robert (Bob) Trenkamp ( 29 June 2013) who
worked on Geodesy of the NW corner of SA for
many years.
REFERENCES
Adamek, S., Frohlich, C., Pennington, W. D. (1988).
Seismicity of the CaribbeanNazca boundary;

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

Constraints on microplate tectonics of the Panama


region, J. Geophys. Res. 93: 2053-2075.
Algar, S. & Pindell, J. L. (1993). Structure and
deformation history of the northern range of Trinidad
and adjacent areas. Tectonics, 12: 814-829.
Alvarado, D., Demets, C., Tikoff, B., Hernndez, D.,
Wawrzyniec, T.F., Pullinger, C., Mattioli, G.,
Turner, H.L., Rodrguez, M., Correa-Mora, F.
(2011). Forearc motion and deformation between El
Salvador and Nicaragua: GPS, seismic, structural,
and paleomagnetic observations. Lithosphere, 3(1):
3-21.
Allen, C.R. (1981). The modern San Andreas Fault.
In: The Geotectonic Development of California,
Prentice-Hall, Englewood Cliffs, N.J., 511-534.
Audemard, F. A. (1993). Notectonique, Sismotectonique
et Ala Sismique du Nord-ouest du Vnzula
(Systme de failles dOcaAncn). PhD thesis,
Universit Montpellier II, France, 369 pp + appendix.
Audemard, F. A. (1997). Holocene and Historical
Earthquakes on the Bocon Fault System, Southern
Venezuelan Andes: Trench Confirmation. Journal of
Geodynamics, 24: 155-167.
Audemard, F. A. (1998). Evolution Godynamique de
la Faade Nord Sud-amricaine: Nouveaux apports
de lHistoire Gologique du Bassin de Falcn,
Vnzula. Proceedings XIV Caribbean Geological
Conference, Trinidad, 1995, 2: 327-340.
Audemard, F. A. (1999). Morpho-structural expression
of active thrust fault systems in humid tropical
foothills of Colombia and Venezuela. Zeitschrift fr
Geomorphologie, 118: 1-18.
Audemard, F. A. (2000). Major Active Faults of
Venezuela. Proceedings of 31st International
Geological Congress, Rio de Janeiro, Brazil (extended
abstract; on CD-ROM).
Audemard, F. A. (2002). Ruptura de los grandes
sismos histricos venezolanos de los siglos XIX
y XX, revelados por la sismicidad instrumental
contempornea. XI Congreso Venezolano de
Geofsica, Caracas, Venezuela, Nov. 17-20, 2002
(8pp; Extended Abstract in CD).
Audemard, F. A. (2003). Geomorphic and geologic
evidence of ongoing uplift and deformation
in the Mrida Andes, Venezuela. Quaternary
International, 101-102C: 43-65. doi: 10.1016/S10406182(02)00128-3
Audemard, F. A. (2009). Key issues on the post-Mesozoic
southern Caribbean plate boundary. In: James, K.H.,
Lorente, M.A. and Pindell, J., (eds) Origin and
Evolution of the Caribbean Plate, Geological Society,
London, Special Publications, 328: 567-584. doi:
10.1144/SP328.23
Audemard, F. A. & Giraldo, C. (1997). Desplaza-mientos

dextrales a lo largo de la frontera meridional de la


placa Caribe, Venezuela septentrional. 8 Congreso
Geolgico Venezolano, Porlamar, Venezuela 1: 101108.
Audemard, F. A., Romero, G., Rendn, H., Cano, V.
(2005). Quaternary fault kinematics and stress tensors
along the southern Caribbean from fault-slip data and
focal mechanism solutions. Earth Science Reviews,
69: 181-233.
Audemard, F. A., Machette, M., Cox, J., Dart, R.,
Haller, K. (2000). Map and Database of Quaternary
Faults and Folds in Venezuela and its Offshore
Regions. USGS Open-File Reports, 00-0018
(accessible from USGS webpage; open file reports
ofr-00-0018).
Audemard, F. A., Ollarves, R., Betchtold, M., Daz,
G., Beck, C., Carrillo, E., Pantosti, D., Diederix,
H. (2008). Trench investigation on the main strand of
the Bocon fault in its central section, at Mesa del
Caballo, Mrida Andes, Venezuela. Tectonophysics,
459: 38-53. doi: 10.1016/j.tecto.2007.08.20.
Audemard, F. A., Pantosti, D., Machette, M., Costa,
C., Okumura, K., Cowan, H., Diederix, H., Ferrer,
C., Sawop Participants. (1999). Trench investigation
along the Merida sectin of the Bocon fault (central
Venezuelan Andes), Venezuela. In: Pavlides, S.,
Pantosti, D. and Peizhen, Z. (Eds.) Earthquakes,
Paleoseismology and Active Tectonics. Selected
papers to 29th General Assembly of the Association
of Seismology and Physics of the Earths Interior
(IASPEI), Thessaloniki, Greece, August 1997.
Tectonophysics, 308: 1-21. doi: 10.1016/S00401951(99)00085-2
Audemard, F. E. (1991). Tectonics of Western Venezuela.
PhD thesis, Rice University, Houston, TX.
Audemard, F. E. & Audemard, F. A. (2002). Structure
of the Mrida Andes, Venezuela: relations with the
South America-Caribbean geodynamic interaction.
Tectonophysics, 345: 299-327.
Babb, S. (1997). Tectonics and sedimentation of the Gulf
of Paria and Northern Basin, Trinidad. PhD thesis,
The University of Texas at Austin, TX.
Baumgartner, P. O., Flores, K., Bandini, A. N., Girault,
F., Cruz, D. (2008). Upper Triassic to Cretaceous
radiolaria from Nicaragua and northern Costa Rica
-The Mesquito Composite Oceanic Terrane. Ofioliti,
33: 1-19.
Beck, C. (1985). Caribbean colliding, Andean drifting
and the Mesozoic-Cenozoic geodynamic evolution of
the Caribbean. Proceedings of VI Congreso Geolgico
Venezolano, Caracas, 10: 6575-6614.
Beltrn, C. (1993). Mapa Neotectnico de Venezuela.
Scale 1 : 2,000,000. Funvisis.
Beltrn, C. 1994. Trazas activas y sntesis neotectnica

61

62

Franck A. Audemard M.

de Venezuela a escala 1 : 2.000.000. Proceedings of


VII Congreso Venezolano de Geofsica, Caracas,
541-547.
Beltrn, C. & Giraldo, C. (1989). Aspectos
neotectnicos de la regin nororiental de Venezuela.
Proceedings of VII Congreso Geolgico Venezolano,
Barquisimeto, 3: 1000-1021.
Benford, B., Demets, C., Calais, E. (2012a). GPS
estimates of microplate motions, northern Caribbean:
evidence for a Hispaniola microplate and implications
for earthquake hazard.
Benford, B., Demets, C., Tikoff, B., Williams, P.,
Brown, L., Wiggins-Grandison, M. (2012b). Seismic
hazard along the southern boundary of the Gnave
microplate: block modelling of GPS velocities from
Jamaica and nearby islands, northern Caribbean.
Geophys. J. Int. (2012) 190: 59-74.
Bentez, S. (1986). Sntesis geolgica del graben de
Jambeli. IV Congreso Ecuatoriano de Geologa,
Minera y Petrleo, 1, 137-160.
Bevis, M. & Isacks, B.L. (1984). Hypocentral trend
surface analysis: probing the geometry of Benioff
zones. J. Geophys. Res., 89: 6153-6170.
Bezada, M. (2010). Crustal and upper mantle
investigations of the Caribbean-South American plate
boundary. PhD Thesis, Rice University, Houston,
USA. 143 pp.
Bezada, M.J., Levander, A., Schmandt, B. (2010).
Subduction in the southern Caribbean: Images from
finite-frequency P wave tomography. J. Geophys.
Res. 115, B12333. doi: 10.1029/2010JB007682.
Biju-Duval, B, Mascle, A., Montadert, L., Wanneson,
J. (1978). SEISMIC INVESTIGATIONS IN THE
COLOMBIA, VENEZUELA and Grenada basins,
and in the Barbados ridge for future drilling. Geologie
Mijnbouw, 57: 105-116.
Bizon, G., Bizon, J. J., Calmus, T., Muller, C., Van
Den Berghe, B. (1985). Stratigraphie du Tertiare
du sud dHispaniola (Grandes Antilles): Influence
de la tctonique dcrochante sur la palogographie
et lhistoire sedimentaire. In: Proceedings
Godynamique des Carabes Symposium, Paris,
France, February, 1985. Editions Technip, Paris, 371380.
Bourdon, E., Eissen, J., Gutscher, M., Monzier, M.,
Hall, M., Cotton, J. (2003). Magmatic response
to early aseismic ridge subduction: the Ecuadorian
margin case (South America). Earth and Planetary
Science Letters, 205: 123-138.
Bouysse, P. (1979). Caractres morphostructuraux et
volution godynamique de larc insulaire des Petites
Antilles (campagne ARCANTE 1). Bull. BRGM, IV
(I): 3-14.
Bouysse, P. (1988). Opening of the Grenada back-arc

Basin and evolution of the Caribbean plate during


the Mesozoic and Early Paleogene. Tectonophysics
149:121-143.
Bouysse, P. & Westercamp, D. (1990). Subduction of
Atlantic aseismic ridges and Late Cenozoic evolution
of the Lesser Antilles island arc. Tectonophysics, 175:
349-380.
Bouysse P., Baubron, J.C., Richard, M., Maury, R.C.,
Andreieff, P. (1985). Evolution de la terminaison
nord de larc interne des Petites Antilles. Bull. Soc.
Gol. Fr., (8)1: 181-188.
Bowin, C.O. (1975). The geology of Hispaniola. In:
Nairn, A.E.M and Stehli F. G. (eds.) The Ocean Basins
and Margins: The Gulf of Mexico and the Caribbean,
Plenum New York, 501-552.
Bowin, C. O. (1976). Caribbean gravity field and plate
tectonics, Geol. Soc. Am. Spec. Pap., 169: 179.
Brown, K. M. & Westbrook, G. K. (1987). The tectonic
fabric of Barbados Ridge accretionary complex. Mar.
Pet. Geol., 4: 71-81.
Burbach, G.V., Frohlich, C., Pennington, W.D.,
Matumoto, T. (1984). Seismicity and tectonics of the
subducted Cocos Plate. J. Geophys. Res. 89: 77197735.
Burkart, B. (1978). Offset across the Polochic fault of
Guatemala and Chiapas, Mexico. Geology, 6: 328332.
Burkart, B. (1983). Neogene North AmericanCaribbean Plate boundary across northern Central
America: Along the Polochic fault. Tectonophysics,
99: 251 270.
Burke, K., Fox, P.J., Sengr, A.M.C. (1978). Buoyant
ocean floor and the evolution of the Caribbean. J.
Geophys. Res., 83 (B8): 3949-3954.
Burke, K., Grippi, J., Sengr, A.M.C. (1980). Neogene
structures in Jamaica and the tectonic style of the
Northern Caribbean Plate Boundary Zone. J. Geol.,
88: 375-386.
Burke, K., Cooper, C., Dewey, J.F., Mann, P., Pindell,
J.L. (1984). Caribbean tectonics and relative plate
motions. In: Bonini, W.E., Hargraves, R.B. and
Shagam, R. (eds) The CaribbeanSouth America
plate boundary and regional tectonics, Geological
Society of America Memoir, 162: 31-64.
Byrne, D., Suarez, G., Mccann, W. (1985). Muertos
Trough subduction-microplate tectonics in the
northern Caribbean? Nature, 317: 420-421.
Calais, E., Mazabraud, Y., Mercier De Lepinay,
B., Mann, P. (2002). Strain partitioning and fault
slip rates in the northeastern Caribbean from GPS
measurements, Geophys. Res. Lett., 29: 1856-1859.
Calais, E., Mercier De Lepinay, B., Saint-Marc, P.,
Butterlin, J., Schaaf, A. (1992). La limite de plaques
dcrochante nord carabe en Hispaniola: volution

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

palogographique et structurale cnozoque. Bulletin


Societ Gologique de France, 163: 309-324.
Calais, E., Freed, A., Mattioli, G., Amelung, F., Jnsson,
S. Jansma, P., Hong, S-H. Dixon, T., Prepetit, C.,
Momplaisir, R. (2010). Transpressional rupture of
an unmapped fault during the 2010 Haiti earthquake.
Nat. Geosci., 3: 794-799. doi: 10.1038/NGEO992.
Calmus, T. (1983). Dcrochement senestre sud-haitien:
Analyses et consequences palogographiques dans
la region de Camp Perrin (Massif de la Macaya,
Presquile du Sud dHaiti). Extrait Ann. Soc. Gol.
Nord (Lille, France), Sance du 10 Juin, 1983: 309316.
Camacho, E., Hutton, W., Pacheco, J. F. (2010). A
new look at evidence for a Wadati-Benioff zone and
active convergence at the north Panama deformed
belt, Bull. Seismol. Soc. Am., 100(1): 343-348, doi:
10.1785/0120090204.
Cantalamessa, G. & Di Celma, C. (2004). Origin and
chronology of Pleistocene marine terraces of Isla
de la Plata and of flat, gently dipping surfaces of
the southern coast of Cabo San Lorenzo (Manab,
Ecuador). Journal of South American Earth Sciences
16, 633-648.
Carr, M. J. (1984). Symmetrical and segmented
variation of physical and geochemical characteristics
of the Central American volcanic front. Geol. Soc.
Am. Bull. 88: 151-156.
Carr, M. J. & Stoiber, R.E. (1990). Volcanism. In
Dengo, G., Case, J.E. (eds.): The Caribbean Region,
vol. H. Geological Society of America, Boulder, 375391.
Case, J. E. (1975). Geophysical studies of the Caribbean
Sea. In: Nairn, A.E.M. and Stehli, F.G. (Eds.): The
Ocean Basins and Margins, Vol. 3. Plenum Press,
New York, N.Y., 663-689.
Case, J. E. & Holcombe, T. L. (1980). Geologic-tectonic
map of the Caribbean region: U.S. Geol. Survey
Miscell. Investig. Series, Map I-100, scale 1:2 500
000.
Case, J. E., Durn, L., Lpez, A., Moore, W. (1971).
Tectonic investigations in western Colombia and
eastern Panam. Bulletin Geological Society of
America, 82: 2685-2712.
Caytrough (1979). Geological and geophysical
investigation of the Mid-Cayman rise spreading
center: Initial results and observations. In: Talwani,
M., Harrison, C. G. and Hayes D. E. (eds) Deep
Drilling Results in the Atlantic Ocean: Ocean Crust.
Maurice Ewing Ser., 2:66-93.
Choy, J., Morandi, M. T., Palme De Osechas, C. (1998).
Determinacin de patrones de esfuerzos tectnicos
para el Oriente de VenezuelaSureste del Caribe a
partir de mecanismos focales. Proceedings of IX

Congreso Venezolano de Geofsica, Caracas (CDROM; paper no. 17)


Colletta, B., Roure, F., De Toni, B., Loureiro, D.,
Passalacqua, H., Gou, Y. (1996). Tectonic inheritance
and structural styles in the Merida Andes (western
Venezuela). 3rd International Symposium on Andean
Geodynamics, Saint-Malo, France, 323-326.
Colletta, B., Roure, F., De Toni, B., Loureiro,
D., Passalacqua, H., Gou, Y. (1997). Tectonic
inheritance, crustal architecture, and contrasting
structural styles in the Venezuelan Andes. Tectonics,
16: 777-794.
Colmenares, L. & Zoback, M. D. (2003). Stress field and
seismotectonics of northern South America. Geology,
31: 721-724.
Cooper, M. A., Addison, F. T., Alvrez, R., Coral, M.,
Graham, R., Hayward, A., Howe, S., Martnez, J.,
Naar, J., Peas, R., Pulham, A. Taboada, A. (1995).
Basin development and tectonic history of the Llanos
Basin, Eastern Cordillera, and Middle Magdalena
Valley, Colombia. Bulletin of the American
Association of Petroleum Geologists, 79: 1421-1443.
Corredor, F. (2003). Seismic strain rates and distributed
continental deformation in the northern Andes and
three-dimensional seismotectonics of the northwestern
South America. Tectonophysics 372: 147-166.
Corts, M. & Angelier, J. (2005). Current states of
stress in the northern Andes as indicated by focal
mechanisms of earthquakes. Tectonophysics 403: 2958.
Cowan, H. (1999). Design earthquakes for the southeast
area of the canal basin, Panam, Consulting report to
the Authoridad Canal de Panam. 35 pp.
Cowan, H., Dart, R. L., Machette, M. N. (1998).
Map and Database of Quaternary faults and folds
of Panama and its offshore regions: A project of the
International Lithosphere Program Task Group II-2,
Major Active Faults of the World, U.S. Geological
Survey Open File Report 98-779. 41 pp.
Deng, J. & Sykes, L.R. (1995). Determination of Euler
pole for contemporary relative motion of Caribbean
and North American plates using slip vectors of
interpolate earthquakes. Tectonics, 14: 39-53. doi:
10.1029/94TC 02547.
Dengo, G. (1968). Estructura geolgica, historia tectnica
y morfologa de Amrica Central. Centro Regional
de Ayuda Tcnica, Agencia para el Desarrollo
internacional (AID), Mxico. 50 pp.
Dengo, G. & Bohnenberg, O. (1969). Structural
development of Central America. Mem. Am. Assoc.
Pet. Geol. 11: 203-220.
Deshon, H., Schwartz, S. Y., Newman, A. V., Gonzlez,
V., Protti, M., Dorman, L. M., Dixon, T. H.,
Sampsom, D. E., Flueh, E. R. (2006). Seismogenic

63

64

Franck A. Audemard M.

zone structure beneath the Nicoya Peninsula, Costa


Rica, from three-dimensional local earthquake P- and
S-wave tomography, Geophys. J. Int., 164, 109124,
doi:10.1111/j.1365-246X.2005.02809.x.
De Toni, B. & Kellogg, J. (1993). Seismic evidence
for blind thrusting of the northwestern flank of the
Venezuelan Andes. Tectonics, 12, 1393-1409.
DeMets, C. (1993). Earthquake slip vectors and estimates
of present-day plate motions: Journal of Geophysical
Research, 98: 6703-6714.
DeMets, C. (2001). A new estimate for present-day
CocosCaribbean plate motion: implications for slip
on the Central American volcanic arc. Geophysical
Research Letters, 28: 4043-4046.
DeMets, C. & Wiggins-Grandison, M. (2007).
Deformation of Jamaica and motion of the Gonave
microplate from GPS and seismic data. Geophys. J.
Int., 168: 362-378.
DeMets, C., Gordon, R. G., Argus. D. F. (2010).
Geologically current plate motions. Geophys. J. Int.
181: 1-80; doi: 10.1111/j.1365-246X.2009.04491.x.
DeMets, C., Gordon, R. G., Argus, D. F., Stein, S.
(1990). Current plate motions, Geophys. J. Int., 101:
425-478.
DeMets, C., Gordon, R. G., Argus, D. F., Stein, S.
(1994). Effect of recent revisions to the geomagnetic
reversal timescale on estimates of current plate
motions: Geophysical Research Letters, 21: 21912194. doi: 10.1029/94GL02118.
DeMets, C., Jansma, P. E., Mattioli, G. S., Dixon,
T. H., Farina, F., Bilham, R., Calais, E., Mann, P.
(2000). GPS geodetic constraints on Caribbean-North
America plate motion. Geophysical Research Letters,
27: 437-440, doi: 10.1029/1999GL005436.
Dewey, J. (1972). Seismicity and tectonics of western
Venezuela. Bulletin Seismological Society of
America, 62: 1711-1751.
Dewey, J. F. & Pindell, J. L. (1985). Neogene block
tectonics of eastern Turkey and northern South
America: Continental applications of the finite
difference method: Tectonics, 4: 71-83.
Dewey, J. F. & Sengr, A. M. C. (1979). Aegean
and surrounding regions: Complex multiplate and
continuum tectonics in aconvergent zone, Geol. Soc.
Am. Bull., 90: 84-92.
Diebold, J. B., Driscoll, N.W., Abrams, L., Donnelly,
T.W., Laine, E. P., Leroy, S. (1995). Tectonic model
for the origin and evolution of the rough/smooth
basement in the Venezuelan Basin: Eos (Transactions,
American Geophysical Union) 76: F615.
Dixon, T. H. & Mao, A. (1997). A GPS estimate of
relative motion between North and South America,
Geophys. Res. Lett., 24: 535-538.
Dixon, T., Gonzlez, G., Katsigris, E., Lichten,

S. (1991). First epoch geodetic measurements


with the Global Positioning System across the
northern Caribbean plate boundary zone: Journal of
Geophysical Research, 96: 2397-2415.
Dixon, T.H., Farina, F., Jansma, P., Mannand, P., Calais,
E. (1998). Relative motion between the Caribbean
and North American plates and related boundary zone
deformation from a decade of GPS observations. J.
Geophys. Res. 103: 15157- 15182.
Dolan, J. F. & Bowman, D. D. (2004). Tectonic and
seismologic setting of the 22 September 2003, Puerto
Plata, Dominican Republic Earthquake: Implications
for earthquake hazard in Northern Hispaniola. Seism.
Res. Lett., 75 (5):582-592.
Dolan, J. F., Mullins, H. T., Wald, D. J. (1998). Active
tectonics of the north-central Caribbean: Oblique
collision, strain partitioning, and opposing subducted
slabs, in Dolan, J. F., and Mann, P., eds., Active
Strike-Slip and Collisional Tectonics of the Northern
Caribbean Plate Boundary Zone: Boulder, Colorado,
Geological Society of America Special Paper 326.
1-61.
Donnelly, T.W. (1973). Late Cretaceous basalts from the
Caribbean, a possible flood basalt province of vast
size. EOS Trans Am. Geophys. Union 54: 1004.
Driscoll, N. W. & Diebold, J. B. (1998). Deformation
of the Caribbean region: One plate or two? Geology,
26 (11): 1043-1046.
Duque-Caro, H. (1978). Major Structural Elements
of Northern Colombia. American Association of
Petroleum Geologists Memoirs, 29: 329-351.
Duque-Caro, H. (1990). The Choc block in the
northwestern corner of South America; structural,
tectonostratigraphic
and
paleogeographic
implications. Journal of South American Earth
Sciences, 3: 1-14.
Dziewonski, A. M., Chou, T-A., Woodhouse, J. H. (1981).
Determination of earthquake source parameters from
waveform data for studies of global and regional
seismicity, J. Geophys. Res., 86, 2825-86,2852.
doi:10.1029/JB086iB04p02825.
Edgar, N.T., Ewing, J.I., Hennion, J. (1971). Seismic
refraction and reflection in Caribbean Sea. Am.
Assoc. Pet. Geol. Bull., 55 (6): 883-870.
Egbue, O. & Kellogg, J. (2010). Pleistocene to Present
North Andean escape. Tectonophysics, 489(1-4):
248-257.
Ego, F., Sbrier, M., Ypes, H. (1995). Is the Cauca-Patia
and Romeral left or right-lateral? J. Geophys. Res.,
22(1): 33-36.
Ego, F., Sebrier, M., Lavenu, A., Yepes, H., Egues, A.
(1996). Quaternary state of stress in the northern
Andes and the restraining bend model for the
Ecuadorian Andes. Tectonophysics, 259: 101-116.

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

Ekstrm, G., Nettles, M., Dziewonski, A. M. (2012). The


global CMT project 2004-2010: Centroid-moment
tensors for 13,017 earthquakes, Phys. Earth Planet.
Inter., 200-201: 1-9. doi:10.1016/j.pepi.2012.04.002.
Erdlac, R.J. Jr. & Anderson, T.H. (1982). The ChixoyPolochic Fault and its associated fractures in western
Guatemala, Geol. Soc. Am. Bull., 93: 57-67.
Escalona, A. & Mann, P. (2011). Tectonics, basin
subsidence mechanisms, and paleogeography of the
Caribbean-South American plate boundary zone.
Marine and Petroleum Geology 28: 8-39.
Escalona, A., Mann, P., Jaimes, M. (2011). Miocene to
recent Cariaco basin, offshore Venezuela: Structure,
tectonosequences, and basin-forming mechanisms.
Marine and Petroleum Geology 28 (1): 177-199.
Fan, G.-W., Beck, S. L., Wallace, T. C. (1991). The
seismic source parameters of the 1991 Costa Rica
aftershock sequence: Evidence for a transcurrent plate
boundary. J. Geophys. Res., 18: 1385-1388.
Farris, D. W. et al. (2011). Fracturing of the Panamanian
Isthmus during initial collision with South America,
Geology, 39: 1007-1010. doi: 10.1130/G32237.1.
Feuillet, N., Manighetti, I., Tapponnier, P., Jacques,
E. (2002). Arc parallel extension and localization of
volcanic complexes in Guadeloupe, Lesser Antilles,
J. Geophys. Res., 107. doi: 10.1029/ 2001JB000308.
Feuillet, N., Leclerc, F., Tapponnier, P., Beauducel,
F., Boudon, G., Le Friant, A., Deplus, C., Lebrun,
J-F., Nercessian, A., Saurel, J-M., Clement, V.
(2010). Active faulting induced by slip partitioning
in Montserrat and link with volcanic activity:
New insights from the 2009 GWADASEIS marine
cruise data, Geophys. Res. Lett., 37, L00E15.
doi:10.1029/2010GL042556.
Fitch, T.J. (1972). Plate convergence, transcurrent faults,
and internal deformation adjacent to southeast Asia
and the western Pacific. J. Geophys. Res., 77: 44324460.
Fox, P. J. & Heezen, B. C. (1975). Geology of the
Caribbean crust, In: Nairn, A. E. M. and F. G. Stehli
(eds) The Ocean Basins and Margins: The Gulf of
Mexico and the Caribbean, Plenum, New York, 421466.
Fox, P. J., Ruddiman, W. F., Ryan, W. B. F., Heezen, B. C.
(1970). The geology of the Caribbean crust, I: Beata
Ridge: Tectonophysics, 10: 495-513.
Freymueller, J. T., Kellogg, J. N., Vega, V. (1993).
Plate motions in the north Andean region. Journal of
Geophysical Research, 98: 21,853-21,863.
Garrison, J.M. & Davidson, J.P. (2003). Dubious case
for slab melting in the Northern volcanic zone of the
Andes. Geology 31: 565-568.
GEORED (2009). http://geored.ingeominas.gov.co/
Gingerich, P. D. (1985). South American mammals in

the Palaeocene of North America. In: Stehli, F. G.


and Webb, S. D. (eds) The Great American Biotic
Interchange. Plenum Press, New York, 123-137.
Globe Task Team And Others (Hastings, D. A., Dunbar,
P. K., Elphingstone, G. M., Bootz, M., Murakami,
H., Maruyama, H., Masaharu, H., Holland, P.,
Payne, J., Bryant, N.A., Logan, T. L., Muller, J.P., Schreier, G. And Macdonald, J. S. (Eds). (1999).
The Global Land One-kilometer Base Elevation
(GLOBE) Digital Elevation Model, Version 1.0.
National Oceanic and Atmospheric Administration,
National Geophysical Data Center, 325 Broadway,
Boulder, Colorado 80305-3328, U.S.A. Digital data
base on the World Wide Web (URL: http://www.ngdc.
noaa.gov/mgg /topo/globe.html) and CD-ROMs.
Goes, S. D. B., Velasco, A.A, Schwartz, S.Y., Lay,
T. (1993). The April 22, 1991, Valle de Estrella,
Costa Rica (Mw = 7.7) earthquake and its tectonic
implications; a broadband seismic study. J. Geophys.
Res., 98: 8127-8142.
Gordon, M.B. & Muehlberger, W.R. (1994). Rotation of
the Chorts block causes dextral slip on the Guayap
Fault. Tectonics 13: 858-872.
Granja, J. L., Carb, A., Muoz-Martn, A.,
Ballesteros, M. G. (2006). Cinturn Deformado de
Los Muertos (Noreste de la Placa Caribe): Anlisis
Morfotectnico y Procesos Activos. E-Complutense
(http://eprints.ucm.es/5880/).
Grindlay, N.R., Mann, P., Dolan, J. (1997). Researchers
investigate submarine faults north of Puerto Rico, Eos
(Transactions), 78: 404 (Abstract).
Grindlay, N. R., Mann, P., Dolan, J. F., Van Gestel, J. P.
(2005). Neotectonics and subsidence of the northern
Puerto Rico-Virgin Islands margin in response to the
oblique subduction of high-standing ridges, in Mann,
P. (ed.): Active Tectonics and Seismic Hazards of
Puerto Rico, the Virgin Islands, and Offshore Areas,
Geol. Soc. Am. Spec. Paper 385: 31-60.
Gutscher, M. A., Malavieille, J., Lallemand, S.,
Collot, J.Y. (1999). Tectonic segmentation of the
North Andean margin; impact of the Carnegie Ridge
collision. Earth and Planetary Science Letters, 168:
255-270.
Guzmn-Speziale, M. (1995). Relative motion of the
Central America forearc sliver. (abstract), EOS
Transact. Am. Geophys.Union 76, F547 (suppl.).
Guzman-Speziale, M. (2001). Active seismic deformation
in the grabens of northern Central America and
its relationship to the relative motion of the North
America-Caribbean plate boundary. Tectonophysics
337, 3951.
Guzman-Speziale, M. (2010). Beyond the Motagua and
Polochic faults: Active strike-slip faulting along the
Western North AmericaCaribbean plate boundary

65

66

Franck A. Audemard M.

zone. Tectonophysics, 496: 17-27.


Guzmn-Speziale, M. & Gmez, J.M. (2002). Comment
on A new estimate for present-day Cocos-Caribbean
plate motion: implications for slip along the Central
American volcanic arc by Charles DeMets. Geophys.
Res. Lett. 29, 6.1-6.2, doi:10.1029/2002GL015011.
Guzmn-Speziale, M. & Meneses-Rocha, J.J. (2000).
The North America-Caribbean plate boundary west
of the MotaguaPolochic fault system: a fault jog
in southeastern Mexico. Journal of South American
Earth Sciences, 13: 459-468.
Guzmn-Speziale, M., Pennington, W, D., Matumoto,
J. (1989). The triple junction of the Cocos, North
America and Caribbean plates: Seimicity and
Tectonics. Tectonics 8: 981-997.
Guzmn-Speziale, M., Valds-Gonzlez, C., Molina,
E., Martn Gmez, J. (2005). Seismic activity along
the Central America volcanic arc: Is it related to
subduction of the Cocos plate? Tectonophysics 400:
241-254.
Harlow, D.H. & White, R.A. (1985). Shallow
earthquakes along the volcanic chain in Central
America: evidence for oblique subduction (abstract).
Earthq. Notes 55, 28.
Hauff, F., Hoernle, K.A., Van Den Bogaard, P.,
Alvarado, G.E., Garbe-Schnberg, D. (2000). Age
and geochemistry of basaltic complexes in western
Costa Rica: Contributions to the geotectonic evolution
of Central America. Geochemistry Geophysics
Geosystems G3-, 1(5). doi 10.1029/1999GC 000020.
Hess, H. (1962). History of Ocean Basins. In: Engel, A.
E. J., James, H. L. and Leonard B. F. (eds.): Petrologic
studies: a volume in honor of A. F. Buddington.
Geological Society of America, New York?, 599-620.
Hess, H. & Maxwell, J. (1953). Caribbean Research
Project. Bulletin Geological Society of America, 64:
1-6.
Heubeck, C. & Mann, P. (1991). Structural geology
and Cenozoic tectonic history of the southeastern
termination of the Cordillera Central, Dominican
Republic. In: Mann, P., Draper, G. and Lewis,
J.F. (eds), Geologic and Tectonic Development of
the North America-Caribbean Plate Boundary in
Hispaniola. Geol. Soc. Am., Spec. Pap., 262: 315-336.
Heubeck, C., Mann, P., Dolan, J., Monechi, S. (1991).
Diachronous uplift and recycling of sedimentary
basins during Cenozoic tectonic transpression,
northeastern Caribbean plate margin. Sediment.
Geol., 70: 1-32.
Hey, R. (1977). Tectonic evolution of the CocosNazca
spreading center. Geol. Soc. Am. Bull., 88: 14041420.
Holcombe, T. L., Ladd, J. W., Westbrook, G., Edgar,
N. T., Bowland, C. L. (1990). Caribbean marine

geology; Ridges and basins of the plate interior.


In: Dengo, G. and Case, J.E. (Eds) The Caribbean
Region, The Geology of North America, vol. H,
Geological Society of America, 231-260.
Hopkins, H. R. (1973). Geology of the Aruba Gap abyssal
plain near DSDP site 153, in Initial Reports of the
Deep Sea Drilling Project, U.S. Government Printing
Office, Washington, D.C., 1039-1050.
Horsfield, W.T. (1974). Major faults in Jamaica. J. Geol.
Soc. Jam., 14: 1-15.
Houtz, R. E. & Ludwig, W. J. (1977). Structure of
Colombia basin, Caribbean Sea, from profilersonobuoy measurements. J. Geophys. Res., 82(30):
4861-4867
Jacob, K. H., Pacheco, J., Santana, G. (1991).
Seismology and tectonics, in Costa Rica earthquake
Reconnaissance Report: Supplement B. Earthquake
Spectra, 7: 15-33.
Jcome, M. I. (1994). Interpretacin geolgica, ssmica y
gravimtrica de un perfil transandino. Undergraduate
Thesis, Universidad Simn Bolvar, Caracas,
Venezuela.
Jansma, P. E. & Mattioli, G. S. (2005). GPS results
from Puerto Rico and the Virgin Islands: constraints
on tectonic setting and rates of active faulting. In:
Mann, P. (ed.) Active Tectonics and Seismic Hazards
of Puerto Rico, the Virgin Islands, and Offshore
Areas, Geol. Soc. Am. Spec. Paper, 385: 13-30, The
Geological Society of America, Boulder, CO. doi:
10.1130/2007.2428(02).
Jansma, P.E., Mattioli, G.S., Lopez, A., Demets, C.,
Dixon, T.H., Mann, P., Calais E. (2000). Neotectonics
of Puerto Rico and the Virgin Islands, northeastern
Caribbean, from GPS geodesy. Tectonics, 19(6):
1021-1037.
Jones, S. M. (1950). Geology of Gatn Lake and vicinity,
Panam. Bull. Seismol. Soc. Am., 61: 893922.
Jordan, T. H. (1975). The present-day motions of the
Caribbean plate, J. Geophys. Res., 80: 4433-4439.
Jouanne, F., Audemard, F. A., Beck, C., Van Welden,
A., Ollarves, R., Reinoza, C., (2011). Presentday deformation along the El Pilar fault in eastern
Venezuela: Evidence of creep along a major strikeslip fault. Journal of Geodynamics, 51: 398-410. doi:
10.1016/j.jog.2010.11.003
Kafka, A. L. & Weidner, D. J. (1979). The focal
mechanisms and depths of small earthquakes as
determined from Rayleigh-wave radiation patterns,
Bull. Seismol. Soc. Am., 69: 1379-1390.
Kafka, A. L. & Weidner, D. J. (1981). Earthquake
focal mechanisms and tectonic processes along the
Southern boundary of the Caribbean Plate. Journal of
Geophysical Research; 86 (B4): 2877-2888.
Kaniuth, K., Drewes, H., Stuber, K., Temel, H.,

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

Hernndez, J.N., Hoyer, M., Wildermann, E., Kahle,


H., Geiger, G. (1999). Position changes due to recent
crustal deformations along the Caribbean-South
American plate boundary derived from CASA GPS
project. General Assembly of the International Union
of Geodesy and Geophysics (IUGG), Birmingham,
U.K. Poster at Symposium G1 of International
Association of Geodesy.
Keigwin, L. D. (1978). Pleistocene closing of the Isthmus
of Panama, based on biostratigraphic evidence from
nearby Pacific Ocean and Caribbean Sea cores.
Geology, 6: 630-634.
Keigwin, L. D. (1982). Isotopic paleoceanography of the
Caribbean and the East Pacific; Role of Panama uplift
in late Neogene time. Science, 217, 350353.
Keller, G., Zenker, C. E., Stone, S. M. (1989). Neogene
history of the Pacific-Caribbean gateway. J. South
Am. Earth Sci., 2: 73-108.
Kellogg, J. & Bonini, W. (1982). Subduction of the
Caribbean Plate and basement uplifts in the overriding
South-American plate. Tectonics, 1: 251-276.
Kellogg, J. N. & Mohriak, W. (2001). Tectonic and
geological environment of coastal South America.
In: Seeliger, U. and Kjerfve, B. (eds.) Coastal Marine
Ecosystems of Latin America: Ecological Studies,
144: 1-16.
Kellogg, J. N. & Vega, V. (1995). Tectonic development
of Panama, Costa Rica, and the Colombian Andes:
constraints from Global Positioning System geodetic
studies and gravity. In: Mann, P. (ed.) Geologic
and Tectonic Development of the Caribbean Plate
Boundary in Southern Central America: GSA Special
Paper, 295: 75-90.
Kolarsky, R. A. & Mann, P. (1995). Structure and
neotectonics of an oblique subduction margin,
southwestern Panama. In: Mann P. (ed.) Geologic
and Tectonic Development of the Carribbean Plate
Boundary in Southern Central America, Geol. Soc.
Am. Spec. Pap., 295: 131-157.
Kolarsky, R. A., Mann, P., Montero-Pohly, W. (1995).
Island arc response to shallow subduction of the
Cocos Ridge, Costa Rica, Spec. Pap. Geol. Soc. Am.,
295: 235-262.
Ladd, J. (1976). Relative motion of South America with
respect to North America and Caribbean tectonics.
Bulletin of the Geological Society of America, 87:
969-976.
Ladd, J.W., Shih, T.C., Tsai, C.J. (1981). Cenozoic
tectonics of central Hispaniola and adjacent Caribbean
Sea, Am. Assoc. Pet. Geol. Bull., 75: 466-489.
Lafemina, P.C., Dixon, T.H., Strauch, W. (2002).
Bookshelf faulting in Nicaragua: Geology, 30: 751
754, doi:10.1130/0091-7613(2002)030<0751:BFIN>
2.0.CO;2.

Lafemina, P.C., Dixon, T.H., Govers, R., Norabuena,


E., Turner, H., Saballos, A., Mattioli, G., Protti,
M., Strauch, W. (2009). Forearc motion and Cocos
Ridge collision in Central America: Geochemistry
Geophysics Geosystems G3, Q05S14, doi:
10.1029/2008GC002181.
Leroy, S. & Mauffret, A. (1996). Intraplate deformation
in the Caribbean region, J. Geodynamics 21: 113-122.
Linkimer , L., Beck, S. L., Schwartz, S. Y., Zandt, G.,
Levin, V. (2010). Nature of crustal terranes and the
Moho in northern Costa Rica from receiver function
analysis. Geochemistry, Geophysics, Geosystems
G3-, 11(1). Q01S19, doi: 10.1029/2009GC002795
Lonsdale, P. & Klitgord, K. D. (1978). Structure and
tectonic history of the eastern Panama basin, Geol.
Soc. Am. Bull., 89: 981-999.
Lopez, A. M., Stein, S., Dixon, T., Sella, G., Calais, E.,
Jansma, P., Weber, J., La Femina, P. (2006). Is there
a northern Lesser Antilles forearc block?, Geophys.
Res. Lett., 33. doi: 10.1029/2005GL025293
Lpez, S. (2010). Deformacin intraplaca en zonas de
subduccin oblicua y colisin: algunas consideraciones
para su estudio con GPS. I Conferencia en Gestin
de Riesgos del piedemonte llanero. Yopal (Dpto. de
Casanare) Colombia (PPT Presentation).
Lugo, J. & Mann, P. (1992). Colisin oblicua y formacin
de una cuenca foreland durante el Paleoceno tardo
al Eoceno medio; Cuenca de Maracaibo, Venezuela.
Proceedings of III Congreso Geolgico de Espaa
and VIII Congreso Latinoamericano de Geologa,
Salamanca, 4: 60-64.
Lundgren, P., Protti, M., Donnellan, A., Heflin, M.,
Hernandez, E., Jefferson, D. (1999). Seismic cycle
and plate margin deformation in Costa Rica: GPS
observations from 1994 to 1997, J. Geophys. Res.,
104(B12): 28,915-28,926.
Luyendyk, B.P. & Macdonald, K.C. (1976).
Physiography and structure of the inner floor of the
FAMOUS rift valley; Observations with a deeptowed instrument package, Geol. Soc. Am. Bull., 88:
648-663.
Lyon-Caen, H., Barrier, E., Lasserre, C., Franco, A.,
Arzu, I., Chiquin, M., Chiquin, L.M., Duquesnoy,
T., Flores, O., Galicia, O., Luna, J., Molina, E.,
Porras, O., Requena, J., Robles, V., Romero, J.,
Wolf, R. (2006). Kinematics of the North AmericanCaribbean-Cocos plates in Central America from new
GPS measurements across the Polochic-Motagua
fault system: Geophysical Research Letters, 33, p.
L19309, doi: 10.1029/2006GL027694.
Macmillan, D.S. & Ma, C. (1999). VLBI measurements
of Caribbean and South American motion.
Geophysical Research Letters, 26: 919-922.
Malav, G. & Surez, G. (1995). Intermediate-depth

67

68

Franck A. Audemard M.

seismicity in northern Colombia and western


Venezuela and its relationship to Caribbean Plate
subduction. Tectonics, 14: 617-628.
Malfait, B.T. & Dinkelman, M.G. (1972). CircumCaribbean tectonic and igneous activity and the
evolution of the Caribbean Plate. Geol. Soc. Am.
Bull., 83: 251-272.
Manaker, D.M., Calais, E., Freed, A. M., Ali, S. T.,
Przybylski, P., Mattioli, G., Jansma, P., Prpetit,
C., De Chabalier, J. B. (2008). Interseismic plate
coupling and strain partitioning in the northeastern
Caribbean, Geophys. J. Int., 174: 889-903. doi:
10.1111/j.1365-246X.2008.03819.x.
Mann, P. & Burke, K. (1980). Neogene wrench faulting
in the Wagwater Belt, Jamaica. Trans. 9th Caribb.
Geol. Conf., 95-97.
Mann, P. & Burke, K. (1984). Neotectonics of the
Caribbean. Reviews of Geophysics and Space
Physics, 22: 309-362.
Mann, P. & Corrigan, J. (1990). Model for late Neogene
deformation in Panama. Geology, 18: 558-562.
Mann, P. & Kolarsky, R. A. (1995). East Panama
deformed belt; structure, age, and neotectonic
significance. In: Mann P. (ed.) Geologic and Tectonic
Development of the Carribbean Plate Boundary in
Southern Central America, Geol. Soc. Am. Spec.
Pap., 295: 111-130.
Mann, P., Burke, K., Matumoto, T. (1984). Neotectonics
of Hispaniola: Plate motion, sedimentation, and
seismicity at a restraining bend. Earth Planet. Sci.
Lett., 70: 311-324.
Mann, P., Demets C., Wiggins-Grandison, M. (2007).
Toward a better understanding of the Late Neogene
strike-slip restraining bend in Jamaica: geodetic,
geological, and seismic constraints. In: Cunningham,
W. D. and Mann, P. (eds) Tectonics of Strike-Slip
Restraining and Releasing Bends. Geological Society,
London, Special Publications, 290, 239-253.
Mann, P., Draper, G., Burke, K. (1985). Neotectonics
of a strike-slip restraining bend system, Jamaica. In:
Biddie, K.T. and Christie-Blick, N. (eds) Strike-slip
Deformation, Basin Formation, and Sedimentation.
SEPM Spec.Publ., 37: 211-226.
Mann, P., Schubert, C., Burke, K. (1990). Review of
Caribbean neotectonics. In: Dengo, G., Case, J.E.
(eds.) The Caribbean Region, vol. H. Geological
Society of America, Boulder, 307-338.
Mann, P., Draper, G., Lewis, J.F. (1991a). An overview of
the geologic and tectonic development of Hispaniola.
In: P. Mann, G. Draper and J.F. Lewis (eds): Geologic
and Tectonic Development of the North AmericaCaribbean Plate Boundary in Hispaniola. Geol. Soc.
Am., Spec. Pap., 262: 1-28.
Mann, P., Mclauglin, P.P., Cooper, J.C. (1991b).

Geology of the Azua and Enriquillo basins, Dominican


Republic; 2, Structure and tectonics. In P. Mann, G.
Draper and J.F. Lewis (eds): Geologic and Tectonic
Development of the North America-Caribbean Plate
Boundary in Hispaniola. Geol. Soc. Am., Spec. Pap.,
262: 367-389.
Mann, P., Tyburski, S.A., Rosencrantz, E. (1991c).
Neogene development of the Swan Islands restrainingbend complex, Caribbean Sea. Geology, 19: 823-826.
Mann, P., Hempton, M. R., Bradley, D. C., Burke, K.
(1983). Development of pull-apart basins, J. Geol.,
91, 529-554.
Mann, P., Taylor, F.W., Edwards, R. L., Ku, T.-L. (1995).
Actively evolving microplate formation by oblique
collision and sideways motion along strike-slip faults:
An example from the northeastern Caribbean plate
margin. Tectonophysics, 246 (1995) 1-69
Mann, P., Calais, E., Ruegg, J. C., Demets, C., Jansma,
P., Mattioli, G. (2002). Oblique collision in the
northeastern Caribbean from GPS measurements and
geological observations. Tectonics, 21(6) 1057. doi:
10.1029/2001TC001304.
Manton, W.I. (1987). Tectonic interpretation of the
morphology of Honduras. Tectonics, 6: 633-651.
Marshall, J. S., Fisher, D. M., Gardner, T. W. (2000).
Central Costa Rica deformed belt: Kinematics of
diffuse faulting across the western Panama block.
Tectonics 19: 468-492.
Mascle, A. & Letouzey, P. (1990). Geological Map of
the Caribbean. Editions Technip, Paris (2 sheets).
Mascle, A., Biju-Duval, B., De Clarens, P., Munsch,
H. (1986). Growth od accretionary prisms: tectonic
processes from Caribbean examples. In: Wezel, F.
C. (ed) The Origin of Arcs. 1 ed., Elsevier Science
Publishers, Amsterdam, 576 pp.
Mascle, A., Biju-Duval, B., Letouzey, J., Bellizzia, A.,
Aubouin, J., Blanchet, R., Stephan, J.F., Beck, C.
(1979). Estructura y evolucin de los mrgenes este y
sur del Caribe (anlisis de los problemas del Caribe).
Bulletin du Bureau de Recherches Gologiques et
Minires, 3/4: 171-184.
Masson, D. G. and Scanlon, K. M. (1991). The
neotectonic setting of Puerto Rico. Geol. Soc. Am.
Bull., 103: 144-154.
Mauffret, A. & Jany, I. (1990). Collision et tectonique
dexpulsion le long de la frontire Nord-Carabe.
Oceanol. Acta, 10: 97-116.
Mauffret, A. & Leroy, S. (1997). Seismic stratigraphy
and structure of the Caribbean igneous province,
Tectonophysics, 283: 61-104.
Mauffret, A. & Leroy, S. (1999). Neogene intraplate
deformation of the Caribbean plate at the Beata
Ridge. In: Mann P. (ed) Sedimentary Basins of the
World, vol. 4, Caribbean Basins, Elsevier Sci., New

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

York, 627-669.
Mccaffrey, R. (1992). Oblique plate convergence, slip
vectors, and forearc deformation, J. Geophys. Res.,
97: 8905-8915.
Mckenzie, D. P. (1972). Active tectonics of the
Mediterranean region. Geophys. J. R. Astron. Soc.,
30: 109-185.
Mckenzie, D. P. (1977). The initiation of trenches: A
fine amplitude instability. In: Talwani, M. and Pitman
W. C. III (Eds) Island Arcs, Deep Sea Trenches, and
Back-arc Basins,), Am. Geophys. Union, Washington,
D.C., 5761.
Mckenzie, D. P. (1978). Active tectonics of the AlpineHimalayan Belt: The Aegean Sea and surrounding
regions, Geophys. J. R. Astron. Soc., 52: 211-254.
Mckenzie, D. P. & Morgan, W. J. (1969). Evolution of
triple junctions. Nature, 224: 125-133Meja, J. and Meyer, H. (2004) Modelo detallado
preliminar de la sismicidad en el Occidente de
Colombia. Observatorio sismolgico del Sur
Occidente. OSSO. Proceedings I Congreso
Latinoamericano de Sismologa. Armenia.
Minster, J. & Jordan, T. (1978). Present-day plate
motions. Journal of Geophysical Research, 83: 53315354.
Molnar, P. & Sykes, L. (1969). Tectonics of the
Caribbean and Middle America Regions from focal
mechanisms and Seismicity. Geological Society of
America Bulletin, 80: 1639- 1684.
Montero-Pohly, W. (1994a). Neotectonics and related
stress distribution in a subduction-collisional zone:
Costa Rica. - In: Seyfried, H. (ed.) Geology of an
evolving island arc: southeastern Central America. Profil, 7: 125-141.
Montero-Pohly, W. (1994b). Sismicidad y neotectnica.
- In: Denyer, P. y Kussmaul, S. (eds.) Atlas geolgico
Gran Area Metropolitana: 147-160. Ed. Tecnolgica
de Costa Rica.
Montero-Pohly, W. (2001). Neotectnica de la regin
central de Costa Rica: frontera oeste de la microplaca
de Panam. Revista Geolgica de Amrica Central,
24: 29-56.
Montero-Pohly, W. & Dewey, J. W. (1982). Shallowfocus seismicity, composite focal mechanism, and
tectonics of the Valle Central, Costa Rica. - Seism.
Soc. Am. Bull. 71: 1611-1626.
Montero-Pohly, W., Denyer, P., Barquero, R.,
Alvarado, G. E., Cowan, H., Machette, M. N.,
Haller, K. M., Dart, R. L. (1998). Map and database
of Quaternary faults and folds in Costa Rica and its
offshore regions. - 63 pgs. + mapa escala 1: 500 000,
Open-File Report 98-481, U. S. Geological Survey,
Boulder.
Montes, C., Bayona, G., Cardona, A., Buchs, D. M.,

Silva, C. A., Morn, S., Hoyos, N., Ramrez, D. A.,


Jaramillo, C. A., Valencia, V. (2012). Arc-continent
collision and orocline formation: Closing of the
Central American seaway, J. Geophys. Res., 117,
B04105, doi:10.1029/2011JB008959.
Moore, G. T. & Fahlquist, D. A. (1976). Seismic profile
tying Caribbean DSDP sites 153, 151, and 152, Geol.
Soc. Am. Bull, 87: 1609-1614,
Moore, J. C. & Biju-Duval, B. (1984). Tectonic synthesis,
Deep Sea Drilling Project Leg 78 A: structural
evolution of offscraped and underthrust sediment,
northern Barbados. Initial Reports of the Deep Sea
Drilling Project, Vol. 78A, U.S. Government Printing
Office, Washington, D.C., 601-621.
Mora-Pez, H. (2011). South America-Caribbean
boundary: Existing geodetic infrastructure in
Colombia GEORED Project (Spatial geodesy
network for geodynamics research)-. Presentation
given at I COCONet Workshop, Puerto Rico
(available from UNAVCO webpage).
Muehlberger, W. R. & Ritchie, A. (1975). CaribbeanAmerican plate boundary in Guatemala and Southern
Mexico as seen on Skylab IV orbital photography,
Geology, 3: 232-235.
Mller, R. D., Royer, J. Y., Cande, S. C., Roest, W.
R., Maschenkov, S. (1999). New constraints on the
Late Cretaceous/Tertiary plate tectonic evolution of
the Caribbean, In: Mann, P. (ed) Sedimentary Basins
of the World, vol. 4, Caribbean Basins, Elsevier Sci.,
New York, 33-59.
ODP LEG 110 SCIENTIFIC PARTY (1987). Expulsion
of fluids from depth along a subduction-zone
decollement horizon. Nature, 326(6115): 785-788.
Ojeda, A. & Havskov, J. (2001). Crustal structure and
local seismicity in Colombia, J. Seismol., 5(4): 575593.
Orihuela, N. & Cuevas, J. (1992). Modelaje
sismogravimtrico de perfiles regionales del Caribe
Central. XIII Caribbean Geological Conference, Cuba
(Abstract).
Palma, M., Audemard, F. A., Romero, G. (2010).
Nuevos mecanismos focales para Venezuela y reas
vecinas 2005-2008: importancia de la densificacin y
distribucin de la red sismolgica nacional. Revista
Tcnica Facultad de Ingeniera de la Universidad del
Zulia, 33(2): 1-14.
Pedoja, K. (2003). Les terrasses marines de la marge
Nord Andine (Equateur et Nord Prou): Relations
avec le contexte godynamique. Ph.D. thesis, Univ.
Paris VI, Paris, 413 pp.
Perfit, M. R., Heezen, B. C., Rawson, M.R., Donnelly,
T.W. (1980). Chemistry, origin, and significance of
metamorphic rocks from the Puerto Rico Trench.
Mar. Geol., 34: 125-156.

69

70

Franck A. Audemard M.

Prez, O. J. & Aggarwal, Y. (1981). Present-day


tectonics of southeastern Caribbean and northeastern
Venezuela. Journal of Geophysical Research, 86:
10,791-10,805.
Prez, O. J., Bilham, R., Bendick, R., Hernndez, N.,
Hoyer, M., Velandia, J., Moncayo, C., Kozuch, M.
(2001a). Velocidad relativa entre las placas del Caribe
y Sudamrica a partir de observaciones dentro del
sistema de posicionamiento global (GPS) en el norte
de Venezuela. Interciencia, 26: 69-74.
Prez, O. J., Bilham, R., Bendick, R., Velandia, J.R.,
Hernndez, N., Moncayo, C., Hoyer, M., Kozuch, M.
(2001b). Velocity field across the southern Caribbean
Plate boundary and estimates of Caribbean/South
American Plate motion using GPS geodesy 19942000. Geophysical Research Letters, 28: 2987-2990.
Prez, O. J., Bilham, R., Sequera, M., Molina, L.,
Gavotti, P., Codallo, H., Moncayo, C., Rodrguez,
C., Velandia, R., Guzmn M., Molnar, P. (2011).
Campo de Velocidades GPS en el Occidente de
Venezuela: Componente lateral derecha asociada
a la Falla de Bocon y componente convergente
perpendicular a Los Andes. Interciencia, 36(1): 39-44.
Petersen, M., Schweig, E., Mueller, C., Harmsen,
S., Frankel A. (2005). Preliminary update of the
probabilistic seismic hazard analysis for sites along
the Panam Canal Zone, 25 p.
Pindell, J. L. & Dewey, J. (1982). Permo-Triassic
reconstruction of western Pangea and the evolution
of the Gulf of Mexico/Caribbean region. Tectonics,
1: 179-211.
Pindell, J. L., Higgs, R., Dewey, J. (1998). Cenozoic
palinspatic reconstruction, paleogeographic evolution
and hydrocarbon setting of the northern margin of
South America. In: Pindell, J. and Drake, C. (eds)
Paleogeographic Evolution and Non-glacial Eustasy,
Northern South America. Society for Sedimentary
Geology, Special Publications, 58: 45-85.
Porras, L. (2000). Evolucin tectnica y estilos
estructurales de la regin costa afuera de las cuencas
de Falcn y Bonaire. Proceedings of VII Congreso
Bolivariano Exploracin Petrolera en las Cuencas
Subandinas, Caracas, Venezuela, 279-292.
Plafker, G. (1976). Tectonic aspects of the Guatemalan
earthquake of 4 February 1976, Science, 193: 12011208.
Plafker, G. & Ward, S. N. (1992). Backarc thrust
faulting and tectonic uplift along the Caribbean
sea coast during the April 22, 1991, Costa Rica
earthquake. Tectonics, 11: 709-718.
Protti, M. & Schwartz, S. (1994). Mechanics of back
arc deformation in Costa Rica: Evidence from an
aftershock study of the April 22, 1991, Valle de la
Estrella, Costa Rica, earthquake (Mw 7:7), Tectonics

13(5): 10931107.
Protti, M., Gendel, F., Mcnally, K. (1994). The
geometry of the Wadati-Benioff zone under southern
Central America and its tectonic significance: Results
from a high-resolution local seismographic network,
Phys. Earth Planet. Inter., 84: 271-287.
Protti, M., Gonzlez, V. Freymueller, J., Doelger, S.
(2012). Isla del Coco, on Cocos Plate, converges with
Isla de San Andrs, on the Caribbean Plate, at 78mm/
yr. Int. J. Trop. Biol., 60 (Suppl. 3): 33-41.
Pubellier, M., Vila, J.-M., Bolsson, D. (1991). North
Caribbean neotectonic events: the Trans-Haitian fault
system. Tertiary record of an oblique transcurrent
shear zone uplifted in Hispaniola, Tectonophysics,
194: 217-236.
Pubellier, M., Mauffret, A., Leroy, S., Vila, J.M.,
Amilcar, H. (2000). Plate boundary readjustment in
oblique convergence: Example of the Neogene of
Hispaniola, Greater Antilles, Tectonics, 19: 630-648.
Reid, J. A., Plumley, P. W., Schellekens, J. H.
(1991). Paleomagnetic evidence for late Miocene
counterclockwise rotation of North Coast carbonate
sequence, Puerto Rico. Geophys. Res. Lett., 18: 565568,.
Robertson, P. & Burke, K. (1989). Evolution of the
southern Caribbean Plate boundary in the vicinity
of Trinidad and Tobago. Bulletin of the American
Association of Petroleum Geologists, 73: 490-509.
Rockwell, T. K., Gath, E., Gonzlez, T., Madden, C.,
Verdugo, D., Lippincott, C., Dawson, T., Owen, L.
A., Fuchs, M., Cadena, A., Williams, P., Weldon,
E., Franceschi, P. (2010). Neotectonics and
paleoseismology of the Limn and Pedro Miguel
faults in Panam: Earthquake hazard to the Panam
Canal. Bull. Seismol. Soc. Am., 100(6): 3097-3129.
doi: 10.1785/0120090342.
Rosencrantz, E. & Mann, P. (1991). SeaMARC II
mapping of transform faults in the Cayman Trough,
Caribbean Sea. Geology, 19: 690-693.
Rosencrantz, E., Ross, M., Sclater, J. (1988). Age
and spreading history of the Cayman Trough as
determined from depth, heat flow, and magnetic
anomalies. Journal of Geophysical Research, 93:
2141-2157.
Ruiz, C., Davis, N., Bentham, P., Price, A., Carvajal,
D. (2000). Structure and tectonic evolution of the
South Caribbean basin, southern offshore Colombia:
a progressive accretionary system. Proceedings of VII
Simposio Bolivariano Exploracin Petrolera en las
Cuencas subandinas, Caracas, Venezuela, 334-355.
Russo, R. (1999). Dynamics and deep structure of
the southeastern Caribbean-South America plate
boundary zone: relationship to shallow seismicity.
AGU Spring Meeting, Boston, MA, S228 (Abstract).

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

Russo, R. & Speed, R. C. (1992). Oblique collision and


tectonic wedging of the South American continent
and Caribbean terranes. Geology, 20: 447-450.
Snchez-Barreda, L. A. (1981). Geologic evolution of
the continental margin of the gulf of Tehuantepec,
southern Mxico. PhD Thesis, University of Texas,
Austin, Texas.
Scholz, C.H. & Small, C. (1997). The effect of seamount
subduction on seismic coupling. Geology, 25: 487490.
Schubert, C. (1979). El Pilar Fault Zone, northeastern
Venezuela: Brief review, Tectonophysics 5(2): 447455.
Schubert, C. (1980). Late Cenozoic pull-apart basins,
Bocon Fault Zone, Venezuelan Andes. J. Struct.
Geol., 2: 463-468.
Schubert, C. (1982a). Neotectonics of Bocon Fault,
western Venezuela. Tectonophysics, 85: 205-220.
Schubert, C. (1982b). Origin of the Cariaco Basin,
southern Caribbean Sea. Mar. Geol., 47: 345-360.
Schubert, C. (1984). Basin formation along the BoconMorn-E1Pilar Fault System, Venezuela, J. Geophys.
Res., 89: 5711-5718.
Schweig, E., Cowan, H., Gomberg, J., Pratt, T., Tenbrink,
A. (1999). Design earthquakes for the evaluation of
seismic hazard at Gatn Dam and vicinity, Report
to the Panam Canal Commission for Interagency
Agreement Number CNP-93786-NN-29 between
the Panam Canal and the U.S. Geological Survey (9
August 1999), 60 p.
Sella, G. F., Dixon, T. H., Mao, A. (2002). REVEL: A
model for recent plate velocities from space geodesy.
Journal of Geophysical Research, 107. B4. doi:
10.1029/2000JB000033.
Singer, A. & Audemard, F. A. (1997). Aportes de
Funvisis al desarrollo de la geologa de fallas activas
y de la paleosismologa para los estudios de amenaza
y riesgo ssmico. Academia de las Ciencias Naturales,
Matemticas y Fsicas, Publicacin Especial, 33: 2538.
Silver, E. A., Case, J., Macgillary, H. (1975).
Geophysical study of the Venezuela borderland.
Bulletin of Geological Society of America, 86: 213226.
Silver, E. A., Galewsky, J., Mcintosh, K. D. (1995).
Variation in structure, style, and driving mechanism
of adjoining segments of the north Panama deformed
belt. In Mann P. (ed.): Geologic and Tectonic
Development of the Caribbean Plate Boundary in
Southern Central America. Geol. Soc. Am. Spec. Pap.
295, 225234.
Silver, E. A., Reed, D. L., Tagudin, J. E., Heil, D. J.
(1990). Implications of the north and south Panama
thrust belts for the origin of the Panama orocline.

Tectonics, 9(2): 261-281.


Simkin, T., Siebert, L., Mcclelland, L., Bridge, D.,
Newhall, C., Latter, J.H. (1981). Volcanoes of the
World. Hutchinson Ross Publishing Co, Stroudsburg,
Pennsylvania. 232 pp.
Sinton, C. W., Duncan, R. A., Storey, M., Lewis, J.,
Estrada, J. J. (1998). An oceanic flood basalt province
within the Caribbean Plate, Earth Planet. Sci. Lett.,
155, 221235, doi:10.1016/S0012-821X(97)00214-8.
Sobiesiak, M., Alvarado, L., Vsquez, R. (2002).
Sismicidad reciente del Oriente de Venezuela.
Proceedings XI Congreso Venezolano de Geofsica
(CD-ROM format).
Sobiesiak, M., Alvarado, L., Vsquez, R. (2005). Recent
seismicity in northeastern Venezuela and tectonic
implications. Revista de la Facultad de Ingeniera de
la Universidad Central de Venezuela,20(4): 43-52.
Soulas, J.-P. (1986). Neotectnica y tectnica activa
en Venezuela y regiones vecinas. Proceedings of VI
Congreso Geolgico Venezolano, 10: 6639-6656.
Soulas, J.-P. (1991). El sismo de Limn-Changuinola
(Costa Rica-Panam). Informe Preliminar de Misin.
22 p.
Speed, R. (1985). Cenozoic collision of the lesser Antilles
arc and continental South America and the origin of
the El Pilar Fault. Tectonics, 4: 41-69.
Spikings, R.A., Winkler, W., Seward D., Handler, R.
(2001). Along-strike variations in the thermal and
tectonic response of the continental Ecuadorian
Andes to the collision with heterogeneous oceanic
crust. Earth and Planetary Science Letters, 186: 57-73
Spikings, R.A., Crowhurst , P.V., Winkler, W.,
Villagomez, D. (2010). Syn- and post-accretionary
cooling history of the Ecuadorian Andes constrained
by their in-situ and detrital thermochronometric
record. Journal of South American Earth Sciences 30:
121-133
Stephan, J.-F. (1982). Evolution godinamique du
domaine Carabe,Andes et chane Carabe sur la
transversale de Barquisimeto (Vnzula). PhD
thesis, Paris, 512 pp.
Stephan, J.-F. (1985). Andes et chane carabe sur la
transversale de Barquisimeto (Vnzula). Evolution
godynamique.
Proceedings
of
Symposium
Godynamique des Carabes, Paris, Editions Technip,
Paris, 505-529.
Stephan, J.-F., Blanchet, R., Mercier De Lepinay, B.
(1986). Northern and southern Caribbean festoons
(Panam, Colombia-Venezuela and Hispaniola-Puerto
Rico), interpreted as pseudosubductions induced
by the East-West shortening of the Pericaribbean
continental frame. In: Wezel, F.C. (ed) The Origin of
Arcs, Elsevier Science Publishers, Amsterdam. 576
pp.

71

72

Franck A. Audemard M.

Stephan, J.-F., Mercier De Lepinay, B., Calais, E.,


Tardy, M., Beck, C., Carfantan, J.-C., Olivet, J.-M.,
Vila, J.-M., Bouysse, P., Mauffret, A., Bourgois, J.,
Thery, J.-M., Tournon, J., Blanchet, R., Dercourt,
J. (1990). Paleogeodynamics maps of the Caribbean:
14 steps from Lias to Present. Bulletin Societ
Gologique de France, 6: 915-919.
Stein, S., Demets, C., Gordon, R. G., Brodholt, J., Engeln,
J. F., Wiens, D. A., Argus, D., Lundgren, P., Stein, C.,
Woods, D. (1988). A test of alternative Caribbean
plate relative motion models: Journal of Geophysical
Research, 93: 3041-3050.
Stewart, R. H., Stewart, J. L., Woodring, W. P. (1980).
Geologic map of the Panam Canal and vicinity,
Republic of Panam, U.S. Geol. Surv. Misc. Invest.
Series, Map I-1232, scale: 1:100,000.
Surez, G., Pardo, M., Domnguez, J., Ponce, L.
Montero-Pohly, W., Boschini, I., Rojas, W. (1995).
The Limn, Costa Rica earthquake of April 22, 1991:
Back arc thrusting and collisional tectonics in a
subduction environment, Tectonics, 14: 518-530.doi:
10.1029/94TC02546.
Suter, F., Sartori, M., Neuwerth, R., Gorin, G. (2008).
Structural imprints at the front of the ChocPanama
indenter: Field data from the North Cauca Basin,
central Colombia, Tectonophysics 460, 134-157.
Sykes, L.R., Mccann, W.R., Kafka, A.L. (1982). Motion
of Caribbean plate during last 7 million years and
implication for earlier Cenozoic movements. J.
Geophys. Res., 87: 10656-10676.
Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A.,
Herv, P., Harmen, B., Olaya, J., Rivera, C. (2000).
Geodynamics of the northern Andes: subductions and
intracontinental deformation (Colombia). Tectonics
19: 787-813.
Tajima, F. & Kikuchi, M. (1995). Tectonic implications
of the seismic ruptures associated with the 1983 and
1991 Costa Rica earthquakes. In Mann, P. (ed.):
Geologic and Tectonic Development of the Caribbean
Plate Boundary in Southern Central America, Spec.
Pap. Geol. Soc. Am., 295: 327-340.
Talwani, M., Windisch, C., Stoffa, P., Buhl, P.,
Houtz, R. (1977). Multichannel seismic study in the
Venezuelan Basin and the Curacao Ridge. In Talwani
M. and Pitman W.C. (eds.): Island Arcs, Deep Sea
Trenches and Back-Arc Basins. Maurice Ewing Ser.,
AGU, Washington, D.C., 1(III): 83-98.
Tapponnier, P., Peltzer, G., Le Dain, A.Y., Armijo, R.,
Cobbold, P. (1982). Propagating extrusion tectonics
in Asia: new insights from simple experiments with
plasticine. Geology, 10 (12): 611-616.
Tarr, A. C., Rhea, S., Hayes, G., Villaseor, A.,
Furlong, K., Benz, H. (COMP.). (2010). Seimicity
of the Earth 1900-2007: Caribbean Plate and Vicinity.

USGS (Poster/Map)
Ten Brink, U., Danforth, W., Pollonini, C., Andrews,
B., Llanes, P., Smith, S., Parker, E., Uozumi, T.
(2004). New seafloor map of the Puerto Rico Trench
helps assess earthquake and tsunami hazards. Eos
(Transactions), 85: 349360.
Ten Brink, U. S., Marshak S., Granja-Brua, J.L. (2009). Bivergent thrust wedges surrounding
oceanic island arcs: Insight from observations and
sandbox models of the northeastern Caribbean plate.
Geological Society of America Bulletin, 121: 15221536.
Trenkamp, R., Kellogg, J. N., Freymueller, J. T., Mora,
H. (2002). Wide plate margin deformation, southern
Central America and northwestern South America,
CASA GPS observations. J. South Am. Earth Sci., 15,
157171, doi: 10.1016/S0895-9811(02)00018-4.
USGS NATIONAL EARTHQUAKE INFORMATION
CENTER AND THE MIDDLE AMERICA
SEISMOGRAPH (MIDAS) CONSORTIUM (1998).
Caribbean Seismicity 1900-1994. USGS ofr 98-223,
Reston, Va, USA (Map).
Van der Hilst, R. (1990). Tomography with P, PP and
pP delay-time data and the three dimensional mantle
structure below the Caribbean region. Geologica
Ultraiectina, vol. 67. University of Utrecht,
Netherlands, 250 pp.
Van der Hilst, R. & Mann, P. (1993). Tectonic
implication of tomographic images of subducted
lithosphere beneath northwestern South America.
AAPG/SVG Internatinal Congress and Exhibition,
Caracas, p. 71 (Abstract).
Van der Hilst, R. & Mann, P. (1994). Tectonic
implication of tomographic images of subducted
lithosphere beneath northwestern South America.
Geology, 22: 451-454.
Van Gestel, J.P., Mann, P., Dolan, J.F., Grindlay, N.R.
(1998). Structure and tectonics of the upper Cenozoic
Puerto Rico-Virgin Islands carbonate platform
as determined from seismic reflection studies. J.
Geophys. Res., 103: 30,505-30,530.
Vargas, C. A. & Mann, P. (2013). Tearing and Breaking
Off of Subducted Slabs as the Result of Collision of
the Panama Arc-Indenter with Northwestern South
America. Bulletin Seismological Society of America,
103(3): 2025-2046. doi: 10.1785/0120120328
Wadge, G. & Burke, K. (1983). Neogene Caribbean
plate rotation and associated Central American
tectonic evolution. Tectonics, 2: 633643.
Wadge, G. & Wooden, J. L. (1982). Late Cenozoic
alkaline volcanism in the northwestern Caribbean:
Tectonic setting and Sr isotopic characteristics, Earth
Planet. Sci. Lett., 57: 35-46,
Weber, J. C., Dixon, T. H., Demets, C., Ambeh, W. B.,

Captulo II - Active Block Tectonics in and Around The Caribbean: A Review

Jansma, P., Mattioli, G., Saleh, J., Sella, G., Bilham,


R., Perez, O. (2001). GPS estimate of relative motion
between the Caribbean and South American plates,
and geologic implications for Trinidad and Venezuela.
Geology, 29: 75-78.
Westbrook, G. K. (1975). The structure of the crust and
upper mantle in the region of Barbados and the Lesser
Antilles. Geophys. J. R. Astron. Soc., 43: 201-242.
Westbrook, G. K. (1982). The Barbados Ridge Complex:
tectonics of a mature forearc system. In: Legett, J.K.
(ed) Trench and Forearc Geology. Spec. Publ., Geol.
Soc. London. 10: 270-275.
Westbrook, G. K., Hardt, N. C., Heath, R. (1995).
Structure and tectonics of the Panama-Nazca
boundary. In: Mann P. (ed) Geologic and Tectonic
Development of the Caribbean Plate Boundary in
Southern Central America, Geol. Soc. Am. Spec.
Pap., 295: 91-109.
Weyl, R. (1980). Geology of Central America. Gebruder
Borntraeger, Berlin. 371 pp.
White, R. A. & Harlow, D. H. (1993). Destructive
upper-crustal earthquakes of Central America since
1900. Bull. Seismol. Soc. Am., 83: 1115-1142.
White, S.M., Trenkamp, R., Kellogg, J. N. (2003).
Recent crustal deformation and the earthquake cycle
along EcuadorColombia subduction zone. Earth and
Planetary Science Letters, 216: 231-242.
Wolters, B. (1986). Seismicity and tectonics of
southern Central America and adjacent regions with
special attention to the surroundings of Panama.
Tectonophysics, 128: 21-46.

Woodring, W. P. (1957). Geology and paleontology of


Canal Zone and adjoining parts of Panama, U.S. Geol.
Surv, Profess. Paper 306-A, 145 p.
Ysaccis, R. (1997). Tertiary evolution of the northeastern
Venezuela offshore. PhD thesis, Rice University,
Houston, TX.
Ysaccis, R., Cabrera, E., Del Castillo, H. (2000). El
sistema petrolfero de la Blanquilla, costa afuera
Venezuela. Proceedings VII Congreso Bolivariano
Exploracin Petrolera en las Cuencas Subandinas,
Caracas, 411-425.

73

74

Franck A. Audemard M.

You might also like