You are on page 1of 11

ARTICLE IN PRESS

Marine and Petroleum Geology 24 (2007) 277287


www.elsevier.com/locate/marpetgeo

Deformation style and basin-ll architecture of the offshore Limon


back-arc basin (Costa Rica)
Christian Brandesa,, Allan Astorgab, Stefan Backc, Ralf Littked, Jutta Winsemanna
a

Institut fur Geologie, Leibniz Universitat Hannover, Callinstr. 30, 30167 Hannover, Germany
b
Escuela Centroamericana de Geologa, San Jose, Costa Rica
c
Geologisches Institut, RWTH Aachen, Wullnerstr. 2, 52062 Aachen, Germany
d
Lehrstuhl fur Geologie, Geochemie und Lagerstatten des Erdols und der Kohle, RWTH Aachen, Lochnerstr. 4-20, 52056 Aachen, Germany
Received 10 March 2007

Abstract
The Limon back-arc basin, located at the eastern coast of Costa Rica, is part of the southern Central American arc-trench system.
Basin evolution started in Late Cretaceous times as response to the onset of the subduction of the Farallon Plate below the Caribbean
Plate. The Limon Basin can be subdivided into a northern and a southern sub-basin separated by the EW trending Trans Isthmic Fault
System. A regional grid of offshore seismic lines allows the comparative analysis of the basin-ll architecture and the deformation style in
both sub-basins. The northern sub-basin shows exclusively normal faults. The basin-ll architecture resembles a passive continental
margin setting showing a distinct wedge-shaped geometry and seaward propagating depositional units. The area of the Rio San Juan
Delta is characterized by gravity-driven, deltaic deformation exhibiting a series of listric growth faults both on the shelf and in slope
areas. In contrast to the northern basin, the southern sub-basin is characterized by the development of thin-skinned fold-and-thrust belt
tectonics best recorded in concentric or asymmetric hangingwall anticlines separated by listric or planar southwestward dipping thrust
faults. Due to the pronounced NE-propagating of folding the shelf is broader and the slopes are steeper in the southern sub-basin,
compared with the northern sub-basin. Thus both sub-basins show a very different tectonic style. In the North Limon Basin an
extensional regime established, whereas in the South Limon Basin compression dominated. However, in both sub-basins the basal
detachment is probably controlled by a lithological change from limestone to shale.
r 2007 Elsevier Ltd. All rights reserved.
Keywords: Basin-ll architecture; Delta tectonics; Limon Basin; Back-arc basin

1. Introduction
The Limon back-arc basin is a complex basin-system. An
extensional back-arc area (the North Limon Basin) and a
compressional retro-arc foreland basin (the South Limon
Basin) can be observed side-by-side. Both sub-basins are
very similar regarding the stratigraphy and lithology of
their ll, but differ signicantly in the structural style and
basin morphology. Previous work on the stratigraphy and
sedimentology of the Limon Basin is published in e.g.
Amann (1993), Bottazzi et al. (1994), Escalante and
Astorga (1994), Mende (2001) and Campos (2001). Few
Corresponding author. Tel.: +49 511 762 4391; fax: +49 511 762 2172.

E-mail address: brandes@gcowi.uni-hannover.de (C. Brandes).


0264-8172/$ - see front matter r 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.marpetgeo.2007.03.002

studies focused on the structural evolution of the onshore


parts of the basin (Fernandez et al., 1994; Suarez et al.,
1995). However, only little is known of the offshore
geology of the Limon Basin (Sheehan et al., 1990;
Barrientos et al., 1997). The study presented is based on
the analysis of a regional grid of offshore 2D seismic
reection lines, which was made available by the Costa
Rican Ministry of Environment and Energy (MINAE).
Use of this data set allows to analyse the basin-ll
architecture and the structural style of the offshore part
of the Limon Basin, including a detailed stratigraphic and
structural subsurface interpretation, supported by well
integration and a 3D visualization of the subsurface
geometries of the Limon Basin. The aim of the study is
to document and discuss differences as well as similarities

ARTICLE IN PRESS
278

C. Brandes et al. / Marine and Petroleum Geology 24 (2007) 277287

of the southern and the northern parts of the Limon Basin,


in particular focusing on the relation between tectonics and
sedimentation and the delineation of the principle controlling factors for deformation.
2. Geological setting
The geology of southern Central America has been
intensively discussed by numerous authors (Weyl, 1980;
Astorga, 1988; Pindell et al., 1988; Ross and Scotese, 1988;
Lundberg, 1991; Seyfried et al., 1991; Winsemann and
Seyfried, 1991; Frisch et al., 1992; Weinberg, 1992;
Winsemann, 1992; Amann, 1993; Krawinkel and Seyfried,
1994; von Huene and Fluh, 1994; Astorga, 1997; Werner
et al., 1999; Maresch et al., 2000; Abratis and Worner,
2001; Grafe et al., 2002; Calvo, 2003; Krawinkel, 2003).
The Limon Basin extends along the present-day coastal
plain and continental shelf of eastern Costa Rica (Fig. 1a).
The basin is approx. 250 km long and 130 km wide,
bounded in the north by the Hess Escarpment, in the west
and in the south by the volcanic arc, and in the east by the
offshore continuation of the Trans Isthmic Fault System
(Fig. 1a). The evolution of the Limon Basin started in Late
Cretaceous times as response to the onset of subduction of
the Farallon Plate below the Caribbean Plate. The Limon
Basin is underlain by thickened oceanic crust of the
Caribbean Plate. The thickened crust is the result of

mantle plume activity in the Mid-Cretaceous (Burke et al.,


1978; Donnelly, 1994; Meschede et al., 2000). The ll of
the Limon Basin consists of Late Cretaceous to Recent
deep-marine, shallow marine, continental and volcaniclastic rocks (Sheehan et al., 1990; Coates et al., 1992,
2003; Amann, 1993; Bottazzi et al., 1994; Fernandez
et al., 1994; McNeill et al., 2000; Mende, 2001; Campos,
2001) (Fig. 1b). The oldest sediments in the Limon Basin,
known from wells and outcrops consist of 1280 m
thick pelagic carbonates and intercalated volcaniclastic
sediments of Late Campanian to Maastrichtian age
(Changuinola Formation) (Campos, 2001). The deep-water
sediments of the Changuinola Formation are overlain by
3000 m thick Paleocene to Early Eocene coarse-grained
volcaniclastic turbidites, debris-ow deposits, lava ows
and tuffs of the Tu s Formation, which form a prograding
apron-system (Mende, 2001) (Fig. 1b). During Eocene
to Oligocene times compressional tectonic caused an
inversion of the basin margin and the formation of
structural highs. On top of these uplifted areas, carbonate
depositional systems were established and 150200 m
thick carbonates of the Middle Eocene to Oligocene Las
Animas Formation were deposited (Amann, 1993;
Mende, 2001). Contemporaneously, 700900 m hemipelagic mudstones, calcareous turbidites, and carbonate debrisow deposits (Senosri Formation) accumulated in the
adjacent basinal areas (Mende, 2001). During the Late

Fig. 1. (a) Geological map of Costa Rica. The Limon Basin extends along the Caribbean coast (modied after Barboza et al., 1997; Fernandez et al.,
1997). (b) Stratigraphy of the Limon Basin (modied after Campos, 2001; Mende, 2001).

ARTICLE IN PRESS
C. Brandes et al. / Marine and Petroleum Geology 24 (2007) 277287

279

Oligocene an unconformity developed, probably caused by


regional-scale uplift of the island-arc and a major sea-levelfall (Amann, 1993). Subsequently, carbonate ramps formed
on top of this unconformity. These carbonate ramps were
overlain by 2000 m thick shallow-water siliciclastic
sediments of the Late Miocene Uscari Formation (Amann,
1993; Mende, 2001), interpreted to represent delta-inuenced shelf deposits (Fig. 1b). During Late Miocene to
Pliocene times, shallow-water mudstones, sandstones and
conglomerates of the R o Banano Formation were
deposited and are locally associated with patch-reefs
(Mende, 2001). The estimated thickness ranges from 380
to 1800 m (Amann, 1993; Bottazzi et al., 1994).
Today the Limon Basin can be subdivided into a
northern and into a southern sub-basin, which are
separated by the EW trending Trans Isthmic Fault
System (Fig. 1). Both sub-basins exhibit a similar
stratigraphy. In contrast to the lithological similarities,
the morphology as well as the structural style of the North
and South Limon Basins are very different. The North
Limon Basin is more than 7 km deep and shows little
evidence of deformation (except for deltaic faulting in the
region of the Rio San Juan Delta). The South Limon Basin
is 79 km deep and the basin-ll has been deformed by
NE-directed folding and thrusting (Bottazzi et al., 1994;
Mende, 2001; Campos, 2001).
3. Database and methods
The data used in this study include a grid of 2D seismic
reection lines, comprising NESW directed in-lines and
NWSE directed cross-lines (Fig. 2). The seismic grid in the
South Limon Basin is denser than in the North Limon
Basin, but covers a smaller area. Stratigraphic and
lithologic control for the seismic interpretation is derived
from one onshore well for each sub-basin. Well 1
penetrates Pleistocene to Eocene sandstones, shales and
limestones, and well 2 penetrates Pleistocene to Miocene
sandstones, shales and limestones. Seismic interpretation
was performed with the software package Kingdom
Suiter. Well 1 is the key well for the interpretation of
the seismic sections in the North Limon Basin. Well 1 was
tied to section 22 (Fig. 2). This section is the basis for the
interpretation of all other seismic lines in the North Limon
Basin. The SW end of section 6 and section 10 is close to
section 22, therefore it was possible to trace the reectors
into the offshore part of the seismic grid. The Cretaceous to
Paleocene parts of the basin-ll that were not drilled were
interpreted on the basis of published outcrop data (Astorga
et al., 1991; Mende, 2001; Campos, 2001). The thicknesses
of these successions were attached to wells. Well 2 is the
key well for the interpretation of the seismic sections of the
South Limon Basin. The long section 4 connects both
basins and shows that the interpretation is consistent.
The seismic interpretation was used as a base for the 3D
interpretation of the basin ll, and the development of a
static 3D model of the study area. Such static models reect

Fig. 2. Seismic grid of the study area. The dotted rectangle indicates the
position of the seismic sections shown in Figs. 4, 5, 6a and 7a. The dotted
squares indicate the position of the 3D models shown in Figs. 6b and 7b.

the present state of a geologic body and include tectonic


deformation and the rotation of strata (Blendinger et al.,
2004). From the seismic data it was also possible to identify
different deformation phases in the offshore part of the
Limon Basin, which can be correlated with the deformation phases that had previously been recognized in the
onshore area of Costa Rica. A few seismic sections were
depth converted to calculate water depth and the subsurface geometries of sedimentary and structural features.
4. Basin-ll architecture
The basin-ll architecture was reconstructed in 3D for
the areas shown in Fig. 2. The uppermost seismic reector
interpreted was the seabottom (see bathymetric map of the
Limon Basin in Fig. 3). The map shows that both the
northern and the southern sub-basin have a at shelf area
with a mean water depth of 75 m, followed by a 45 km wide
slope. The average slope-angles in the central North Limon
Basin are 2:523 . In the northern portion of the North
Limon Basin, near the Rio San Juan Delta, the slope has
angles of 323:5 . In the South Limon Basin, the slopeangle is 426:5 , and the shelf is broader as compared with
the North Limon Basin.

ARTICLE IN PRESS
C. Brandes et al. / Marine and Petroleum Geology 24 (2007) 277287

280

4.1. North Limon Basin


The North Limon Basin is characterized by partially
continuous depositional units, which extend over several

tens of kilometres on the NWSE directed cross-lines. The


NESW orientated in-lines display sedimentary units with
a pronounced seaward dip (Fig. 4). Maximum sediment
thickness is observed in the vicinity of the present-day
coastline. Large listric NWSE orientated normal faults
characterize the shelf areas (Figs. 4 and 5). They bound a
small graben structure in the northern part of the study
area (Fig. 4). Plio to Pleistocene deposits show an increased
thickness in the graben-ll. All synthetic normal faults sole
into one sub-horizontal detachment. Borehole data indicate that the position of the detachment is probably
controlled by the lithological change from limestone of the
Senosri Formation to shale of the overlying Uscari
Formation. Locally, the synthetic normal faults are
associated with a rollover (Fig. 4). In the very north of
the study area, four seismic lines provide insights into the
Rio San Juan Delta (Fig. 2). These sections display several
large listric normal faults in the shelf and slope area, which
are associated with small basins (Fig. 5). At the base of the
slope some evidence is found for deltaic toe thrusts,
indicating local compression. As their counterparts further
south, the listric normal faults trend NWSE and sole into
a common detachment probably formed by the transition
from the Senosri Formation into overlying Uscari Formation deposits. The normal faults in the shelf area show
activity until Plio-Pleistocene times. The faults in the slope
area seem to have been active in the Late Miocene to Early
Pliocene. The tip lines of the faults are buried under Late
Pliocene deposits, indicating that the activity already
ceased in the Pliocene.
4.2. South Limon Basin

Fig. 3. Bathymetric map of the study area.

The South Limon Basin is dominated by the Limon foldand-thrust belt. The best insight into the architecture of
1this fold-and-thrust belt is provided by the NESW
directed seismic lines. The Limon fold-and-thrust belt is

Fig. 4. Slope parallel seismic section of the North Limon Basin (part of seismic line 6 in Fig. 2). Large listric normal faults of Pleistocene age are located in
the shelf area. These normal faults can be attributed to a Pleistocene to Recent extensional phase.

ARTICLE IN PRESS
C. Brandes et al. / Marine and Petroleum Geology 24 (2007) 277287

281

Fig. 5. Seismic line from the San Juan Delta (seismic line 3 in Fig. 2). The section displays several large listric normal faults in the shelf and slope area.
These faults are associated with small basins. At the base of the slope is some evidence for thrusting, indicating compression. The major listric normal
faults in the shelf and slope area trend NWSE and sole into a common detachment. The normal faults in the shelf area show Plio-Pleistocene activity. The
faults in the slope area were active in the Late Miocene to Early Pliocene. The tip lines are buried under Late Pliocene deposits, indicating that the activity
already ceased in the Pliocene. For the base of the slope two different interpretations are offered because of the limited data base. (a) The section without
interpretation. (b) The structure at the base of the slope are interpreted as toe thrust. (c) Alternative interpretation, with more toe thrusts.

characterized by concentric or asymmetric hangingwall


anticlines and large southwestward dipping listric or planar
thrusts (Fig. 6a). Below the detachment, no signicant
deformation can be observed. As in the North Limon
Basin, the location of the detachment seems to be
controlled by the lithological change from limestone of
the Senosri Formation to shale of the Uscari Formation.
Small and lenticular piggy-back basins are associated with
the fold-and-thrust belt, which contain some of the
youngest sediments of the South Limon Basin (Fig. 6a).
The 3D model of Fig. 6(b) gives a more detailed impression

of the internal architecture of the fold-and-thrust belt and


the orientation of the thrust faults. Furthermore it shows
that the detachment has a very constant depth throughout
the study area. Thrusts located in a more internal position
within the fold-and-thrust belt are generally steeper and
have greater offsets than the more external and probably
younger thrusts. There is no evidence for out-of-sequence
thrusting. Towards the north, the thrusts seem to shift from
a NE direction to a NW direction. An antiformal structure,
referred to as Mo n High, is located between the North and
South Limon Basin (Fig. 7a). A strong reector envelopes

ARTICLE IN PRESS
282

C. Brandes et al. / Marine and Petroleum Geology 24 (2007) 277287

Fig. 6. (a) Seismic line from the Limon fold-and-thrust belt (seismic line 16 in Fig. 2). The section displays four southwestward dipping thrusts. All thrusts
sole into a sub-horizontal detachment (near base Middle Miocene). Behind the rst thrust a deep piggy-back basin developed, lled with Plio-Pleistocene
sedimentary rocks. (b) 3D model of the Limon fold-and-thrust belt. It can be derived that the detachment has a very constant depth. Only towards the
north it is in a slightly higher position. Thrusts located in a more internal position within the fold-and-thrust belt are generally steeper and have greater
offsets than more external and probably younger thrusts.

the Mo n High and separates it from the surrounding and


overlying sedimentary rocks. Below this reector the
structure seems weakly stratied. It is draped with Middle
Miocene and younger deposits, older units display an onlap
pattern. Well 1 was drilled on the northern ank of the
structure and terminates in Middle Eocene carbonates,
which are unconformably overlain by Early Miocene shales
(Fig. 7b). Upper Eocene and Oligocene deposits are
lacking. The wedging of reector packages against the

Mo n High implies that these units are present in the deeper


parts of the basin. A small graben structure characterizes
the crestal area of the Mo n High (Fig. 7b). Based on this
data set the Mo n High is interpreted as an anticline
structure, which evolved during Late Eocene to Oligocene
times, probably as basement-cored anticline. The central
graben is interpreted as crestal collapse. The 3D visualization implies that the Mo n High has an elliptic outline and
trends approx. NorthSouth (Fig. 7c).

ARTICLE IN PRESS
C. Brandes et al. / Marine and Petroleum Geology 24 (2007) 277287

283

5. Deformation phases in the Limon Basin


Structural interpretation of the seismic data implies that
the study area has been affected by several deformation
phases since the Paleocene. A key element in the analysis of
the deformation phases is the Mo n High, which is
interpreted to have developed in response to a rst
compressional deformation event affecting the study area
between the Middle Eocene and the Early Miocene. This
phase caused the uplift of the High as indicated by the lack
of Eocene and Oligocene deposits on the northern ank of
the Mo n High and the onlap of Lower to Middle Miocene
deposits against the structure. A second, but less strong
phase of uplift occurred between the late Middle Miocene
and Pliocene as indicated by the upward bending of seismic
reectors. Since then the Mo n High was draped with
sediments. The reector pattern indicates that the crestal
graben structure might has been still active in the Pliocene
(Fig. 7a). This late activity is probably related to the
activity of the NWSE trending listric normal faults of the
North Limon Basin that indicate a strong Plio-Pleistocene
extensional phase.
For the South Limon Basin, thickness variations across
the thrust faults provide the possibility to infer the age of
the deformation in greater detail. The offshore part of the
South Limon Basin was affected by a compressional phase
since the Pliocene, but most of the deformation occurred
during the Pleistocene (Brandes et al., in press). Topographic breaks on the sea-oor located at the tip of some of
the thrusts are interpreted as fault-scarps, likely indicating
young thrust movements. Another hint for ongoing
deformation in the South Limon Basin is the earthquake
activity. The Limon earthquake in 1991 is attributed to
movements along a blind thrust (Suarez et al., 1995).
6. Discussion

Fig. 7. (a) Seismic section from the Mo n High (part of seismic line 22 in
Fig. 2). (b) The interpretation in a line drawing. The Mo n High is a convex,
mound-like antiformal structure interpreted as a basement high (Barrientos
et al., 1997). A strong reector envelopes the Mo n High, delineating it from
the surrounding sedimentary rocks. Below this reector, the Mo n High is
very weakly layered or completely structureless. Some sections, however,
show a more distinct layered reector pattern especially in the upper part of
the structure. The lack of Oligocene deposits at the western ank of the
structure might indicate vertical movements between Eocene and Miocene
times. (c) 3D reconstruction of the Mo n High. The deformed South Limon
Basin is located to the left of the Mo n High. The undeformed North Limon
Basin is on the right-hand side. The depth of the seismic lines is in two-waytravel time (5 s). The central portion of the Mo n High is in a depth of
1.81.9 s. View is towards the west.

The offshore part of the North Limon Basin shows


several characteristics of passive continental margins. The
wedge-shape geometry of the accumulated sediments, the
seaward-dipping sedimentary units and the large listric
normal faults are well known features of these settings.
Many authors described these characteristics from different
passive margins in the circum Atlantic area (e.g. Hutchinson et al., 1982; Watts and Thorne, 1984; Bond et al.,
1995). The slope-angle in the central part of the North
Limon Basin 2:523 is low compared with the global
average of 5 (Emery, 1965). In the area of the San Juan
Delta, the slope-angle is 323:5 and the delta shows intense
deformation with a series of listric normal faults in the
shelf area and strong extension on the continental slope
area. There is even evidence for toe thrusting at the base of
the slope (Fig. 5). These features are well-known from e.g.
the Niger Delta and other comparable settings (e.g. Doust
and Omatsola, 1989; Bilotti and Shaw, 2005). Passive
continental margins are known as the place of synsedimentary deformation (e.g. Carver, 1968; Prestholm

ARTICLE IN PRESS
284

C. Brandes et al. / Marine and Petroleum Geology 24 (2007) 277287

and Walderhaug, 2000; Broucke et al., 2004). In several


cases this deformation is interpreted to be gravity induced
(e.g. Rowan et al., 2004). The large listric and seaward
dipping normal faults in the central and northern North
Limon Basin are characteristic for extensional deformation
at passive margins. Based on the seismic data it is difcult
to decide whether the extension is related to tectonic forces
sensu stricto or whether it is gravity induced. Gravity
induced deformation is very likely for the San Juan Delta
because of the linked extensional and compressional fault
system.
For Costa Rica, where no salt is in the subsurface and no
major deltas trigger the gravitational tectonics a comparison with the Mexican Ridges seems to be suitable. The
Mexican Ridges are a deep-water fold-and-thrust belt,
which developed in the western Gulf of Mexico. Bufer
et al. (1978) and Trudgill et al. (1999) showed that the
detachment in the Mexican Ridges occurred in a thick
shale interval. This is quite similar to structural style of
NE Costa Rica. Like the Mexican Ridges, the area of the
Rio San Juan Delta shows a linked extensional and
compressional system. The deformation in the Mexican
Ridges is addressed to massive gravity sliding, triggered by
regional uplift and sediment loading (Bufer et al., 1978).
Similar to the Mexican Ridges, the deformation in the
Rio San Juan Delta can be interpreted as a consequence of
sediment loading. In this case the sediment masses
accumulated at the mouth of the Rio San Juan. The
deformation further to the south might be triggered by
thick Pleistocene deposits, but there is no delta present.
Today several rivers transport sediments to this area. Such
a situation is very likely for the Plio-Pleistocene. In the
North Limon Basin, the Plio-Pleistocene sediments close to
the present-day coastline reach a thickness of 1 km. This
appears to be sufcient to trigger gravity sliding, but an
inuence of tectonic forces cannot be ruled out.
Various types of faults develop due to deformation at
passive margins (e.g. Bruce, 1973). The evolution of these
faults is attributed to differential compaction caused by
lithological changes, e.g. from shale to sandstone (Carver,
1968). Deformation at passive margins is also strongly
inuenced by at detachment surfaces especially caused by
evaporitic layers (e.g. Letouzey et al., 1995; Peel et al.,
1995; Vendeville, 2005; Gaullier and Vendeville, 2005).
Gemmer et al. (2004, 2005) simulated the effect of such salt
layers on the tectonic evolution of the sedimentary wedge
of a passive margin. In this model the material moves on
top of the viscous salt layer. This causes extension in updip
regions and compression in downdip areas. The low slopeangle in the North Limon Basin might be a hint to low
basal friction along the detachment. Continuous sediment
progradation leads to the formation of landward extension
and seaward compression (Gemmer et al., 2005). The
seismic section from the San Juan Delta shows comparable
structures (Fig. 5). The rheology at the detachment in the
North Limon Basin is interpreted to be controlled by the
lithological change from limestone to shale, which was

detached from the underlying limestone. This situation is


different to salt tectonics, because the salt itself starts to
ow under applied stress, but the results are quite similar.
In the San Juan Delta the locally higher and more rapid
sediment input associated with a high water content of the
sediments probably intensied the deformation. In the
northern part of the North Limon Basin the passive margin
seems to have reached a later stage with stronger
deformation compared with the central part that lacks
the inuence of a major river.
In contrast, the offshore South Limon Basin is interpreted as a retro-arc foreland basin (Campos, 2001). The
basin is dominated by folding and thrusting characterized
by thin-skinned tectonics on a sub-horizontal detachment
located within the sedimentary rocks. Taking the results of
the seismic interpretation into account, it is very likely that
the internal architecture of the observed folds and thrusts is
largely determined by the friction at the basal detachment.
Experimental studies have shown that low basal friction
along a detachment mainly leads to frontal accretion of
thrust sheets (Gutscher et al., 1996). In contrast, high basal
friction causes increasing underplating of weakly deformed
thrusts and generates a steeper slope angle (Gutscher et al.,
1996). The Limon fold-and-thrust belt consists of repeated
thrust sheets. No evidence for underplating or for duplex
formation can be found on the seismic sections. Therefore,
it is assumed that the basal detachment is controlled by low
friction.
It is well-known that the geometry and position of
structural elements such as faults and thrusts in a
sedimentary basin are largely controlled by the lithology
and initial architecture of the basin-ll. Many studies have
shown the effects of a multi-layered stratigraphy and the
occurrence of weak decollments on the deformation style
(e.g. Turrini et al., 2001; Costa and Vendeville, 2002).
Verges et al. (1992), Gutscher et al. (1996) and Ford (2004)
described in great detail the impact of low friction along
the basal detachment on the geometry of fold-and-thrust
belts. The North and South Limon Basins provide the
possibility to study these lithology controlled effects in two
different tectonic regimes. The seismic data imply that the
location and geometry of the large-scale structural elements
in both basins are independently controlled by the preexisting stratigraphy and lithological pattern.
Flat and laterally continuous depositional units can
favour the development of a horizontal detachment. In
addition strong rheologic contrasts between the individual
depositional units support the formation of such a
detachment surface. For the Limon Basin seismic lines
and well data imply that a lithologic contrast between
limestone and shale led to the formation of a basal
detachment in the South Limon Basin and probably in the
southern part of the North Limon Basin.
The evolution of the structural elements of the Limon
back-arc basin reects several deformation phases.
Seyfried et al. (1991) and Krawinkel et al. (2000) gave
a comprehensive overview of the deformation phases

ARTICLE IN PRESS
C. Brandes et al. / Marine and Petroleum Geology 24 (2007) 277287

observed in southern Central America, identifying three


main phases of deformation and uplift of the island-arc in
Early Campanian, Oligocene and Miocene times. Campos
(2001) observed three unconformities in the onshore part of
the South Limon Basin in Eocene, Oligocene and MioPliocene times. The seismic interpretation of this study
supports the observations of these authors in so far as the
rst compressional phase probably occurred between the
Middle Eocene and Late Miocene. This phase caused the
onset of uplift of the Mo n High and correlates very well
with the data of Seyfried et al. (1991), Amann (1993),
Bottazzi et al. (1994), Mende (2001) and Campos (2001).
The Mo n High has an elliptic outline and trends roughly
NS. This implies an EW directed compression. Barrientos et al. (1997) described a main convergence of North and
South America in the Late EoceneEarly Oligocene, which
is interpreted to have caused a narrowing of the Caribbean
Plate. The lack of Oligocene deposits on the northern ank
of the Mo n High can be interpreted as response to this
convergence phase by folding and uplift. In addition, the
seismic data from the Mo n area show a second phase of
uplift for the Late Middle Miocene and Late Miocene
(Fig. 7a). This observation correlates very well with
the observations of Amann (1993) and Mende (2001).
Krawinkel et al. (2000) interpreted this deformation phase
as a consequence of increased plate coupling between the
Cocos Plate and the Caribbean Plate. Furthermore the
seismic interpretation of Fig. 4 and Fig. 5 shows evidence
of a Pleistocene extensional regime in the North Limon
Basin. This observation corresponds with the interpretation of Mende (2001), who expected ongoing subsidence in
the North Limon Basin.
The Miocene deformation described from the onshore
record (Bottazzi et al., 1994; Campos, 2001) might have
also been the consequence of a successively decreasing
subduction angle. De Boer et al. (1995) proposed a
lowering of the subduction angle of the Cocos Plate since
the Miocene in combination with a thickening of the upper
plate. However, the offshore part of the South Limon
Basin was affected by a compressional regime only since
the Pliocene, but most of the deformation occurred since
the Pleistocene (Brandes et al., in press). Active deformation is indicated by fault-scarps on the sea oor. The
subduction of the Cocos Ridge started 3.6 Myr ago. At ca.
1.6 Ma the subducted ridge should have reached the study
area (Collins et al., 1995). The Pliocene to recent
deformation in the offshore part of the South Limon
Basin, which is observed on the seismic lines can therefore
be attributed to the shallow subduction of the Cocos Ridge
and the related back-arc compression.
In Costa Rica stress conditions change from a low-stress
subduction zone associated with an extensional back-arc
basin in the north to a high stress subduction zone with a
deformed retro-arc area in the south over a very short
distance. The main driving mechanism for this deformation
pattern seems to be the very low subduction angle in southern
Costa Rica, which is a direct effect of the young age of the

285

oceanic lithosphere (Norabuena et al., 2004). In contrast in


northern Costa Rica, the Cocos Plate is subducted with 60 .
The much steeper subduction in northern Costa Rica is
associated with the development of an extensional back-arc
basin. The Trans Isthmic Fault System separates the two
different island-arc segments. Along this strike-slip fault
zone, the extensional northern arc segment is decoupled from
the compressional southern arc segment.
7. Conclusions
The unique geological setting of the Limon Basin with
the two different sub-basins allows to study the inuence of
the basin-ll architecture on the deformation style of
sedimentary basins in great detail. Despite the back-arc
position, the offshore part of the North Limon Basin shows
characteristics of passive continental margins with a
wedge-shaped geometry, seaward propagating depositional
units and large listric normal faults. The area of the Rio
San Juan Delta shows stronger, probably gravity induced
deformation with a series of listric normal faults in the shelf
and slope area. At the base of the slope the seismic data
give evidence for compressional deformation with deltaic
toe thrusts. The data imply that the intensity of deformation at the passive margin is controlled by the amount of
sediment input. Locally higher input leads to increased
loading and can induce slope failure.
The South Limon Basin is dominated by the thinskinned Limon fold-and-thrust belt, which is characterized
by concentric or asymmetric hangingwall anticlines and
large listric or planar southwestward dipping thrusts. The
offshore part of this fold-and-thrust belt developed due to
back-arc compression since the Pliocene and the structural
style of this fold-belt is largely determined by the position
and friction at the basal detachment. The Plio-Pleistocene
evolution of both sub-basins drastically diverges. The
North Limon Basin is characterized by extension, whereas
the South Limon Basin is dominated by compression.
Seismic lines and well data give evidence for a connection
of basin-ll architecture and the geometry and position of
structural elements in the North and South Limon Basin.
Probably the laterally continuous depositional units and
the rheologic contrasts between limestone and shale are
important controlling factors for the spatial development
of these elements. Especially the occurrence of at
detachment surfaces in the same stratigraphic level in both
tectonic regimes underlines the possible inuence of the
initial basin-ll architecture. This detachment largely
controls the deformation, independently from the tectonic
regime. Costa Rica can be subdivided into two arc
segments, which are separated by the Trans Isthmic Fault
System. The northern arc segment shows a steep low-stress
subduction zone associated with an extensional back-arc
basin. In the south there is a high-stress subduction zone
with a deformed retro-arc area. Along the Trans Isthmic
Fault System, the extensional northern arc segment is
decoupled from the compressional southern arc segment.

ARTICLE IN PRESS
286

C. Brandes et al. / Marine and Petroleum Geology 24 (2007) 277287

Acknowledgements
We would like to thank the Costa Rican Ministry of
Environment and Energy (MINAE) for providing the data
base. We are indebted especially to Alvaro Aguilar and
Gustavo Segura for logistic help. Financial support from
the German Research Foundation (DFG) Project Wi 1844/
6-1 and a graduate scholarship from the University of
Hannover are gratefully acknowledged. Seismic Micro
Technologies is gratefully thanked for the sponsoring of
Kingdom Suiter. Many thanks to Peter Blisniuk for
correcting an early draft of the manuscript. Careful reviews
by Bert Bally and David Roberts helped to improve the
manuscript. Ulrich Asprion, Franz Binot, Lolita Campos,
Frederic Flerit, Christoph Gaedicke, Stefan Ladage,
Andreas Mende, Klaus Reicherter and Imke StruX are
gratefully acknowledged for constructive discussions.
References
Abratis, M., Worner, G., 2001. Ridge collision, slab-window formation,
and the ux of Pacic asthenosphere into the Caribbean realm.
Geology 29, 127130.
Amann, H., 1993. Randmarine und terrestrische Ablagerungsraume des
neogenen Inselbogen-systems in Costa Rica (Mittelamerika). Prol,
vol. 4, 161pp.
Astorga, A., 1997. El puente-istmo de America Central y la evolucion de la
Placa caribe (con efasis en el Mesozoico). Prol, vol. 12, 201pp.
Astorga, A., 1988. Geodinamica de las cuencas del Cretacico Superior
Paleogeno de la region forearc del Sur de Nicaragua y Norte de
Costa Rica. Revista Geologica de America Central 9, 140.
Astorga, A., Fernandez, J.A., Barboza, G., Campos, L., Obando, J.,
Aguilar, A., Obando, L.G., 1991. Cuencas sedimentarias de Costa
Rica: evolucion geodinamica y potential de hidrocarburos. Revista
Geologica de America Central 13, 2559.
Barboza, G., Fernandez, A., Barrientos, J., Bottazzi, G., 1997. Costa Rica:
petroleum geology of the Caribbean margin. Leading Edge 16,
17871794.
Barrientos, J., Bottazzi, G., Fernandez, A., Barboza, G., 1997. Costa
Rican data synthesis indicates oil, gas potential. Oil and Gas Journal,
May 12, 7680.
Bilotti, F., Shaw, J.H., 2005. Deep-water Niger delta fold and thrust belt
modelled as a critical-taper wedge: the inuence of elevated basal uid
pressure on structural styles. AAPG Bulletin 89, 14751491.
Blendinger, W., Brack, P., Norborg, A.K., Wulf-Pedersen, E., 2004.
Three-dimensional modelling of an isolated carbonate buildup
(Triassic, Dolomites, Italy). Sedimentology 51, 297314.
Bond, G.C., Kominz, M.A., Sheridan, R.E., 1995. Continental terraces
and rises. In: Busby, C.J., Ingersoll, R.V. (Eds.), Tectonics of
Sedimentary Basins. Blackwell Science, Oxford, pp. 149178.
Bottazzi, G., Fernandez, A., Barboza, G., 1994. Sedimentolog a e historia
tectono-sedimentaria de la cuenca Limon Sur. In: Seyfried, H.,
Hellmann, W. (Eds.), Geology of an Evolving Island Arc, The Isthmus
of Southern Nicaragua, Costa Rica and Western Panama. Prol, vol.
7, pp. 351389.
Brandes, C., Astorga, A., Blisniuk, P., Littke, R., Winsemann, J., in press.
Anatomy of anticlines, piggy-back basins and growth strata: a case
study from the Limon fold-and-thrust belt, Costa Rica, IAS Special
Publication.
Broucke, O., Temple, F., Rouby, D., Robin, C., Calassou, S., Naplas, T.,
Guillocheaun, F., 2004. The role of deformation processes on the
geometry of mud-dominated turbiditic systems, Oligocene and LowerMiddle Miocene of the Lower Congo Basin (West African margin).
Marine and Petroleum Geology 21, 327348.

Bruce, C.H., 1973. Pressured shale and related sediment deformation:


mechanism for development of regional contemporaneous faults.
AAPG Bulletin 57, 878886.
Bufer, R.T., Shaub, F.J., Watkins, J.S., Worzel, J.L., 1978. Anatomy of
the Mexican Ridges, Southwestern Gulf of Mexico. In: Watkins, J.S.,
Montadert, L., Dickerson, P.W. (Eds.), Geological and Geophysical
Investigations of Continental Margins. AAPG Memoir, vol. 29,
pp. 319327.
Burke, K., Fox, P.J., Sengor, A.M.C., 1978. Buoyant ocean oor and the
evolution of the Caribbean. Journal of Geophysical Research 83,
39493954.
Calvo, C., 2003. Provenance of plutonic detritus in cover sandstones of
Nicoya Complex, Costa Rica: cretaceous unroong history of a
Mesozoic ophiolite sequence. GSA Bulletin 115, 832844.
Campos, L., 2001. Geology and basins history of middle Costa Rica: an
intraoceanic island arc in the convergence between the Caribbean and
the central pacic plates. Tubinger Geowissenschaftliche Arbeiten,
Reihe A, B and 62, 138pp.
Carver, R.E., 1968. Differential compaction as a cause of regional
contemporaneous faults. AAPG Bulletin 52, 414419.
Coates, A.G., Jackson, J.B.C., Collins, L.S., Cronin, T.M., Dowsett, H.J.,
Bybell, L.M., Jung, P., Obando, J.A., 1992. Closure of the Isthmus of
Panama: the near-shore marine record of Costa Rica and western
Panama. GSA Bulletin 104, 814828.
Coates, A.G., Aubry, M-P., Berggren, W.A., Collins, L.S., Kunk, M.,
2003. Early Neogene history of the Central American arc from Bocas
del Toro, western Panama. GSA Bulletin 115, 271287.
Collins, L.S., Coates, A.G., Jackson, J.B.C., Obando, J.A., 1995. Timing
and rates of emergence of the Limon and Bocas del Toro basins:
Caribbean effects of Cocos Ridge subduction? In: Mann, P. (Ed.),
Geologic and Tectonic Development of the Caribbean Plate Boundary
in Southern Central America. GSA Special Publication, vol. 295,
pp. 263289.
Costa, E., Vendeville, B.C., 2002. Experimental insights on the geometry
and kinematics of fold-and-thrust belts above weak, viscous evaporitic
decollement. Journal of Structural Geology 24, 17291739.
De Boer, J.Z., Drummond, M.S., Bordelon, M.S., Defant, M.J., Bellon,
H., Maury, R.C., 1995. Cenozoic magmatic phases of the Costa Rican
island arc (Cordillera de Talamanca). In: Mann, P. (Ed.), Geologic and
Tectonic Development of the Caribbean Plate Boundary in Southern
Central America. GSA Special Publication, vol. 295, pp. 3555.
Donnelly, T.W., 1994. The Caribbean Cretaceous basalt association: a
vast igneous province that includes the Nicoya Complex of Costa Rica.
In: Seyfried, H., Hellmann, W. (Eds.), Geology of an Evolving Island
Arc, The Isthmus of Southern Nicaragua, Costa Rica and Western
Panama. Prol, vol. 7, pp. 1745.
Doust, H., Omatsola, E., 1989. Niger Delta. In: Edwards, S.D.,
Santograssi, P.A. (Eds.), Divergent/Passive Margin Basins. AAPG
Memoir, vol. 48, pp. 201238.
Emery, K.O., 1965. Characteristics of continental shelves and slopes.
AAPG Bulletin 49, 13791384.
Escalante, G., Astorga, A., 1994. Geolog a del este de Costa Rica y el
norte de Panama. Revista Geologica de America Central, Vol. esp.
Terremoto de Limon, 114.
Fernandez, J.A., Bottazzi, G., Barboza, G., Astorga, A., 1994. Tectonica y
estratigraa de la Cuenca Limon Sur. Revista Geologica de America
Central, Vol. esp. Terremoto de Limon, 1528.
Fernandez, J., Alvaro, A., Guillermo, B., Bottazzi, G., Campos, L.,
Obando, J., Tejera, R., Arrieta, L., Barrientos, J., Bustos, I., Escalante,
G., Pizarro, D., Valer n, E., Astorga, A., Bolanos, X., Calvo, C.,
Laurito, C., Rojas, J., Valerio, A., 1997. Mapa Geologico de Costa
Rica. Ministerio del Ambiente y Energ a, Costa Rica.
Ford, M., 2004. Depositional wedge tops: interaction between low basal
friction external orogenic wedges and exural foreland basins. Basin
Research 16, 361375.
Frisch, W., Meschede, M., Sick, M., 1992. Origin of the Central American
ophiolites: evidence from paleomagnetic results. GSA Bulletin 104,
13011314.

ARTICLE IN PRESS
C. Brandes et al. / Marine and Petroleum Geology 24 (2007) 277287
Gaullier, V., Vendeville, B.C., 2005. Salt tectonics driven by sediment
propagation: part IIradial spreading of sedimentary lobes prograding above salt. AAPG Bulletin 89, 10811089.
Gemmer, L., Ings, S.J., Medvedev, S., Beaumont, C., 2004. Salt tectonics
driven by differential sediment loading: stability analysis and niteelement experiments. Basin Research 16, 199218.
Gemmer, L., Beaumont, C., Ings, S.J., 2005. Dynamic modelling of
passive margin salt tectonics: effects of water loading, sediment
properties and sedimentation patterns. Basin Research 17, 383402.
Grafe, K., Frisch, W., Villa, I.M., Meschede, M., 2002. Geodynamic
evolution of southern Costa Rica related to low-angle subduction of
the Cocos Ridge: constraints from thermo-chronology. Tectonophysics 348, 187204.
Gutscher, M.-A., Kukowski, N., Malavieille, J., Lallemand, S., 1996.
Cyclic behaviour of thrust wedges: insights from high basal friction
sandbox experiments. Geology 24, 135138.
Hutchinson, D.R., Grow, J.A., Klitgord, K.D., Swift, B.A., 1982. In:
Deep Structure and Evolution of the Carolina Trough. AAPG
Memoir, vol. 34, pp. 129152.
Krawinkel, J.J., 2003. Struktur und Kinematik am konvergenten
Plattenrand der sudlichen Zentralamerikanischen Landbrucke
(Zentral- und Sud-Costa Rica, West-Panama). Prol, vol. 20, 36pp.
Krawinkel, J., Seyfried, H., 1994. A review of plate-tectonic processes
involved in the formation of the southwestern edge of the Caribbean
Plate. In: Seyfried, H., Hellmann, W. (Eds.), Geology of an Evolving
Island Arc, The Isthmus of Southern Nicaragua, Costa Rica and
Western Panama. Prol, vol. 7, pp. 4761.
Krawinkel, H., Seyfried, H., Calvo, C., Astorga, A., 2000. Origin and
inversion of sedimentary basins in southern Central America.
Zeitschrift fur Angewandte Geologie SH 1, 7177.
Letouzey, J., Collett, B., Vially, R., Chermette, J.C., 1995. Evolution of
salt-related structures in compressional settings. In: Jackson, M.P.A.,
Roberts, D.G., Snelson, S. (Eds.), Salt Tectonics: A Global
Perspective. AAPG Memoir, vol. 65, pp. 4160.
Lundberg, N., 1991. Detrital record of the early Central American
magmatic arc: petrography of intraoceanic forearc sandstones, Nicoya
Peninsula, Costa Rica. GSA Bulletin 103, 905915.
Maresch, W.V., Stockhart, B., Baumann, A., Kaiser, C., Kluge, R.,
Kruckhans-Lueder, G., Brix, M.R., Thomson, S., 2000. Crustal
history and plate tectonic development in the southern Caribbean.
Zeitschrift fur Angewandte Geologie SH 1, 283290.
McNeill, D.F., Coates, A.G., Budd, A.F., Borne, P.F., 2000. Integrated
paleontologic and paleomagnetic stratigraphy of the Upper Neogene
deposits around Limon, Costa Rica: a coastal emergence record of the
Central American Isthmus. GSA Bulletin 112, 963981.
Mende, A., 2001. Sedimente und Architektur der Forearc- und Backarc-Becken
von Sudost-Costa Rica und Nordwest-Panama. Prol, vol. 19, 130pp.
Meschede, M., Frisch, W., Chinchilla Chavez, A.L., Lopez Saborio, A.,
Calvo, C., 2000. The plate tectonic evolution of the Caribbean plate in
the Mesozoic and early Cenozoic. Zeitschrift fur Angewandte Geologie
SH 1, 275281.
Norabuena, E., Dixon, T.H., Schwartz, S., DeShon, H., Newman, A.,
Protti, M., Gonzales, V., LeRoy, D., Fueh, E.R., Lundgren, P., Pollitz,
F., Sampson, D., 2004. Geodetic and seismic constraints on some
seismogenic zone processes in Costa Rica. Journal of Geophysical
Research 109, B11403, doi:10.1029/2003JB002931.
Peel, F.J., Travis, C.J., Hossack, J.R., 1995. Genetic structural provinces and
salt tectonics of the Cenozoic offshore US Gulf of Mexico: a preliminary
analysis. In: Jackson, M.P.A., Roberts, D.G., Snelson, S. (Eds.), Salt
Tectonics: A Global Perspective. AAPG Memoir, vol. 65, pp. 153175.
Pindell, J.L., Cande, S.C., Pitman, W.C., Rowley, D.B., Dewey, J.F.,
Labrecque, J., Haxby, W., 1988. A plate-kinematic framework for
models of Caribbean evolution. Tectonophysics 155, 121138.

287

Prestholm, E., Walderhaug, O., 2000. Synsedimentary faulting in a


Mesozoic deltaic sequence, Svalbard, arctic Norway-Fault geometries,
faulting mechanisms, and sealing properties. AAPG Bulletin 84,
505522.
Ross, M.I., Scotese, C.R., 1988. A hierachical tectonic model of the
Gulf of Mexico and the Caribbean region. Tectonophysics 155,
139168.
Rowan, M.G., Peel, F.J., Vendeville, B.C., 2004. Gravity-driven fold belts
on passive margins. In: McClay, K.R. (Ed.), Thrust Tectonics and
Hydrocarbon Systems. AAPG Memoir, vol. 82, pp. 157182.
Seyfried, H., Astorga, A., Amann, H., Calvo, C., Kolb, W., Schmidt, H.,
Winsemann, J., 1991. Anatomy of an evolving Island Arc: tectonic and
eustatic control in the south Central American forearc area. In:
MacDonald D.I.M. (Ed.), Sea-level Changes at Active Plate
Margins: Processes and Products. IAS Special Publication, vol. 12,
pp. 273292.
Sheehan, C.A., Peneld, G.T., Morales, E., 1990. Costa Rica geologic
basins lure wildcatters. Oil and Gas Journal, April 30, 7479.
Suarez, G., Pardo, M., Dom nguez, J., Ponce, L., Montero, W., Boschini,
I., Rojas, W., 1995. The Limon, Costa Rica earthquake of April 22,
1991: back arc thrusting and collisional tectonics in a subduction
environment. Tectonics 14, 518530.
Trudgill, B.D., Rowan, M.G., Fiduk, J.C., Weimer, P., Gale, P.E., Korn,
B.E., Phair, R.L., Gafford, W.T., Roberts, G.R., Dobbs, S.W., 1999.
The Perdido fold belt, Northwestern Deep Gulf of Mexico, Part 1:
structural geometry, evolution and regional implications. AAPG
Bulletin 83, 88113.
Turrini, C., Ravaglia, A.S., Perotti, C.R., 2001. Compressional structures
in a multilayered mechanical stratigraphy: insights from sandbox
modelling with three-dimensional variations in basal geometry and
friction. In: Koyi, H.A., Mancktelow, N.S. (Eds.), Tectonic Modeling:
A Volume in Honor of Hans Ramberg. GSA Special Publication,
vol. 193, pp. 153178.
Vendeville, B.C., 2005. Salt tectonics driven by sediment propagation: part
Imechanics and kinematics. AAPG Bulletin 89, 10711079.
Verges, J., Munoz, J.A., Mart nez, A., 1992. South Pyrenean fold and
thrust belt: the role of foreland evaporitic levels in thrust geometry. In:
McClay, K.R. (Ed.), Thrust Tectonics. Chapman & Hall, London,
pp. 255263.
von Huene, R., Fluh, E., 1994. A review of marine geophysical studies
along the Middle America Trench off Costa Rica and the problematic
seaward terminus of continental crust. In: Seyfried, H., Hellmannm,
W. (Eds.), Geology of an Evolving Island Arc, The Isthmus of
Southern Nicaragua, Costa Rica and Western Panama. Prol, vol. 7,
pp. 143159.
Watts, A.B., Thorne, J., 1984. Tectonics, global changes in sea-level and
their relationship to stratigraphic sequences at the U.S. Atlantic
continental margin. Marine and Petroleum Geology 1, 319339.
Weinberg, R.F., 1992. Neotectonic development of western Nicaragua.
Tectonics 11, 10101017.
Weyl, R., 1980. Geology of Central America. Borntraeger, Berlin, 371pp.
Werner, R., Hoernle, K., van den Bogaard, P., Ranero, C., von Huene, R.,
1999. Drowned 14-m.y.-old Galapagos archipelago off the coast of
Costa Rica: implications for tectonic and evolutionary models.
Geology 27, 499502.
Winsemann, J., 1992. Tiefwasser-Sedimentationsprozesse und -produkte
in den Forearc-Becken des mittelamerikanischen Inselbogensystems:
eine sequenz-stratigraphische Analyse. Prol, vol. 2, 218pp.
Winsemann, J., Seyfried, H., 1991. Response of deep-water forearc
systems to sea-level changes, tectonic activity and volcaniclastic input
in central America. In: MacDonald D.I.M. (Ed.), Sea-level Changes at
Active Plate Margins: Processes and Products. IAS Special Publication, vol. 12, pp. 217240.

You might also like