You are on page 1of 21

Earth and Planetary Science Letters 260 (2007) 444 464

www.elsevier.com/locate/epsl

Segmentation of the Nazca and South American plates along the


Ecuador subduction zone from wide angle seismic profiles
Audrey Gailler a,, Philippe Charvis a , Ernst R. Flueh b
a
Gosciences Azur, Unit mixte de recherche Universit de Nice Sophia Antipolis, Institut de Recherche pour le Dveloppement (IRD),
Centre National de la Recherche Scientifique (CNRS), Universit Pierre et Marie Curie, F-06235, BP 48 Villefranche-sur-mer, France
b
IFM-GEOMAR, Research Center for Marine Sciences, 24148, Kiel, Germany
Received 10 October 2006; received in revised form 23 May 2007; accepted 23 May 2007
Available online 2 June 2007
Editor: R.D. van der Hilst

Abstract

We describe the deep structure of the south Colombiannorthern Ecuador convergent margin using travel time inversion of wide-
angle seismic data recently collected offshore. The margin appears segmented into three contrasting zones. In the North Zone, affected
by four great subduction earthquakes during the 20th century, normal oceanic crust subducts beneath the oceanic Cretaceous
substratum of the margin underlined by seismic velocities as high as 6.06.5 km/s. In the Central Zone the subducting oceanic crust is
over-thickened beneath the Carnegie Ridge. A steeper slope and a well-developed, high velocity, Cretaceous oceanic basement
characterizes the margin wedge. This area coincides with a gap in significant subduction earthquake activity. In the South Zone, the
subducting oceanic crust is normal. The fore-arc is characterized by large sedimentary basins suggesting significant subsidence.
Velocities in the margin wedge are significantly lower and denote a different nature or a higher degree of fracturing.
Even if the distance between the three profiles exceeds 150 km, the structural segmentation obtained along the Ecuadorian
margin correlates well with the distribution of seismic activity and the neotectonic zonation.
2007 Published by Elsevier B.V.

Keywords: wide-angle seismic data; traveltime tomography; Ecuadorian margin; subduction zone; segmentation of active margin; subduction
earthquakes

1. Introduction and tectonic setting at a rate of 58-mm/yr (Trenkamp et al., 2002). The
margin undergoes shortening perpendicular to the trench
The Ecuador margin (3.0N3.5S) represents a axis, involving the north Andean block expulsion at
region of intense crustal deformation, related to the 6 mm/y (Trenkamp et al., 2002) to the Northeast,
subduction of the oceanic Nazca plate beneath the South- parallel to the margin (Fig. 1).
American continental plate, along a N90E direction and The oceanic Nazca plate in this area has an age to
less than 26 My (Lonsdale, 2005), and carries the
Carnegie Ridge (CR), which testifies to the interaction
between the Galapagos hotspot (GHS) and the Cocos
Corresponding author. Nazca spreading centre since 20 My (e.g., Sallares
E-mail address: gailler@geoazur.obs-vlfr.fr (A. Gailler). and Charvis, 2003). Its subduction beneath the
0012-821X/$ - see front matter 2007 Published by Elsevier B.V.
doi:10.1016/j.epsl.2007.05.045
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464 445

Fig. 1. General map of the eastern part of the Nazca plate and of the southern Colombia and Ecuador margin. Seafloor topography is derived from
satellite altimetry and ship depth soundings (Smith and Sandwell, 1997) completed by available swath mapping (Flueh et al., 2001; Collot et al.,
2002). It is contoured at 500 m intervals between 5000 and 500 m, and 25 m intervals between 250 and the coastline. Heavy solid black lines
show locations of the three wide-angle seismic profiles used in this work. White circles: locations of OBSs along each profile. Bold circles: same as
white circles for sections shown in Fig. 25. White triangles: land station locations. White line is wide-angle seismic refraction line published in
(Sallares et al., 2003). Stars: epicentres of the six great subduction earthquakes that occurred along the margin since 1901, including four in the
northern part (3.5N0.5S). The most important (1906, Mw = 8.8) (Kanamori and McNally, 1982) has partially been reactivated three times from
south to north: 1942 (Mw = 7.9), 1958 (Mw = 7.8) and 1979 (Mw = 8.2) (Beck and Ruff, 1984; Mendoza and Dewey, 1984). In contrast, no major
earthquakes have been reported in the southern part of the study area (0.5S3S), except two poorly documented events in 1901 and 1953. Dashed
contours: rupture zone of the 1958 and 1979 earthquakes (Collot et al., 2004). Bold arrow: relative convergence between Nazca Plate and North
Andean Block from GPS data. DGM: DoloresGuayaquil Megashear. Location of Cayo and Pion formation outcrops and their equivalent in
Colombia (Diabasico formation) are in red and orange (after the geologic maps of Ecuador and Colombia).
446 A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464

Ecuadorian coast induces lateral variations of seismicity, channel seismic data (MCS) that enabled us to image the
arc magmatism, deformations, vertical motion and structure of both plates as well as the geometry and
sediment distribution along the margin (Gutscher et properties of the inter-plate zone.
al., 1999; Collot et al., 2002). The Nazca plate is also This paper, based on the tomographic inversion of 3
marked by huge scarps such as the Yaquina Graben, wide-angle seismic profiles shot across the margin
interpreted as a transform fault associated with the (Fig. 1), provides the first overview of the structure of
remnant East Malpelo spreading axis (Hey, 1977; the crust and upper mantle of the Nazca and South Ameri-
Lonsdale and Klitgord, 1978; Lonsdale, 2005) (Fig. 1). can plates at the South ColombiaEcuador subduction
The Ecuador margin consists of the oceanic North- zone. We will then tentatively correlate long wavelength
Andean block which was accreted to the Brazilian structural variations observed along the margin with sub-
continental craton during the Cretaceous (Jaillard et al., duction processes and large subduction earthquakes.
1997). These oceanic terranes forming the Ecuadorian
margin substratum seem to extend southward into 2. Data acquisition
northern Peru (i.e., Banco Peru; Fig. 1) (Shepherd and
Moberly, 1981). They consist of tholeitic/andesitic During the SISTEUR cruise, wide-angle seismic data
basalts and pillow lavas associated with siliceous were acquired using a seismic source composed of an
sediments, defined as the Pion Formation. In Ecuador, array of eight 16-l airguns, shot at 125 m intervals. During
the Pion Formation is overlain by an intra-oceanic the SALIERI cruise, the seismic source was composed of
volcanic arcs series, such as the Cayo formation (e.g., up to three 32-l airguns; in this case, the shot interval was
Jaillard et al., 1995; Reynaud et al., 1999). To the north set at a time of 60 s leading to a 140 m shot spacing. 24
(south Colombia, Gorgona Island), outcrops of Creta- UTIG-IRD OBS (Nakamura and Garmany, 1991) and up
ceous oceanic terranes (mafic to ultramafic rocks) are to 10 land-seismometers were used during the SISTEUR
associated with the Diabasico Formation (Goosens cruise, whereas 26 UTIG-IRD OBS and 24 GEOMAR
et al., 1977; Marriner and Millward, 1984; Mamberti OBH and OBS (Flueh et al., 2001) were used during the
et al., 2004). SALIERI cruise.
Several studies on deformation of Andean coastal Most of the wide-angle data were of good quality,
basins (e.g. Deniaud, 2000) have shown that the margin showing clear intracrustal refracted phases (Pg), deeper
has undergone a generally compressive regime since arrivals refracted in the upper mantle (Pn), and
accretion of the oceanic terranes to the Ecuadorian reflections at the boundary between the subducting
paleo-margin ( 90 My ago). This compressive regime oceanic crust and the mantle (PmP) (Figs. 24, Tables 1,
appears largely disturbed in the Gulf of Guayaquil (GG) 2, 3). Pg here is used as a generic term to indicate all the
area, where the tectonic structures testify to an NS first arrivals that are not refracted in the mantle
trending opening of the GG basin. The development of (sediments and basement).
this deepest sedimentary basin of the Ecuadorian fore- Significant variations were observed between record
arc during the last 5 Ma is believed to have resulted from sections for instruments located in deep water, shallow-
the escape of the north Andean block to the NE (Fig. 1) water and on-land instruments. This is probably related
(Witt et al., 2006). to the strong change in the bathymetry (more than
Recent coastal uplift facing CR is indicated by 3000 m) across the margin but also to the strong lateral
Pliocene marine terraces (Tablazos) exposed at 200 variation in the crustal structure along each profile.
300-m elevations (De Vries, 1988; Pedoja et al., 2006).
To the north, the morphology of the margin, the 2.1. Record sections for OBS located seaward from the
diversion of drainage networks and increases in coastal trench
accretion support an active subsidence in relation to the
seismic activity of the margin (Dumont et al., 2006). To OBS located on top of the Nazca plate, seaward from
the south, the GG area appears also as a globally the trench, show Pg phases with apparent velocities
subsiding zone (e.g. Witt et al., 2006). ranging from 2.5 to 6.0 km/s. South of the CR, Pg
Our group conducted two major marine seismic phases rapidly reach apparent velocities higher than
experiments along the south ColombiaEcuador con- 5 km/s at near offset (Fig. 4, OBH 45 and Fig. 3, OBS
vergent margin: the SISTEUR cruise in 2000 (Collot 04). In contrast, the apparent velocity of Pg phases is
et al., 2002), and the SALIERI cruise in 2001 (Flueh less than 4 km/s north of the CR (OBH 116, Fig. 2).
et al., 2001). These surveys provided a unique set of At larger offsets, apparent velocities of first arrivals
high-quality wide-angle reflection/refraction and multi- generally exceed 7.07.5 km/s on most record sections.
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464
Fig. 2. Examples of wide-angle seismic data, picked and calculated travel times and corresponding ray paths for the best fitting velocity model (Fig. 5a) along profile SAL-6 (North Ecuador) for OBH
116 (left) and OBS 121 (right). (top) Seismic record section for vertical geophone after data preprocessing (deconvolution, 515 Hz Butterworth filter, equalization); (middle) picked (solid circles with
error bars) and predicted (open circles) travel times for first arrival phases (Pg and Pn) and refraction from the Moho (PmP); (bottom) Corresponding ray tracing for all identified seismic phases. The
black arrows point at the trench location.

447
448
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464
Fig. 3. Same as Fig. 2 for profile SIS-4 (Carnegie Ridge), OBS 04 (left) and OBS 10 (right) for the best fitting velocity model (Fig. 5b).
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464
Fig. 4. Same as Fig. 2 for profile SAL-2 (Gulf of Guayaquil), OBH 45 (left) and OBS 57 (right) except hydrophone component is shown for OBH 45. Best fitting velocity model (Fig. 5c).

449
450 A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464

This has been interpreted to correspond to refraction in floating reflector (not associated with a change in
the upper mantle (Pn), except westward of OBS 04 velocity in the model). In this study, we choose to invert
(Fig. 3) located on top of the CR along profile SIS-4 the reflector corresponding to the crustmantle bound-
where the Pn phase is lacking. In this section (Fig. 3) the ary (PmP phase). Detailed descriptions of this method as
apparent velocity, up to 40 km offset, is 7 km/s and well as parameters used for the regularization of the
probably corresponds to refraction within the subduct- inversion can be found in (Korenaga et al., 2000).
ing oceanic crust. Good quality PmP arrivals are also Using this inversion procedure means that we have to
observed. take into account three major issues:
The first issue is due to the inherent non-uniqueness
2.2. Record sections for OBS located landward from the of the inversion solution. The final tomographic model
trench depends on the starting velocity model chosen for the
inversion, on the method of linearization and on the
For OBS located on the continental slope and shelf geometry of the experiment. The second issue concerns
Pg arrivals are characterized by low apparent velocities the estimate of uncertainties in the final velocity model,
(from 2 to 4 km/s) up to 50 km along profile SAL-2 i.e. how well-resolved are the different parts of the
(OBS 57, Fig. 4). A shadow zone is observable on model. This is done by estimating the degree of
many instruments located along profiles SAL-6 and dependence of the obtained solution on the initial
SIS-4 (Figs. 2 and 3; OBS 121 and 10 respectively). In starting model, as well as to estimate uncertainties, via
both cases, this shadow zone occurs between Pg Monte-Carlo analysis. The method involves a number of
phases with apparent velocities of 4 km/s and inversions with a variety of initial models using travel
6 km/s. Along profile SAL-6, the shadow zone is time picks with random errors applied. If all the Monte-
0.5 s thick. Carlo realizations have the same probability and if the
At larger offsets (N5060 km), Pn arrivals are initial models cover the full region of non-null
identified as first arrivals on most record sections probability within the space of parameters, the standard
along the three profiles. PmP arrivals are observed to be deviation of the obtained solutions can be interpreted as
variable in quality (e.g., SAL-6; Fig. 2, OBS 121). a measure of the final model parameters uncertainty. The
third issue concerns evaluating the trade-off between
2.3. Travel time picking depth and velocity parameters linked to the reflected
phase travel time inversion and depending on the
Travel times of refracted (Pg, Pn) and reflected source-receiver geometry. A conventional method to
(PmP) phases identified in the record sections were estimate the degree of velocitydepth ambiguity is to
picked manually (Tables 1, 2, 3). Most of the time, the perform sensitivity tests, such as checkerboard tests
PmP cannot be identified on the shallow-water instru- analysis. This procedure consists of using synthetic
ments. The estimated picking error takes into account anomalies of various sizes at specific places in the
the quality of the phase and a possible systematic shift in models to evaluate at what point these anomalies are
the arrival identification in order to weight the different well-resolved by the travel time inversion.
picks during the inversion. A 4050 ms and 6070 ms
picking error was assigned respectively to near offsets 3.2. Model construction and parameterization
and far offset Pg arrivals. A 8090 ms error was
assumed for Pn phases, and 100 ms for PmP phases. In order to take into account the issues detailed
above, the inversion is conducted using a three step
3. Travel time tomography approach on each of the three profiles: (1) estimation of
the 1D reference velocity model; (2) generation of a set
3.1. Method of 100 1D Monte Carlo starting models; and (3) iterative
inversion of all the starting models. This approach is
A joint refraction and reflection travel time tomog- similar to that in Korenaga et al. (2000) (see Fig. 5 in
raphy method was used to build a two-dimensional Sallares et al., 2003 for more details).
velocity model along the three seismic lines (Korenaga The main difference between our and Sallares et al.'s
et al., 2000). The method allows the simultaneous (2003) approach is that we adopt as final velocity model
inversion of first arrival travel times (refracted phases) the best result of all the Monte-Carlo realizations,
and a reflection phase, in order to simultaneously characterized by the best RMS-2 couple (Fig. 5), rather
determine a 2-D velocity field and the geometry of a than the average of all the inversion models. The
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464 451

velocity features modelled along each profile are similar 3.3. Resolution and accuracy
in both the final best-fitting models (Fig. 5) and the
mean final models derived from the Monte-Carlo 3.3.1. Model parameter uncertainties
approach (Fig. 6-a1, b1, c1) but they are smoothed in The derivative weight sum (DWS, Fig. 6) (Toomey
the latter. and Foulger, 1989), allows a concise representation of

Fig. 5. Results of tomographic inversion of Pg, Pn and PmP arrivals. Iteration steps and RMS travel time misfits are displayed in Table 2. Models are
displayed from north to south : (a) profile 6, Northern Ecuador; (b) profile 4, Carnegie Ridge; (c) profile 2, Gulf of Guayaquil. White circles: OBS and
land stations locations; solid black lines: location of the Moho interface inverted from PmP arrivals; white dashed line: base of sedimentary cover
derived from coincident MCS data (Fig. 8); black dots: top of the oceanic crust from coincident MCS data converted to depth; whites dots: top of the
subduction channel (decollement) derived from coincident MCS data converted to depth; red dots: interpreted splay fault derived from coincident
MCS data converted into depth; the vertical grey line correlate the trench location from one profile to another.
452 A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464

the ray coverage in the best final model (Fig. 5). Along all profiles, the shape of the velocity
However, no quantitative measurement concerning the anomalies is well recovered beneath the OBS network
resolution and accuracy of the model can be associated ( 10-km spacing) and only a few small artefacts
with this type of representation. The mean deviations (b0.2 km/s) near the crustmantle boundary interface
of velocity and depth parameters from each Monte- are observed. The differences are especially signifi-
Carlo ensemble averaged can be interpreted as a cant in profile SIS 4, in agreement with the sparser
measure of the models uncertainties and therefore acquisition geometry and the less clear PmP and Pn
used to estimate the robustness of our iterative scheme phases recorded along this line, which affect the
(Fig. 6), which correlates well with the ray coverage resolvability of the deep parts of the model (Fig. 7 b).
imprint (DWS). Beneath zones where OBS are more than 15 km
Along profile SIS-4 (Fig. 6-b2 and b3), the N 10 km spaced, the pattern of velocity anomalies spreads in
OBS spacing seaward from the trench can be associated the direction of the mean ray paths (Fig. 7 c2), and
with a lack of short offset refractions and consequently velocity artefacts are larger and higher (0.20.4 km/s;
to a larger uncertainty in the upper part of the oceanic Fig. 7-c2).
crust (down to 7-km depth). Beyond 100-km distance, Along profile SAL-6, synthetic tests enable check-
only refractions sample the lower part of the model so it ing of the resolvability of the low velocity zone ob-
cannot be constrained properly between 15 and 22-km served in both the best fitting model and the mean
depth. final model. The location and amplitude of velocity
Along profile SAL-2 (Fig. 6-c2 and c3), the anomalies are reasonably well recovered within model
amplitude of arrivals recorded by OBS deployed on uncertainty limits (Fig. 7-a2), indicating that the ex-
the continental margin becomes rapidly weak due to a periment geometry and data set allow a distinc-
thick sedimentary cover, so no clear PmP or deep tion between positive and negative variations at 6 to
refracted phase can be identified (Fig. 4, OBS 57). Rays 17-km depth (Fig. 7-a2). The sensitivity of the data
poorly sample the model, from 90 km to the landward is therefore considered sufficient to resolve structures
end of the line, and consequently the Moho depth and with size and amplitudes similar to the low velocity
seismic velocities are poorly constrained in this area zone at this depth range implying that the low-velocity
beneath 14-km depth. zone observed in the final models is not an artefact
Along profile SAL-6 (Fig. 6-a2 and a3), the part of linked to the inversion procedure; even if the absolute
the model located between 50 and 70-km distance and value of velocity in this area is poorly constrained by
including the lower part of the continental margin the data set.
(below 10-km depth) and the subducting oceanic plate
(up to 15-km depth) shows the largest velocity 3.4. Results of tomographic inversion
uncertainties (0.20.25 km/s). Absolute velocities are
not well constrained there. In this zone the DWS shows The tomographic inversion of travel times provides
a lack in ray coverage associated with the shadow zone three 2D velocity models across the Ecuador convergent
observed on several record sections (Fig. 2, OBH 116 margin and the geometry of the Moho of the Nazca plate
and OBS 121). This shadow zone appears as a thick low (Fig. 5).
velocity zone that encloses the lower part of the margin Since there is no clear reflector separating the
and the inter-plate contact zone in both the best fitting overriding margin from the oceanic subducting plate in
model (Fig. 5a) and the final averaged Monte-Carlo the wide-angle seismic data, except along profile SIS-4
model (Fig. 6-a1). The 100 2D final models set linked to (Graindorge et al., 2004), we have digitized and converted
the Monte-Carlo analysis shows that the low velocity to depth the inter-plate contact reflectors observed along
zone occurs systematically and independently of the coincident available MCS sections (Collot et al., 2002;
initial velocity model considered. Checkerboard tests Marcaillou, 2003; Collot et al., 2004; Calahorrano, 2005).
will help to estimate the resolvability of this low The top of the subducting plate and the decollement,
velocity zone. which mark respectively the base and the top of the
subduction channel, are displayed as well as several major
3.3.2. Checkerboard tests faults affecting the margin (Fig. 5). Only long-wavelength
A number of checkerboard tests (Fig. 7) has been velocity anomalies are determined from travel time
performed in order to determine the size and amplitude inversion and so it is likely that velocity in the subduction
of synthetic velocity anomalies resolved in different channel is not properly calculated. Consequently the
parts of the velocity models. thickness of the subduction channel observed in the
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464
Fig. 6. Results of the Monte-Carlo analysis and ray tracing representation for (a) profile SAL-6; (b) profile SIS-4; (c) profile SAL-2. (1) Final velocity model derived from averaging all Monte-Carlo
ensembles, the average models are computed by taking the mean values of the 100 final models at each point of the grid; the two dashed lines indicates the variation in Moho location for the 100
Monte-Carlo inversions. (2) Corresponding velocity and Moho depth uncertainties (contours interval is set at 0.05 km/s). The velocity uncertainties correspond to the average of the absolute difference
between the average velocity derived from the 100 final models at each node of the model and the velocity obtained for each of the 100 final models at the same node. (3) Derivative weight sum (DWS)
provides a quantitative representation of the ray coverage (in white) calculated in the best final velocity models (Fig. 5).Uncertainties are much larger in the lower part of the models (imaged by PmP
and Pn phases only), than in the upper crust ( 0.15 to 0.20 km/s) along all profiles. In profile SIS-4, the number of Pn and PmP picked was too small to allow a regular sampling of the crustmantle
boundary (as confirmed by the DWS) so the lower crust is poorly resolved. Uncertainty of Moho depths is usually lower than 5 km, the maximum is reached beneath the margin where the Moho is
deeper and velocity uncertainties higher than 0.15 km/s. In the shallower part of the floating reflector (seaward each profile) the depth uncertainty is generally smaller than 3 km.

453
454
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464
Fig. 7. Results of the checkerboard tests for (a) profile SAL-6; (b) profile SIS-4; (c) profile SAL-2. The synthetic models consist of the addition of two sinusoidal velocity anomalies (8 5 km in the
upper part of the model; and 25 12 km in the lower part) featuring periodic positive and negative amplitudes (up to 5%) to the best fitting velocity models (Fig. 5). The same source receiver geometry
is used to compute synthetic travel times but random noise is added. Subsequently, these synthetic data sets are inverted, starting from the corresponding unperturbed models. (1) True model
perturbation showing the amplitude and geometry of velocity anomalies; (2) recovery of anomalies obtained after four iterations. The Moho geometry is successfully recovered with an uncertainty of
less than 1 km. Thus, we can suggest that the velocitydepth ambiguity raised by the tomography is insignificant in our study and that the data sets have enough spatial resolution to resolve velocity
anomalies with the size and amplitude of the ones considered here.
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464 455

Table 4
Synthetic view on the results obtained along the three profiles

tomographic models must be considered as an upper- These differences will be discussed in the next sections.
bound. Nevertheless, the comparison of wide-angle The structure of the oceanic crust is discussed using a
seismic models and MCS data enables the inter-plate classical 3 layer scheme: oceanic layer 1 consisting of
contact in the velocity model to be located, helping to sediments with velocities lower than 3.5 km/s; oceanic
correlate and interpret velocity structures. layer 2 (ol2) characterized by velocities ranging from
The uncertainty on the Moho depth is 1 km in the 3.5 to 6.5 km/s and a high vertical velocity gradient
better-constrained central part of the model (Fig. 6). The and oceanic layer 3 (ol3) characterized by velocities
crustmantle transition is marked by a transition from higher than 6.5 km/s, its base defined by the Moho
crustal velocities (b7 km/s) to mantle velocities (7.5 to discontinuity.
8.0 km/s). The agreement between the velocity According to the velocity models, we find consider-
transition and the Moho location is quite good in the able contrasts in structure, morphology, and sedimentary
central part of the model, where both parameters are cover for both plates. We propose a zonation of the
well constrained according to the outcome of the South ColombiaEcuador convergent margin into three
uncertainty analysis. The dip of the subducting plate is contrasted areas (Table 4, Fig. 1): the North Zone (2N
estimated from the dip of the Moho along the different 1N), the Central Zone (1N2.5S) and the South Zone
profiles, which is consistent with the dip of the top (2.53.5S).
of the oceanic crust derived from MCS data (Fig. 5,
Table 4). 4.1. The North Zone: profile SAL-6 (north Ecuador,
2N1N)
4. Interpretation
Along profile SAL-6, the velocity structure of the
The tomographic models are not split into discrete oceanic crust is well constrained from 10 to 40 km
layers. Nevertheless, to support the geological interpre- distance because of the 7 OBS located west of the trench
tation of the models, several model bodies or layers are (Fig. 5a). A 1.5 to 2.0-km thick sedimentary cover
differentiated based mainly on the velocities. (Vp b 3.5 km/s) overlays oceanic layer 2 characterized by
Along profile SAL-6 the base of the sedimentary a large vertical velocity gradient, ranging from 3.5 km/s
cover inferred from the coincident MCS profile SIS-44 at 5-km depth to 6.5-km/s at 7-km depth. The small
(Fig. 8) correspond to velocities ranging from 3.5 to velocity gradient in the lower part of the oceanic crust
4.0 km/s. At 4-km depth seismic velocities reach (Vp from 6.5 to 7.2 km/s) is in good agreement with
3.6 km/s for sands (30% porosity) or marine shales average oceanic layer 3 velocities (Mutter and Mutter,
(Japsen et al., 2007). In this paper, velocities lower than 1993). The Moho is 10 to 11-km deep, leading to a 6.0 to
3.5 km/s are assumed to represent sedimentary rocks. 7.0-km thick igneous oceanic crust with an ol2/ol3
Based on this assumption it is likely that the thickness of thickness ratio of 0.59.
the sedimentary cover is underestimated because at The velocity structure is consistent with normal
depth velocities in highly compacted sediments are oceanic crust showing no evidence of hotspot influence
likely to exceed 3.5 km/s. (e.g. White et al., 1992). The hypothesis suggesting a
The margin basement displays velocities ranging continuation of the CocosMalpelo Ridge east of the
from 3.5 to 5.5 km/s along profile SAL-2 but velocities Yaquina Graben, between 2N and 4N (Sallares and
reach 6.2 km/s at depth along profiles SAL-6 and SIS-4. Charvis, 2003), is not supported by our dataset.
456 A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464

Fig. 8. Multichannel seismic reflection line SIS-44 across the margin wedge (Collot et al., 2004). This MCS line is coincident with profile SAL-6 (see
location in Fig. 1). (Top) Line drawing of line SIS-44 showing the splay fault, the decollement and the top of oceanic crust as interpreted from (Collot
et al., 2004) and used in this study (Fig. 5). (Bottom) Close-up showing the splay fault that is connected to a summit graben associated with a gentle
rise of seafloor. The splay fault appears to connect to the decollement at depth.

East of the trench (distance 5090 km, Fig. 5a), margin. This estimate is consistent with previous studies
the Moho dips at 8.0 0.5. The base of the crust (Marcaillou, 2003; Collot et al., 2004; Calahorrano,
can be followed landward for a distance of 8090 km 2005) along the south Colombia margin, indicating that
with a good accuracy (Fig. 5a). Seismic velocities the trench infill is probably due to recent and massive
immediately above the Moho remain constant at turbiditic intakes, via canyons, in association with
around 7.07.5 km/s even if they are likely over- hemipelagic sedimentation.
estimated due the tomographic approach used. On The continental slope immediately east of the trench
the contrary, the velocity increases rapidly beneath is covered by 2.03.5-km thick sedimentary deposit.
the MCS reflector marking the top of the subducting This correlates with a small, probably ancient, accretion-
oceanic crust (black dots, Fig. 5a). Its velocity increases ary prism which has been described in this area using
from 5.0 km/s at 55-km to 6.0 km/s at 95-km distance MCS data (Fig. 8) (Marcaillou, 2003; Collot et al., 2004).
in the down-dip direction, which possibly indicates On the continental margin, the thickness of the sedi-
modifications of the upper part of the oceanic plate mentary cover is quite uniform around 2.5 km on
during its burial. average. Two sedimentary basins lie on the fore-arc
The sedimentary layers observed along MCS line (5590-km and 90125-km distance, Fig. 5a), and the
SIS-44 (Fig. 8) are consistent with velocities lower than sedimentary basin becomes thinner at the front of the
3.5 km/s modelled along the coincident wide-angle margin.
seismic line SAL-6 (Fig. 5a). The trench shows a 2 to 3- The continental platform is characterized by a
km thick sedimentary infill along the north Ecuador 600 m deep and 90-km wide morphological re-
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464 457

entrant limited landward by a subsiding coastal area The oceanic crust velocity structure is followed up
(Dumont et al., 2006). The subsidence of the continental to 90-km westward from the trench (Fig. 5b). The total
platform is supported by the presence of the 23-km thickness of the crust is 15 km here, based on the
thick sedimentary basins covering the margin basement. Moho location obtained from PmP travel times
Correlations with on-land drills made in the fore-arc inversion (15 to 18-km depth, Fig. 6-b1). ol2 is 3.5-
basins on top of the margin (Mountney and Westbrook, km thick, similar to that observed along SAL-6 ( 3 km)
1997) suggest a Palaeogene to Neogene calcareous, clay whereas ol3 is 10-km thick, two fold the thickness of ol3
and sand origin for this sedimentary infill (e.g. Deniaud, beneath line SAL-6.
2000). A travel time tomography analysis along a wide-
The margin basement is characterized by velocities angle line extending from the top of CR to the oceanic
ranging from 5.4 to 6.2 km/s (Fig. 9) that occupies a basin across its northern flank (Sallares et al., 2005)
large part of the deep margin, and can be followed from (see location of refraction line in Fig. 1) reveals that
35 km east of the trench to the coast at 7-km the maximum crustal thickness beneath the crest of
depth (see Table 5). The highest basement velocities the ridge is 19 km ( 100 km north of profile SIS-4)
( 6.2 km/s) are reached at 50-km distance from the with a similar velocity structure. If we compare this
trench, at 10-km depth. profile with profile SIS-4, it appears that the thick-
The reflector interpreted from the MCS data (Fig. 8) ness of ol2 remains uniform through CR whereas
as a major splay fault dipping to the east (Collot et al., the thickness of ol3 increases with increasing total
2004) marks the top of a large low velocity zone inferred thickness.
from our dataset (Fig. 5a). It clearly soles out at the inter- The crustal thickening is dominated by ol3 thicken-
plate boundary at 1314-km depth (45 km from the ing (ol2/ol3 thickness ratio b0.3), which confirms that
trench), and bounds the westward limit of the high the CR has originated from the interaction between the
velocity (Vp 6.0 km/s) upper plate basement. The GHS and the Galapagos spreading centre (Hey et al.,
seaward section of this system forms in fact a large low 1977; Wilson and Hey, 1995; Barckhausen et al., 2001;
velocity zone and appears significantly disturbed at the Sallares and Charvis, 2003), generating increased melt
deformation front (Fig. 8). production within the mantle source.
Due to the scarcity of data, the model is not well
4.2. The Central Zone: profile SIS-4 (Carnegie Ridge, constrained east of 130-km distance. Reflection from the
1N2.5S) top of the oceanic crust along coincident MCS lines and
modelling of intra-crustal wide-angle reflections pro-
The final velocity model we obtain along profile vide a dip ranging from 4 beneath the slope to 10 dip
SIS-4, located on the southern flank of the CR, is in beneath the Ecuadorian coast at 15-km depth
agreement with the layered model based on the forward (Graindorge et al., 2004). Deeper and farther inland
modelling of the same dataset (Graindorge et al., 2004) the WadatiBenioff zone dip varies between 35 and
but exhibits some differences. 25 between 1N and 1.5S, reaching at least a 150-

Fig. 9. Velocity logs showing vertical velocity variation below the top of basement (assumed to start at 4 km/s) versus depth across the margin 25 km
east of the trench (left) and 60 km east of the trench (right) for profile SAL-6 (red), profile SIS-4 (green) and profile SAL-2 (blue). The thickness of the
curves represents the uncertainty of the velocity inferred from the previous Monte-Carlo analysis. (Dashed lines): location of the inter-plate contact
zone.
458 A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464

Fig. 10. Seismicity of the EcuadorColombia margin. (a) Hypocenters from 1973 to 2006 are from the USGS National Earthquakes Information
Center. On line data available at http://neic.usgs.gov/neis/epic/epic.html. (b) Earthquake fault plane solutions from the Global Centroid Moment
Tensor (CMTS) database from 1976 to 2006 (Dziewonski et al., 1981). On line data available at http://www.globalcmt.org/.

km depth in the north and 200-km depth in the south et al., 1999; Collot et al., 2002; Mix et al., 2002;
(Guillier et al., 2001). Michaud et al., 2005).
On the CR and in the oceanic basin south of it, the At the southern flank of the CR, the margin wedge
sedimentary cover and trench fill are as thin as 0.5 to consist of a body with velocities that reach 5.7 km/s at
1.0 km assuming a velocity lower than 3.5 km/s. The 30 km east of the trench and 6.2 km/s at 40 km east of
lack of sediments on top of the CR and on its southern the trench, at 5-km depth (Table 5). These velocities
flank could be explained by the prominent topographic are higher than velocities described in similar context
feature of the CR, which acts as a barrier and influences and will be discussed in the next section.
marine current trajectories, sedimentary flux, deposition
(Gutscher et al., 1999; Collot et al., 2002; Mix et al., 4.3. The South Zone: profile SAL-2 (Gulf of Guayaquil,
2002; Michaud et al., 2005), and the erosive power of 2.5S3.5S)
strong marine currents.
Similarly, the continental shelf shows a very thin Along profile SAL-2 the structure of the oceanic
sedimentary cover (less than 1-km thick), which plate is followed up to 47 km westward from the
confirms that no subductionaccretion occurs along this trench. Seaward from the trench, the Moho is located at
segment (Gutscher et al., 1999; Collot et al., 2002; Mix 11.5-km depth (Fig. 5c). The crust of the oceanic plate
et al., 2002; Michaud et al., 2005). The coastal area is 8-km thick, only slightly thicker than along the
facing the CR has been uplifted as indicated by northern line (SAL-6). Major differences are observable
Quaternary marine terraces exposed at 200300 m in the very thin sedimentary cover which has a
elevation (De Vries, 1988). As a consequence, no rivers maximum thickness of 0.5 km (for velocities
coming from the Andes cross the Coastal Cordillera and b3.5 km/s), a thinner ol2 ( 2.5-km thick) with
precludes sediment influx towards the trench (Gutscher velocities from 3.5 to 6.0 km/s, and a 5-km thick
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464 459

ol3 (Vp rising from 6.2 to 7.0 km/s). Despite a velocity thinner sedimentary cover is associated with a progres-
structure close to normal oceanic crust, the low ol2/ol3 sive uplift of the margin basement possibly indicating
thickness ratio ( 0.40) and the slight thickening of ol3 weak compressive deformation expressed at the wedge
( 1 km thicker) suggest the crust underwent some (Calahorrano, 2005).
influence of the GHS activity during its emplacement The subsidence of the margin fore-arc in front of the
(Mutter and Mutter, 1993). GG, made evident by the thick sedimentary basin,
East of the trench, the oceanic Moho dips at supports a concept of large scale tectonic erosion. In
5.0 0.5. The thickness of the subducting oceanic such a context, the strong dip of the margin front appears
crust deduced from depth-migrated MCS reflectors is as a consequence of frontal erosion. Moreover, several
6.57.0 km. The velocity structure of ol2 and ol3 can be MCS sections acquired in this area show basal erosion
followed from the trench to 95-km distance and is associated with the subduction of down going sea-
characterized by isovelocity curves parallel to the Moho mounts carried by the oceanic plate (Sage et al., 2006).
at the base of the crust. The velocity at the top of the In this zone, seismic velocities ranging from 3.5
oceanic crust increases slightly during the down-warping to 5.2 km/s characterize the basement of the margin.
process (from 4 km/s at 4-km depth to 5.5 km/s at 13-km Here, velocities in the margin wedge are lower than the
depth), similar to our observations along SAL-6. ones described along northern and central Ecuador
The tomographic model shows the margin sedi- profiles (Fig. 5, Table 5). From the trench to 80-km
mentary cover (Vp 3.5 km/s) progressively thickening distance the continental shelf shows a progressive uplift
from 1 km at margin front to up to 7 km in the inner of this layer up to 1 km below the seafloor. At 100-km
part of the GG (between 50-km to 135-km distance). distance another high point of this layer reaches up to
Because of the compaction processes that likely occur at 2.5 km beneath the seafloor. It separates the upper slope
depth, it is possible that velocities higher than 3.5 km/s from the eastern basin. Velocities obtained for the
characterize the lower part of the sedimentary pile. margin substratum, 60 km east of the trench, are as low
Consequently, the sedimentary thickness in the basins as 5.2 km/s at 10-km depth, with a constant vertical
might be underestimated but this will not affect our gradient of 0.2 km/s/km (Fig. 9).
interpretation. Interpretation of industrial multichannel The GG roughly separates the northern fore-arc units
seismic profiles show that the continental slope (North Andean Block) made of oceanic terranes from the
(distance 50115 km, Fig. 5) is affected by a series of southern fore-arc units of continental composition
NS trending listric normal faults controlled by the (Mourier et al., 1988; Jaillard et al., 2002) but the limit
subduction processes (Witt et al., 2006; Witt, 2007). On between oceanic and continental basement is not precisely
the contrary, the continental platform (east of 115-km known. The Pion formation outcrops on-land in northern
distance along profile SAL-2) is mainly controlled by Peru and possibly extends off-shore beneath Banco Peru
the northward tectonic escape of the North Andean (Fig. 1) (Shepherd and Moberly, 1981).
Block since the Pleistocene. But both segments seem to Shadow zones appear along several wide-angle
have been subsiding zones since the Miocene (Witt seismic sections (e.g., Fig. 4). They cannot be modelled
et al., 2006; Witt, 2007). According to this interpreta- from inversion of first arrival travel times alone but
tion, up to 7-km thick eastern basin located on the possibly denote a low velocity subduction channel that
platform (distance 105145 km) may be related both to supports the basal erosion of the margin's underside. In
subductionerosion processes (i.e. tectonic erosion of fact, the resulting debris dragged along the inter-plate
the upper plates underside) since the Miocene, and to contact zone generally create a small velocity inversion
the NS extensive regime affecting the GG since the which gradually disappears with the down-going
Pleistocene. Whereas the sedimentary basin lying on the process(e.g., Sage et al., 2006).
upper part of the slope (distance 80110 km, Fig. 5c) is
more likely to be related to tectonic erosion at the base 5. Discussion
of the margin (Witt et al., 2006; Witt, 2007).
Westward, a thin sedimentary cover and infill In this study, we corroborate that the Moho beneath
characterize the deformation front and the trench area. the CR presents a strong dissymmetry with a sharp
Thus, the important sediment influx coming from the transition from the thick crust associated with volcanic
Andes does not seem to reach the trench, but is ridges to normal oceanic crust associated with the
predominantly trapped in the large sedimentary basins oceanic basin beneath the northern flank ( Sallares et al.,
lying on the continental platform and on the slope, in the 2005) (whereas the crust of the ridge thins progressively
region of greater subsidence. At the deformation front, a southward Sallares and Charvis, 2003; Sallares et al.,
460 A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464

2005). We show that the oceanic crust south of the CR Mamberti et al., 2004), prove a similar oceanic nature
southern flank (profile SAL-2) is slightly thicker than for the basement wedge which is part of the North
normal oceanic crust probably because of the Andean Block.
influence of the GHS, that is outlined by numerous Along the Peru margin, despite rather high seismic
seamounts described from multibeam bathymetric data velocities, the margin wedge consists of Palaeozoic or
(Fig. 1) (Sage et al., 2006). This is consistent with the Precambrian foliated metamorphic rocks of continental
present-day thermal anomaly associated with the GHS origin (Krabbenhoft et al., 2004). Velocities in the
that influences the generation of new oceanic crust margin wedge beneath the Gulf of Guayaquil are lower
within at least 350-km radius (Canales et al., 2002; than those observed along the central and north Ecuador
Sallares et al., 2005). margin and could denote either a different nature or a
To the north (profile SAL-6), we find normal higher degree of fracturing.
oceanic crust thickness which precludes the continua- A progressive increase of seismic velocities in the
tion of the Cocos-Malpelo Ridge east of the Yaquina backstop from 3.03.5 km/s near the trench to 4.5
Graben (Sallares et al., 2005). 6.2 km/s over a horizontal distance of 20 km is observed
The basement wedge offshore central Ecuador near in all of the well-resolved model based on forward
La Plata Island exhibits higher velocities at similar modelling or on tomographic inversion (see Table 5).
distance from the trench than any other well-constrained Velocities lower than 3.5 km/s are most likely to
margin wedge (Table 5). Margins with similar velocities represent sedimentary rocks either accreted at the
are the Nicoya Peninsula (Christeson et al., 1999) margin or deriving from the continent. Low velocities
consisting of a well-known ophiolitic complex, the observed at the margin toe are usually interpreted as
Chile margin offshore Antofagasta where the basement associated to fracturing, or alteration processes that
consists of basaltic to andesitic rocks (Sallares and affect the rocks of the margin basement and increase the
Ranero, 2005), the Shumagin Islands margin wedge porosity of these rocks (Christeson et al., 1999; Ranero
which consists of interstratified volcanic and sedimen- and von Huene, 2000; Krabbenhoft et al., 2004; Sallares
tary rocks (Lewis et al., 1988) and the Cascadia margin and Ranero, 2005). These processes decrease landward
offshore Oregon associated with the Siletz terrane of where the margin wedge is thicker and more mechan-
oceanic origin (Trehu et al., 1994; Gerdom et al., 2000; ically resistant.
Graindorge et al., 2003). The seismic velocities obtained along profile SAL-2
The margin of Ecuador is part of the North Andean for the margin basement of the GG cannot be
Block and mainly consists of oceanic cretaceous rocks immediately linked to accreted oceanic terranes as the
outcropping extensively on-land and on La Plata Island GG probably represents the southern limit of the North
immediately north of profile SIS-4 (Fig. 1). Then, the Andean Block (Witt, 2007). Velocities in the margin
margin wedge consists undoubtedly of the same type of basement range reach 4.5 to 5.0 km/s at 13-km depth.
oceanic rocks. Furthermore, the velocities encounted in They are similar to velocities in the OSA Peninsula
the margin basement are similar to the one determined in (Costa Rica) where the rocks outcropping are a middle
the upper crust of numerous oceanic plateaus like the mid-Eocene/Miocene melange dominated by basalt,
Kerguelen Plateau (Charvis et al., 1995), the Ontong chert, and limestone resulting from accretion of
Java Plateau (Gladczenko et al., 1997), the Caribbean seamounts (Vannucchi et al., 2006). It is also compa-
Plateau (Sallares et al., 1999) or the Cocos and Carnegie rable to the velocity structure described along the
ridges (Sallares et al., 2003, 2005) (refer to the Western Nankai Trough, where the margin basement is
compilation in Christeson et al., 1999 for more details). interpreted as an old metamorphosed accretionary
Velocities in the margin wedge offshore North complex, the Shimanto Belt (Kodaira et al., 2000;
EcuadorSouth Colombia are comparable to those Takahashi et al., 2002).
determined in the Cascadia margin and Eastern Nankai So, it is not possible from seismic velocity alone to
Trough, which include accreted oceanic fragment determine the nature of the basement beneath the GG. The
(Nakanishi et al., 1998; Dessa et al., 2004). They are low velocities obtained along profile SAL-2 are lower
not as high as the velocities observed in front of the than velocities described northward in front of the
Carnegie ridge at the toe of the margin. Nevertheless, Carnegie Ridge (this study) interpreted to be related to
outcrops of oceanic rocks on-land at the border between the oceanic North Andean Block. But they are also lower
Colombia and Ecuador, as well as on Gorgone Island than velocities described in the basement of the margin off
less than 140 km north of profile SAL-6 (Goosens et al., northern Peru (Krabbenhoft et al., 2004) which are related
1977; Marriner and Millward, 1984; Nivia, 1987; to the metamorphic basement. We assume that these low
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464 461

velocities in the GG basement should be mainly related to South of the CR, the strong subsidence of the margin,
a high degree of fracturing associated with the NS the low seismic velocity of the possibly highly fractured
extension in the Gulf itself and with the erosion of the basement and evidence for low velocity material in the
outer part of the margin (Witt, 2007). subduction channel despite the lack of sediments in the
Tectonic erosion is favoured in regions where trench support an intense erosion of the margin associated
convergence rates exceed 60-mm/yr and where the with a northsouth extension of the continental platform.
sedimentary cover is b1 km (Clift and Vannucchi, 2004). In this area, we suggest basal erosion is related to high
Rates of subductionerosion can be computed efficiently pressure fluids rather than a strong coupling.
from the offshore subsidence record and the geometry of The segmentation of the margin is primarily related
the margin (Kukowski and Oncken, 2006). Along the to the subduction of the CR that controls the sediment
margin, from Colombia to the GG, the slope of the margin distribution and increases significantly the interplate
is 15 to 20 (Figs. 1 and 5). Subsidence can be inferred by coupling in the Central Zone. The North Zone presents a
the thickness of low velocities and non-compacted well defined sequence of 4 earthquakes rupturing the
sediments. The steepness of the slope, the lack of an seismogenic interface during the 20th century (Fig. 1)
accretionary prism except near the Colombian border, (Kanamori and McNally, 1982; Mendoza and Dewey,
subsidence of the middle slope and of the fore-arc basin 1984; Beck and Ruff, 1984). The regular shallow
suggest significant erosion at the base of the overriding seismicity along the margin, based on worldwide
plate (Kukowski and Oncken, 2006). seismic networks (Fig. 10), is concentrated immediately
Subductionerosion can be envisaged as being East of the trench mainly with thrust focal mechanisms.
linked to strong coupling between the overriding and This supports that the subduction is partially locked
subducting plate, although processes like hydro-fractur- north of 2S as inferred from geodetic measurements
ing at the base of the overriding fore-arc crust are also and modelling (Trenkamp et al., 2002).
identified as a convincing explanation for basal tectonic The oceanic nature of the basement of the overriding
erosion. Fluid overpressure, finally, is thought to gene- plate in the Central and North Zone contributes to
rate fractures in the basal region of the upper plate increase the interplate coupling because of its higher
(Lallemand et al., 1994; von Huene et al., 2004; Ku- density. The contact between the top of the downgoing
kowski and Oncken, 2006). This loosens fragments that oceanic crust with strong material (velocities N5 km/s)
are then dragged into the subduction channel where they was interpreted as an estimate of the region of seismic
may undergo further downward transport (Lallemand onset (Pacheco et al., 1993). In our case, the lack of
et al., 1994; von Huene et al., 2004; Kukowski and shallow seismicity (b 20 km) in the GG (Fig. 10) is
Oncken, 2006). consistent with the change in the velocity structure of
In north Ecuador, a small accretionary prism observed the overriding plate and also possibly with a different
at the front of the margin suggests accretion whereas the process of basal erosion associated with fluid overpres-
strong subsidence of the margin supports basal erosion at sure. We suggest this region is characterized by a low
the deep part of the basement. The margin appears to be seismic coupling. The two large earthquakes that
at the limit between subductionaccretion which is more occurred in 1901 and 1953 are located near the northern
developed northward (towards Colombia) and subduc- and southern border of the GG (Fig. 10) and suggest a
tionerosion as it has been suggested from previous rapid change in seismic coupling to the north and to the
studies (Gutscher et al., 1999; Collot et al., 2002; south of the gulf.
Marcaillou, 2003; Collot et al., 2004).
At the CR, the increased buoyancy of the Nazca plate 6. Conclusions
is likely to cause the observed general uplift of the coast
in this area (Pedoja et al., 2006). There is no evidence of Wide angle seismic models obtained by travel time
subductionaccretion along this segment of the margin inversion at the Ecuadorian convergent margin reveal
on the MCS data (Collot et al., 2002). The high velocity variations in the crustal structure of the subducting
oceanic basement, which extends westward only 30 km Nazca plate and the overriding South American plate
from the trench (Fig. 5b) and seems to extend beneath the that strongly control the deformation of the margin and
whole margin, as well as the lack of fore-arc basin, do not the seismic cycle :
support a basal erosion scenario. Furthermore, the MCS
data reveal basal erosion related to the presence of (1) The North Zone exhibits normal oceanic crust
seamounts on the subducting plate, which affects the with no evidence for the influence of the GHS.
margin south of latitude 140 S (Sage et al., 2006). The overriding plate is characterized by a well
462 A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464

developed Cretaceous oceanic substratum, de- campaigns. Finally we thank Tim Minshull for his
scribed on land as the Diabasico and Pion detailed review which helped us to improve the initial
formations, underlain by high seismic velocities version of this manuscript.
(5.56.2 km/s). The structure of the margin
suggests it was affected by transient phases of Appendix A. Supplementary data
erosion and accretion.
The inner part of the margin shows a large low Supplementary data associated with this article can
velocity zone. Its upper limit correlates with a be found, in the online version, at doi:10.1016/j.
major splay fault that originates on the inter-plate epsl.2007.05.045.
contact zone and could have favoured the
propagation of the 1958 great earthquake rupture References
(Collot et al., 2004).
(2) The Central Zone reveals an over-thickened Barckhausen, U., Ranero, C.R., von Huene, R., Cande, S.C., Roeser,
oceanic crust (15 to 19 km) beneath the CR. The H.A., 2001. Roeser, revised tectonic boundaries in the Cocos Plate
off Costa Rica: implications for the segmentation of the convergent
Cretaceous oceanic basement is well-developed margin and for plate tectonic models. J. Geophys. Res., [Solid
and is likely to have been uplifted beneath the Earth] 106, 1920719220.
margin. The buoyancy of the Nazca plate and the Beck, S.L., Ruff, L.J., 1984. The rupture process of the Great 1979
high density of the margin basement both Colombia Earthquake evidence for the Asperity Model. J. Geophys.
contribute to an increased interplate coupling. Res. 89, 92819291.
Calahorrano, A. 2005. Structure de la marge du Golfe de Guayaquil
The fore-arc shows neither evidence of subsidence (Equateur) et proprits physiques du chenal de subduction partir
nor of frontal accretion. de donnes de sismique marine rflexion et rfraction, Thse de
(3) The South Zone is characterized by a 7 km thick Doctorat, Universit Pierre et Marie Curie (Paris VI).
oceanic crust, which was slightly influenced by Canales, J.P., Ito, G., Detrick, R.S., Sinton, J., 2002. Crustal thickness
along the western Galapagos Spreading Center and the compen-
the GHS during its emplacement 20 My ago. The
sation of the Galapagos hotspot swell. Earth Planet. Sci. Lett. 203,
continental slope and the platform show clues of 311327.
subsidence, possibly related to intense tectonic Charvis, P., Recq, M., Operto, S., Brefort, D., 1995. Deep structure of
erosion of the upper plate's base. We suggest that the KerguelenHeard Plateau and hot spot-related activity.
the low seismic velocities described for the Geophys. J. Int. 122, 899924.
margin basement could be related to an intense Christeson, G.L., McIntosh, K.D., Shipley, T.H., Flueh, E.R., Goedde, H.,
1999. Structure of the Costa Rica convergent margin, offshore Nicoya
fracturing associated with the presence of fluids Peninsula. J. Geophys. Res., [Solid Earth] 104, 2544325468.
and low seismic coupling. Clift, P., Vannucchi, P., 2004. Controls on tectonic accretion versus
erosion in subduction zones: implications for the origin and
Acknowledgements recycling of the continental crust. Rev. Geophys. 42.
Collot, J.-Y., Charvis, P., Gutscher, M.-A., Operto, S., 2002. The
Sisteur scientific party, exploring the EcuadorColombia active
The French scientific party was supported by Institut margin and inter-plate seismogenic zone. EOS 83 (185), 189190.
de Recherche pour le Dveloppement (IRD), Institut Collot, J.-Y., Marcaillou, B., Sage, F., Michaud, F., Agudelo, W.,
National des Sciences de lUnivers (INSU) and Institut Charvis, P., Graindorge, D., Gutscher, M.A., Spence, G., 2004. Are
Franais pour lExploitation de la Mer (IFREMER), rupture zone limits of great subduction earthquakes controlled by
upper plate structures? Evidences from Multichannel seismic
which provided ship time during the SISTEUR
reflection data acquired across the northern Ecuadorsouthwest
experiment. The SALIERI-2001 experiment was funded Colombia margin. J. Geophys. Res. 109 (14 pages).
by the German Ministry of Education, Research, Deniaud, Y., 2000. Enregistrement sdimentaire et structural de
Science and Technology (BMBF) under project no. lvolution godynamique des Andes quatoriennes au cours du
03G0159A. We are also indebted to H. Andersen, Nogne: tude des bassins davant-arc et bilans de masse. Universit
master of the R/V Sonne, and its crew for their support Joseph Fourier.
Dessa, J.X., Operto, S., Kodaira, S., Nakanishi, A., Pascal, G., Uhira,
during the SALIERI experiment and to Captain R. K., Kaneda, Y., 2004. Deep seismic imaging of the eastern Nankai
Davila and the crew of R/V Orion (INOCAR) from trough, Japan, from multifold ocean bottom seismometer data by
which OBS were operated during the SISTEUR combined travel time tomography and prestack depth migration.
experiment. During the Sisteur experiment the R/V J. Geophys. Res., [Solid Earth] 109.
Atalante (chief scientist Jean-Yves Collot) shot over the De Vries, T., 1988. The geology of late Cenozoic marine terraces
(tablazos) in northwestern Peru. J. South Am. Earth Sci. 1, 121136.
OBS networks. Also special thanks to Y. Hello, A. Dumont, J.F., Santana, E., Valdez, F., Tihay, J.P., Usselmann, P.,
Anglade and B. Yates who successfully operated the Iturralde, D., Navarette, E., 2006. Fan beheading and drainage
OBS at sea and to the scientific parties of both diversion as evidence of a 32002800 BP earthquake event in the
A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464 463

EsmeraldasTumaco seismic zone: a case study for the effects of from a wide-angle ocean bottom seismic survey. J. Geophys. Res.,
great subduction earthquakes. Geomorphology 74, 100123. [Solid Earth] 105, 58875905.
Dziewonski, A.M., Chou, T.A., Woodhouse, J.H., 1981. Determination Korenaga, J., Holbrook, W.S., Kent, G.M., Kelemen, P.B., Detrick,
of earthquake source parameters from waveform data for studies of R.S., Larsen, H.C., Hopper, J.R., Dahl-Jensen, T., 2000. Crustal
global and regional seismicity. J. Geophys. Res. 86, 28252852. structure of the southeast Greenland margin from joint
Flueh, E.R., Vidal, N., Ranero, C.R., Hojka, A., von Huene, R., Bialas, refraction and reflection seismic tomography. J. Geophys.
J., Hinz, K., Cordoba, D., Danobeitia, J.J., Zelt, C., 1998a. Seismic Res., [Solid Earth] 105, 2159121614.
investigation of the continental margin off- and onshore Valpar- Krabbenhoft, A., Bialas, J., Kopp, H., Kukowski, N., Hubscher, C.,
aiso, Chile. Tectonophysics 288, 251263. 2004. Crustal structure of the Peruvian continental margin from
Flueh, E.R., Fisher, M.A., Bialas, J., Jonathan, R., Klaeschen, D., wide-angle seismic studies. Geophys. J. Int. 159, 749764.
Kukowski, N., Parsons, T., Scholl, D.W., ten Brink, U., Trehu, A.M., Kukowski, N., Oncken, O., 2006. Subduction erosion the normal
Vidal, N., 1998b. New seismic images of the Cascadia subduction mode of fore arc material transfer along the Chilean margin? In:
zone from cruise SO108-ORWELL. Tectonophysics 293, 6984. Oncken, O., Chong, G., Franz, G., Giese, P., Gtze, H.J., Ramos,
Flueh, E.R., Bialas, J., Charvis, P., 2001. Cruise Report SO159 V.A., Strecker, M.R., Wigger, P. (Eds.), The Andes active
Salieri. Geomar, Kiel, p. 256. subduction orogeny. . Frontiers in Earth Science Series, vol. 1.
Gerdom, M., Trehu, A.M., Flueh, E.R., Klaeschen, D., 2000. The Springer-Verlag, Berlin, pp. 217236.
continental margin off Oregon from seismic investigations. Lallemand, S.E., Schnurle, P., Malavieille, J., 1994. Coulomb theory
Tectonophysics 329, 7997. applied to accretionary and nonaccretionary wedges possible
Gladczenko, T.P., Coffin, M.F., Eldholm, O., 1997. Crustal causes for tectonic erosion and or frontal accretion. J. Geophys.
structure of the Ontong Java Plateau: modeling of new gravity Res., [Solid Earth] 99, 1203312055.
and existing seismic data. J. Geophys. Res., [Solid Earth] 102, Lewis, S.D., Ladd, J.W., Bruns, T.R., 1988. Structural development of an
2271122729. accretionary prism by thrust and strikeslip faulting Shumagin
Goosens, P.J., Rose, W.I., Flores, D., 1977. Geochemistry of tholeiites Region, Aleutian Trench. Geol. Soc. Amer. Bull. 100 767&.
of the basic igneous complex of Northwestern South America. Lizzaralde, D. 1997. Crustal structure of rifted and convergent
Geol. Soc. Amer. Bull. 88, 17111720. margins: The U.S. East Coast and Aleutian margins, PhD Thesis,
Graindorge, D., Spence, G., Charvis, P., Collot, J.-Y., Hyndman, R., Woods Hole, Mass.
Trehu, A., 2003. Crustal structure beneath the Strait of Juan de Lonsdale, P., 2005. Creation of the Cocos and Nazca plates by fission
Fuca and Southern Vancouver Island from seismic gravity of the Farallon plate. Tectonophysics 404, 237264.
analyses. J. Geophys. Res. 108, 2484. Lonsdale, P., Klitgord, K.D., 1978. Structure and tectonic history of
Graindorge, D., Calahorrano, A., Charvis, P., Collot, J.-Y., Bethoux, the eastern Panama Basin. Geol. Soc. Amer. Bull. 89, 981999.
N., 2004. Seismic evidences of Carnegie Ridge subduction and the Mamberti, J.M., Lapierre, H., Bosch, D., Jaillard, E., Hernandez, J.,
Ecuadorian margin structure at 1.4S. Geophys. Res. Lett. 31. Polv, M., 2004. The Early Cretaceous San Juan Plutonic Suite,
Guillier, B., Chatelain, J.L., Jaillard, E., Yepes, H., Poupinet, G., Fels, Ecuador: a magma chamber in an oceanic plateau? Can. J. Earth
J.F., 2001. Seismological evidence on the geometry of the orogenic Sci. 41, 12371258.
system in central-northern Ecuador (South America). Geophys. Marcaillou, B., Rgimes tectoniques et thermiques de la marge Nord
Res. Lett. 28, 37493752. EquateurSud Colombie (03.5N) Implications sur la
Gutscher, M.A., Malavieille, J., Lallemand, S., Collot, J.Y., 1999. sismogense, Thse de Doctorat, Universit Pierre et Marie
Tectonic segmentation of the North Andean margin: impact of the Curie (Paris VI), 2003.
Carnegie Ridge collision. Earth Planet. Sci. Lett. 168, 255270. Marriner, G.F., Millward, D., 1984. The petrology and geochemistry
Hey, R., 1977. Tectonic evolution of the CocosNazca spreading of Cretaceous to recent volcanism in Colombia the magmatic
center. Geol. Soc. Amer. Bull. 88, 14041420. history of an accretionary plate margin. J. Geol. Soc. 141, 473486.
Hey, R., Johnson, G.L., Lowrie, A., 1977. Recent Plate Motions in McIntosh, K.D., Nakamura, Y., 1998. Crustal structure beneath the
Galapagos Area. Geol. Soc. Amer. Bull. 88, 13851403. Nanao forearc basin from TAICRUST MCS/OBS Line 14. Terr.
Jaillard, E., Ordonez, M., Benitez, S., Berrones, G., Jimenez, M., Atmos. Ocean. Sci. 9, 345362.
Montenegro, G., Zambrano, I., 1995. Basin development in an Mendoza, C., Dewey, J.W., 1984. Seismicity associated with the Great
accretionary, oceanic-floored fore-arc setting : southern coastal ColombiaEcuador Earthquakes of 1942, 1958, and 1979
Ecuador during late Cretaceaouslate Eocene time. In: Tankard, A.J., implications for Barrier Models of Earthquake Rupture. Bull.
Suarez, R., Welsink, H.J. (Eds.), Petroleum Basins of South America. Seismol. Soc. Am. 74, 577593.
Volume AAPG Memoire, vol. 62, pp. 615632. Michaud, F., Chabert, A., Collot, J.-Y., Sallars, V., Flueh, E.R.,
Jaillard, E., Benitez, S., Mascle, G.H., 1997. Palaeogene deformations Charvis, P., Graindorge, D., Gustcher, M.-A., Bialas, J., 2005.
of the forearc zone of south Ecuador in relation to the geodynamic Fields of multi-kilometer scale sub-circular depressions in the
evolution. Bull. Soc. Gol. Fr. 168, 403412. Carnegie ridge sedimentary blanket: effect of underwater carbon-
Jaillard, E., Herail, G., Monfret, T., Worner, G., 2002. Preface ate dissolution ? Mar. Geol. 216, 205219.
Andean geodynamics: main issues and contributions from the 4th Mix, A.C., Tiedemann, R., Blum, P., shipboard scientists (Eds.), 2002.
ISAG, Gottingen. Tectonophysics 345, 115. Southeast Pacific Paleoceanographic Transects Sites 12321242,
Japsen, P., Mukerji, T., Mavko, G., 2007. Constraints on velocitydepth ODP Leg 202.
trends from rock physics models. Geophys. Prospect. 55, 135154. Mountney, N.P., Westbrook, G.K., 1997. Quantitative analysis of
Kanamori, H., McNally, K.C., 1982. Variable rupture mode of the Miocene to Recent forearc basin evolution along the Colombian
subduction zone along the EcuadorColombia Coast. Bull. convergent margin. Basin Res. 9, 177196.
Seismol. Soc. Am. 72, 12411253. Mourier, T., Laj, C., Megard, F., Roperch, P., Mitouard, P., Medrano,
Kodaira, S., Takahashi, N., Park, J.O., Mochizuki, K., Shinohara, M., A.F., 1988. An accreted continental terrane in Northwestern Peru.
Kimura, S., 2000. Western Nankai Trough seismogenic zone: Results Earth Planet. Sci. Lett. 88, 182192.
464 A. Gailler et al. / Earth and Planetary Science Letters 260 (2007) 444464

Mutter, C.Z., Mutter, J.C., 1993. Variations in thickness of layer 3 dominate Smith, W.H.F., Sandwell, D.T., 1997. Global sea floor topography
oceanic crustal structure. Earth Planet. Sci. Lett. 117, 295317. from satellite altimetry and ship depth soundings. Science 277,
Nakamura, Y., Garmany, J. (Eds.), 1991. Development of Upgraded 19561962.
Ocean Bottom Seismograph. University of Texas Institute for Stavenhagen, A.U., Flueh, E.R., Ranero, C., McIntosh, K.D., Shipley,
Geophysics Technical Report, p. 45. T., Leandro, G., Schulze, A., Danobeitia, J.J., 1998. Seismic wide-
Nakamura, Y., McIntosh, K., Chen, A.T., 1998. Preliminary results of angle investigations in Costa Rica: a crustal velocity model from
a large offset seismic survey west of Hengchun Peninsula, southern the Pacific to the Caribbean coast. Zentralbl. Geol. Paleont. 3-6,
Taiwan. Terr. Atmos. Ocean. Sci. 9, 395408. 393408.
Nakanishi, A., Shiobara, H., Hino, R., Kodaira, S., Kanazawa, T., Takahashi, N., Kodaira, S., Nakanishi, A., Park, J.O., Miura, S., Tsuru,
Shimamura, H., 1998. Detailed subduction structure across the T., Kaneda, Y., Suyehiro, K., Kinoshita, H., Hirata, N., Iwasaki, T.,
eastern Nankai Trough obtained from ocean bottom seismographic 2002. Seismic structure of western end of the Nankai trough
profiles. J. Geophys. Res., [Solid Earth] 103, 2715127168. seismogenic zone. J. Geophys. Res., [Solid Earth] 107.
Nivia, A. 1987. Geochemistry and origin of the amaime and volcanic Toomey, D.R., Foulger, G.R., 1989. Tomographic inversion of local
sequences southwestern Colombia, Unpublished M. Phil. thesis, earthquake data from the HengillGrensdalur Central Volcano
University of Leicester. Complex, Iceland. J. Geophys. Res., [Solid Earth and Planets] 94,
Pacheco, J.F., Sykes, L.R., Scholz, C.H., 1993. Nature of seismic 1749717510.
coupling along simple plate boundaries of the subduction type. Trehu, A.M., Asudeh, I., Brocher, T.M., Luetgert, J.H., Mooney, W.D.,
J. Geophys. Res., [Solid Earth] 98, 1413314159. Nabelek, J.L., Nakamura, Y., 1994. Crustal Architecture of the
Pedoja, K., Dumont, J.F., Lamothe, M., Ortlieb, L., Collot, J.Y., Cascadia Fore-Arc. Science 266, 237243.
Ghaleb, B., Auclair, M., Alvarez, V., Labrousse, B., 2006. Plio- Trenkamp, R., Kellogg, J.N., Freymueller, J.T., Mora, H.P., 2002.
Quaternary uplift of the Manta Peninsula and La Plata Island and Wide plate margin deformation, southern Central America and
the subduction of the Carnegie Ridge, central coast of Ecuador. northwestern South America, CASA GPS observations. J. South
J. South Am. Earth Sci. 22, 121. Am. Earth Sci. 15, 157171.
Ranero, C.R., von Huene, R., 2000. Subduction erosion along the Vannucchi, P., Fisher, D.M., Bier, S., Gardner, T.W., 2006. From
Middle America convergent margin. Nature 404, 748752. seamount accretion to tectonic erosion: formation of Osa Melange
Reynaud, C., Jaillard, E., Lapierre, H., Mamberti, M., Mascle, G.H., and the effects of Cocos Ridge subduction in southern Costa Rica.
1999. Oceanic plateau and island arcs of southwestern Ecuador: Tectonics 25.
their place in the geodynamic evolution of northwestern South von Huene, R., Ranero, C.R., Vannucchi, P., 2004. Generic model of
America. Tectonophysics 307, 235254. subduction erosion. Geology 32, 913916.
Sage, F., Collot, J.Y., Ranero, C.R., 2006. Interplate patchiness and White, R.S., McKenzie, D., O'Nions, R.K., 1992. Oceanic crustal
subductionerosion mechanisms: Evidence from depth-migrated thickness from seismic measurements and rare earth element
seismic images at the central Ecuador convergent margin. Geology inversions. J. Geophys. Res. 97, 1968319715.
34, 9971000. Wilson, D.S., Hey, R.N., 1995. History of rift propagation and
Sallares, V., Charvis, P., 2003. Crustal thickness constraints on the magnetization intensity for the CocosNazca spreading center.
geodynamic evolution of the Galapagos Volcanic Province. Earth J. Geophys. Res., [Solid Earth] 100, 1004110056.
Planet. Sci. Lett. 214, 545559. Witt, C., Constraints on the tectonic evolution of the North Andean
Sallares, V., Ranero, C.R., 2005. Structure and tectonics of the Block trailing edge: evolution of the Gulf of GuayaquilTumbes
erosional convergent margin off Antofagasta, north Chile (23 Basin and the intermountain basins of the Central Ecuadorian
degrees 30 S). J. Geophys. Res., [Solid Earth] 110. Andes, Thse de Doctorat, Universit Pierre et Marie Curie, 2007.
Sallares, V., Daobeitia, J.J., Flueh, E.R., Leandro, G., 1999. Seismic Witt, C., Bourgois, J., Michaud, F., Ordonez, M., Jimenez, N., Sosson,
velocity structure across the middle American landbridge in M., 2006. Development of the Gulf of Guayaquil (Ecuador) during
northern Costa Rica. Geodynamics 27, 327344. the Quaternary as an effect of the North Andean block tectonic
Sallares, V., Charvis, P., Flueh, E.R., Bialas, J., 2003. Seismic structure escape. Tectonics 25.
of Cocos and Malpelo Volcanic Ridges and implications for hot Ye, S., Flueh, E.R., Klaeschen, D., vonHuene, R., 1997. Crustal
spotridge interaction. J. Geophys. Res., [Solid Earth] 108. structure along the EDGE transect beneath the Kodiak shelf off
Sallares, V., Charvis, P., Flueh, E.R., Bialas, J., 2005. Seismic structure Alaska derived from OBH seismic refraction data. Geophys. J. Int.
of the Carnegie ridge and the nature of the Galapagos hotspot. 130, 283302.
Geophys. J. Int. 161, 763788.
Shepherd, G.L., Moberly, R., 1981. Coastal Structure of the
Continental-Margin, Northwest Peru and Southwest Ecuador.
Geol. Soc. Am., Mem. 154, 351391.

You might also like