Sie sind auf Seite 1von 15

Available online at www.sciencedirect.

com

Earth and Planetary Science Letters 267 (2008) 453 467


www.elsevier.com/locate/epsl

Nonlinear variations of the physical properties along the southern Ecuador


subduction channel: Results from depth-migrated seismic data
Alcinoe Calahorrano B. a,, Valent Sallars b , Jean-Yves Collot a ,
Franoise Sage a , Cesar R. Ranero c
a

Gosciences Azur, Universit de Nice Sophia-Antipolis, Institut de Recherche pour le Dveloppement (IRD), Universit Pierre et Marie Curie, Centre National de
la Recherche Scientifique (CNRS), BP 48, 06235 Villefranche-sur-Mer cedex, France
b
Unidad de Tecnologa Marina-CMIMA, Consejo Superior de Investigaciones Cientficas (CSIC), Passeig Maritm de la
Barceloneta 37-49, 08003, Barcelona, Spain
c
Instituci Catalana de Recerca i Estudis Avanats (ICREA) at Instituto de Ciencias del Mar-CMIMA, Consejo Superior de Investigaciones Cientficas (CSIC),
Passeig Maritm de la Barceloneta 37-49, 08003, Barcelona, Spain
Received 15 June 2007; received in revised form 25 October 2007; accepted 29 November 2007
Available online 1 February 2008
Editor: C.P. Jaupart

Abstract
We use two high-quality pre-stack depth-migrated multichannel seismic profiles acquired to quantify physical properties variations of underthrust
sediments along the first ~ 30 km of subduction off the erosional southern Ecuadorian margin. Seismic data show three zones along the subduction
channel (referred to as Zones I, II and III) characterized by distinct velocity and velocity-derived physical properties, which are in agreement with
values estimated from experimental results of deformation in granular media. These three zones result from transformational changes of underthrust
sediments governed by fundamentally different physical processes that control their mechanical behavior at increasing confining pressures. Based on
our observations and its comparison with experimental results, we argue that the transformations undergone by underthrust sediments as they dip into
the subduction zone are the following: within Zone I, progressively increasing velocity (and decreasing velocity-derived porosity) indicates
continuous sediment compaction, which must be accompanied by effective fluid drainage along the dcollement and/or across the accretionary
wedge. The underthrust material is here unconsolidated from a mechanical point of view. Laboratory experiments indicate that the dominant
processes at this range of pressures are grain rolling, particle rotation and frictional slip at grain contacts. Within Zone II, velocity (and porosity)
remains constant for ~ 16 km (SIS-72) and ~ 12 km (SIS-18). This suggests undrained conditions resulting in growing fluid overpressure at the
subduction channel. Grain deformation is similar to Zone I. Within Zone III, velocity increases and porosity falls rapidly, indicating sediment
compaction and subsequent release of over-pressured fluids, where grain deformation is likely to be elastic. This might be the dominant process until
the grains attain their crushing strength, resulting in granular cataclasis and, eventually, in the collapse of the system. We suggest that over-pressured
fluid release may induce hydrofracturation and it is likely to increase inter-plate coupling down from Zone III.
2007 Elsevier B.V. All rights reserved.
Keywords: subduction channel; velocity inversion; fluid overpressure; grain deformation

1. Introduction

Corresponding author. Instituto de Ciencias del Mar-CMIMA, Consejo


Superior de Investigaciones Cientficas (CSIC), Passeig Maritm de la
Barceloneta 37-49, 08003, Barcelona, Spain. Tel.: +34 93 230 9500; fax: +34
93 230 95 55.
E-mail address: alcinoe@cmima.csic.es (A. Calahorrano B.).
0012-821X/$ - see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.epsl.2007.11.061

In convergent margins, when subduction takes place, part of


the oceanic and continental sediments cumulated in the trench is
commonly dragged with the downgoing plate beneath the
margin. This downgoing sediment form the so-called subduction channel (SC), a poorly consolidated and fluid-rich layer
that is structurally squeezed between upper and lower plates

454

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

(Scholl et al., 1977; Shreve and Cloos, 1986). The SC has been
clearly imaged using multichannel seismic (MCS) data along
tens of kilometers beneath several accretionary margins like
Nankai (e.g. Moore et al., 2001; Bangs et al., 2004), Barbados
(e.g. Westbrook et al., 1982; Moore and Shipley, 1988) and
Cascadia (e.g. Clowes et al., 1987; Davis and Hyndman, 1989;
Hyndman et al., 1990).
Physical and mechanical properties of subducting material
strongly influence the shape and tectonic deformation of the
margin (Davis et al., 1983; Lallemand et al., 1994). The
subducting material contains pore-filling fluids in variable
amounts depending on the nature and thickness of the underthrust sediment deposit. In high-permeability conditions, pore
fluids are expelled as a response to the rising pressure resulting
from the landward-increasing load of the overriding-plate.
Sediment framework will mainly support this load and will be
progressively compacted. In the case that a significant part of
fluids remain trapped within pores, which typically occurs when
subducting sediments are rapidly buried beneath the margin and
permeability conditions are low, the overburden pressure is
transferred into increasing fluid over-pressures along the dcollement and/or within the sediment column (e.g. Bangs et al.,
1990). Fluid pressure variations are believed to play a major
role in controlling deformation processes and fault dynamics
along subduction zone megathrusts (Moore, 1989; Le Pichon
et al., 1993; Moore and Saffer, 2001; Sage et al., 2006). Such
deformation processes include frontal accretion, wedge thickening by out-of-sequence thrusting, subduction erosion, and
underplating (Cloos and Shreve, 1988). Understanding the
physical behaviour of underthrust material is critical because it
controls mechanical processes such as inter-plate friction,
hydrofracturing (Sibson, 1981), and the location of the
dcollement (e.g. Le Pichon et al., 1993; McIntosh and Sen,
2000). In addition, they also influence mass and fluid budgets
(e;g. Saffer and Bekins, 1998), heat transfer (e. g. Hyndman
et al., 1995) and the down-dip physical and chemical
transformations of subducted material (e.g. Kastner et al.,
1991). These transformations that occur as the plate drives
deeper into the subduction zone are believed to play an
important role on both the location of the seismogenic zone
(e.g., Vrolijk, 1990) and the amount of co-seismic slip
propagation (e.g., Moore and Saffer, 2001).
Physical properties of the SC material deeply rely on the
kinematics, margin stress regime, sediment supply and water
content, in combination with the age and crustal structures of
both the downgoing and overriding-plates. At present day, the
knowledge of the physical properties of the SC material is
limited to direct measurements of sediment porosity/density and
seismic velocity obtained during scientific drilling at the leading
edge of sediment prism (b ~ 4 km from the trench and b 2 km
below sea floor, bsf). This is the case, for example, of the
accretionary prisms of Nankai (Moore et al., 2001) and Barbados (Mascle et al., 1988; Moore et al., 1995), as well as the
frontal prism of Costa Rica (Bolton et al., 2000). In contrast,
only few indirect estimates of subduction channel porosity and
fluid content at greater depths and distances from the
deformation front (N10 km) are available to date. Some of the

few examples are the accretionary complexes of Barbados


(Bangs et al., 1990) and Oregon (Cochrane et al., 1994; Yuan
et al., 1994), where SC porosity and fluid pressures have been
estimated based on Normal Move Out (NMO) velocity analysis
of MCS data. Likewise, von Huene et al. (1998) used Prestack
Depth Migration (PSDM) of MSC data to estimate porosity and
dewatering at the accretionary margin of Alaskan. Finally,
numerical modeling of consolidation and dewatering based on
borehole chemical and physical data measured at the margin's
toe has been also made in Barbados (Stauffer and Bekins, 2001)
and Nankai (Saffer and Bekins, 1998).
In this paper, we use MCS data acquired across the erosional
margin of southern Ecuador during the SISTEUR-2000 survey
(Collot et al., 2002) to (1) identify the main structures of the
margin, (2) obtain the seismic velocity field by means of PSDM,
and correspondent depth-images, (3) calculate velocity-derived
porosity at SC, as well as effective and pore pressure to quantify
fluid overpressure variations and dewatering along some
~ 30 km of subduction, and (4) discuss the causes and
implications of estimated physical properties, both across and
along the strike of the margin.
2. Tectonic setting
At the South Ecuador margin, offshore the Gulf of Guayaquil, the Nazca plate subducts eastwards beneath SouthAmerica at ~ 55 mm/yr (Trenkamp et al., 2002) (Fig. 1). In
addition to normal plate convergence, the Ecuador margin, as
part of the so-called North Andean Block, moves north-eastwards in response to subduction obliquity in Colombia and
strain partitioning of north-western South American plate. The
North Andean Block motion occurs along a major NE-trending
right-lateral strikeslip fault system (e.g. Winter et al., 1993;
Ego et al., 1996) at a rate of ~ 6 2 mm/yr (Trenkamp et al.,
2002). This motion is believed to have favored the opening the
Gulf of Guayaquil by NNE extension (Ego et al., 1996; Witt
et al., 2006) (Fig. 1).
Recent work based on MCS and swath bathymetry data
acquired during the SISTEUR survey shows that the Ecuador
margin is dominantly erosional and characterized by extension
(Collot et al., 2002; Calahorrano, 2005; Sage et al., 2006).
Despite this overall erosional behaviour, the margin off the Gulf
of Guayaquil is fronted by a ~ 310 km-wide sediment prism.
The volume and nature of terrestrial sediment supplied to the
trench or conveyed by the Nazca plate vary considerably along
this area. This variability is mainly associated to the erosion and
sediment transport of the Andes, the presence of the Grijalva
Fracture Zone (GFZ) and the Carnegie Ridge, and the marine
currents. The erosion of the Andean Western Cordillera provides
several hundred meters of terrigenous sediments transported
northwards to the trench by the Esmeraldas river, and southwards by the Guayas river up to the Gulf of Guayaquil (Fig. 1).
Sand turbidites with abundant wood fragments characterizing
the surficial trench deposits in the northern margin (Collot et al.,
2005) corroborate the erosion/transport of clastic material, and
illustrate sediment variation along the trench while contrasting
this deposits with dark-green hemipelagic mud with high

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

455

Fig. 1. Multibeam bathymetry of the Ecuadorian margin. Black lines are the track of SIS-72 and SIS-18, located off the Gulf of Guayaquil. White lines indicate the portion of
profile used in this study. Black dash line corresponds to the East limit of the North Andean Block (NAB). Thin dash lines correspond to the two major paths of sediment
transport to the trench: the Esmeraldas River (ER) and Guayas River (GR). Carnegie Ridge is subducting almost perpendicularly below the central-north Ecuadorian margin,
while the Grijalva Fracture Zone (GFZ) subducts obliquely under the Gulf of Guayaquil. NF shows normal, east-dipping faults in the southern flank of the Carnegie Ridge.

organic content, ash-rich turbidites and basal graded units of


reworked foraminifera in the central margin (Lonsdale, 1978).
The GFZ and the Carnegie Ridge are two major structural
features of the Nazca Plate that enter the Ecuador trench. The
GFZ is a NE-trending, 700 m-high scarp that separates the ~ 3234 Ma, smooth morphology of the ancient Farallon lithosphere
to the south (Lonsdale and Klitgord, 1978; Barckhausen et al.,
2001), from the younger (~ 024 Ma) and rougher Nazca
lithosphere to the north (Handschumacher, 1976; Hey, 1977).
Sediment accumulation is favored south of the GFZ-trench
intersection where the trench is the deepest (Fig. 1). The Carnegie Ridge, located north of the GFZ, is a prominent ~ 2 kmhigh, ~ 200 km-wide, EW-trending volcanic ridge resulting
from the interaction between the Galapagos hotspot and the

NazcaCocos spreading centre since ~ 20 Ma (Sallars et al.,


2005; Lonsdale and Klitgord, 1978; Sallars and Charvis,
2003). The southern flank of the Carnegie Ridge is spotted by
numerous seamounts that collide with the Ecuador margin
supplying the trench with abundant mass-wasting deposits
(Sage et al., 2006). The cores of site 157 from DSDP Leg 16,
and sites 1238 and 1239 from ODP Leg 202 in the southern
flank of the easternmost Carnegie Ridge indicate that in this
region, the basalt basement is draped by a ~ 400500 m thick
sediment sequence mainly corresponding to diatom nannofossil
ooze and nannofossil diatom ooze, with varying abundance of
clay, foraminifers siliceous microfossils, and some local
concentration of organic carbon, carbonate and ash layers.
Chert and micrite, indicate diagenesis of opal and carbonate at

456

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

Fig. 2. Line SIS-72. A) PSDM image of the front of the margin. B) Geological interpretation of seismic line. TSC, Top of subduction channel reflection. TOC, Top of
oceanic crust reflection. C) 2D velocity model. Notice yellowish green colors along the subduction channel, denoting the velocity inversion.

the base of the pelagic sequence. Additionally, the Carnegie


Ridge is a physical barrier controlling the circulation of the
marine currents that flow along the trench affecting the
sediment deposition and erosion (Lonsdale, 1976).
3. Data
During the SISTEUR survey, some ~ 6000 km of deep
MCS lines were collected using a 45-L airgun source tuned in
a single-bubble mode, and a 360-channel, 4.5 km-long
streamer (Collot et al., 2002). Airgun shots were fired every
50 m to provide 45-fold common depth point data. Here we
use lines SIS-18 and SIS-72, which extend perpendicularly to
the southern Ecuadorian trench, north and south of the GFZ
respectively (Fig. 1). These lines were processed up to
Prestack Depth Migration (PSDM) to obtain depth-images
(e.g. Fig. 2A) and accurate 2D seismic velocity model (e.g.
Fig. 2B).

4. PSDM and velocity model uncertainties


This study focuses on the frontal margin, beneath the lower
slope, which is a complex target for seismic imaging because it
usually shows strong lateral and vertical velocity variations. To
overcome difficulties related to the inherent complexity of the
medium, we performed PSDM (Yilmaz, 2001), which is at
present day the most effective method to accurately image the
subsoil considering its spatial heterogeneities (Guo and Fagin,
2002). We used the SIRIUS-2.0 software package (GX
Technology) to perform PSDM based on the Kirchhoff
algorithm. This method allows building 2D velocity models
by focus analysis and iterative velocity picking. The resolution
and accuracy of the obtained velocity model depend basically
on 1) the quality and dip of reflections, and 2) the maximum
sourcereceiver offset/depth ratio and the dominant source
frequency (Lines, 1993; Ross, 1994). Bartolom et al (2005)
showed that coherent velocity changes for depth-focusing

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

457

Fig. 3. Line SIS-18. A) PSDM image of the front of the margin. B) Geological interpretation. C) 2D velocity model. Note that velocity along the subduction channel is
slightly lower than that of the overlaying upper-plate basement.

analysis can be observed until ~ 15 km depth, when using the


same seismic source and acquisition system that we use here.
Consequently, this depth can be considered as a sort of limit for
reliable velocity modeling. In lines SIS-72 and SIS-18 the
deepest reflections outlining the subduction channel are strong
and continuous, facilitating the application of the PSDM
method up to ~ 7 km depth (e.g. Figs. 2A and 3A). To further
verify the accuracy of the obtained velocity model and constrain
the errors, we performed an uncertainty test using the Velocity
Scan module of SIRIUS-2.0. This module generates a set of
velocity profiles by increasingly perturbing the reference model
within the region of interest. A total of 16 velocity models, with
a 5% step with respect to the reference velocity model, were
used to generate the corresponding depth-migrated images and
corrected Common Reflection Point (CRP) gathers. CRP
gathers enable estimating gather flattening that affects in turn
the quality of depth sections. The comparison of the images
obtained with the different perturbed velocity models allows
estimating the range of accuracy of the velocity models

considered. The results indicate that, depending on the segment


of the profile, velocity perturbed by more than 510% of the
original velocity do not flatten CRP gather reflections and
reduce considerably the quality of migrated images. Thus, the
maximum velocity uncertainties range between 85 m/s at the
trench axis, where the average SC velocity is ~ 1700 m/s, to less
than 200 m/s at a ~ 5 km bsf and ~ 32 km landward from the
trench, where the velocity is ~ 4000 m/s. This velocity
uncertainty is small considering the depth of the target, and
supports the reliability of the velocity inversion at the SC,
serving as a base to depth-migrate seismic data and to associate
uncertainty bounds to the velocity-related physical properties of
SC material. An ultimate velocity validation was done when
Agudelo (2005) performed a new depth migration of line SIS72 using the ray+Born pre-strack depth migration/inversion
method (Thierry et al., 1999). He obtained a similar velocity
distribution, particularly along the SC, and equivalent depth
images that corroborate the accuracy of the velocity model used
in this work.

458

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

5. Imaging the margin structure and subduction channel


5.1. Line SIS-72
This line images a small frontal prism, the overriding-plate
basement and slope sediment, and a well-defined subduction
channel (Fig. 2A and B). Fig. 2A is remarkable because it clearly
shows the SC extending from the trench to ~ 30 km landward, in
an erosional margin.
At the toe of the margin, a ~ 8-km-wide and ~ 1 km-thick
frontal prism is formed by three main landward-dipping imbricated thrust sheets abutting at the apex of the overriding-plate
basement. The top of the basement is associated to a strong but
discontinuous seaward-dipping reflection that underlies the
stratified reflections of the slope sediment. The overriding-plate
basement is cut by east-dipping reflections probably related to
ancient faults.
The deformation front is marked by a thin reflection of the
current thrust separating the frontal-prism sediments and the
horizontal reflections of trench infill. This reflection continues
beneath the frontal prism and branches landward to a thick,
strong and continuous reflection interpreted as the dcollement
(Fig. 2A and C). The dcollement reflection undulates and dips
to the East marking the top of the subduction channel and
bounding the underside of the overriding-plate basement. It is
called TSC (Top of the subduction channel) reflection from here
on. Seaward dipping reflectors at the bottom part of the upperplate basement appear to be truncated by the TSC, supporting
basal tectonic erosion. P-wave seismic velocity (Vp) of the
upper-plate basement backstop (3.53.8 km/s) contrasts with
1.82.7 km/s of the frontal prism (Fig. 2B and C), which results
from off-scraping sediment of the 0.6 km-thick wedge of wellstratified, flat-laying trench infill deposits. These deposits are
interpreted as turbidites, due to its proximity to the discharge
zone of submarine canyons, and by analogy with sand turbiditic
deposits off the northern margin (Collot et al., 2005). These
turbidites show a Vp = 1.8 km/s and lie unconformably over an
east-dipping, 0.20.4-km-thick layer with stratify to transparent
facies of pelagic to hemipelagic sediment that drapes the
Paleogene oceanic crust (Fig. 2).
The top of the oceanic crust is marked by a clear reflection
(TOC, Top of the Oceanic Crust reflection) that extends under
the margin marking the base of the SC (Fig. 2A). The TOC
shows 45 km-wide, 0.40.5 km-relief highs and lows
representing subducting horsts and grabens, suggesting the
vertical reactivation of oceanic crust normal faults (e.g. Ranero
et al., 2003). Fig. 1 shows some of these east-dipping normal
faults in the southern flank of the Carnegie Ridge.
The SC, bounded at top and bottom by the TSC and the
TOC, dips ~ 4 landward for about 32 km from the trench, and
down to a depth of ~ 5 km bsf. Its thickness varies from less
than 0.2 km beneath the frontal prism, to ~ 0.80.9 km on most
of the section before thinning to only ~ 0.20.3 km between
29 km to ~ 32 km from the trench. The origin of these
thickness variations is unconstrained, but we suggest that they
may arise from changes in sediment input from the submarine
canyons located nearby.

The SC shows a well-constrained low velocity (2.72.8 km/s)


that produces a velocity inversion when compared with the higher
velocities (3.53.8 km/s) of the overlaying upper-plate basement
(Fig. 2C). This velocity inversion results from the contrast between
the basement density and the low density of unconsolidated subducting material, and it use to be accompanied by a phase inversion of seismic data polarity (e.g. Bangs et al., 1999). Over
pressured fluids are frequently related with these phenomena, but it
is not always the case (e.g. Shipley et al., 1990). Nevertheless, in
our data, the TSC reflection shows discontinuous polarity
inversion and changing amplitude that could be associated with
variations in water content, as suggested by Bangs et al. (2004).
5.2. Line SIS-18
Line SIS-18, located north of the GFZ-trench intersection
(Fig. 1), shows different structures and seismic facies when
compared with line SIS-72. Line SIS-18 characterizes an
erosional margin, fronted by a ~ 3-km-wide sediment prism with
little internal structure suggesting slope debris accumulated at
the deformation front (Fig. 3A and B).
The flat-bottom trench is underlain by a graben formed by a
normal fault probably caused by plate bending. There, the trench is
twice narrower and shallower than in line SIS-72. Its ~0.3 kmthick infill deposit mainly shows structure-less seismic facies, and
an average Vp of 2.02.2 km/s (Fig. 3C). The Neogene oceanic
crust is covered by a thin layer of pelagic to hemipelagic sediments
b0.2 km that is hardly visible at the base of the trench. The chaotic
deposits filling the trench may be related with a semicircular scar
observed in the inner trench slope bathymetry and with an abrupt
seaward thinning and truncation of the ~0.5 km-thick slope
sediment near CDP 9200, to support mass wasting of the lower
slope as the trench infill mechanism (Fig. 3A).
Similarly to line SIS-72, the TOC and TSC reflections delineate
respectively the top of the oceanic crust and the dcollement, thus
clearly defining the SC. The overriding-plate basement is
internally poorly reflective with a few faint reflections dipping
either landward or seaward, occasionally parallel to the TSC. The
SC dips landward at ~6 and is thinner than in line SIS-72, varying
from ~0.2 km beneath the frontal prism to a maximum of ~0.6 km
landward. Here, changes in the slope sediment thickness indicate
that mass wasting may cause changes in trench sediment input.
The SC shows higher velocity than on line SIS-72, and a weak Vp
inversion of ~200 m/s with respect to the overriding-plate
basement velocity, which is within the Vp uncertainty range
(Fig. 3C). The different range of velocity within the SC between
lines SIS-18 and SIS-72 suggest a different nature of underthrust
material, and agrees with the observation of seismic images.
6. Quantifying fluid pressure within the SC
Estimating physical properties of the SC material to quantify
fluid pressure variations requires: 1) to infer the nature of underthursting sediment based on the geological setting and coring
information, 2) to use PSDM seismic velocity (Vp) along the SC
and sediment composition to infer porosity () based on existing
Vp- relationships, 3) to calculate fluid content and fluid flow

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

459

Table 1
Initial porosity (0), shale fraction (fsh) range, average fsh and b value of lines
SIS-72 and SIS-18 used for pressure estimation
Line

fsh range

Average fsh

b
(MPa 1)

SIS-72
SIS-18

0.45
0.30

0.300.50
0.200.30

0.40
0.25

0.052
0.04

Critical porosity (c) is 0.31 as considered in the normal compaction equation of


Erickson and Jarrard (1998).

along the different segments of the SC, and 4) to calculate the fluid
pressure and its variations within the SC. For these calculations,
our first assumption was to consider the SC as a zone decoupled
from the overriding-plate stress regime by the dcollement (Le
Pichon et al., 1993). This decoupling, shown by drilling and MCS
data in other convergent margins (e.g. Westbrook et al., 1982;
Taira et al., 1992), implies that below the dcollement (i.e. within
the SC), the horizontal stress resulting from plate convergence is
largely exceeded by the vertical stress from lithostatic pressure,
and in consequence it can be neglected in the calculations.
We based on a portion of PSDM velocity models of lines SIS72 and SIS-18 to obtain the average velocity value within the SC
from the trench to ~30 and ~16 km landward (Figs. 2C and 3C).
Velocity and velocity-derived physical properties were calculated
at 100 m interval and subsequently smoothed by applying a 1 kmwide Gaussian-type sliding window to filter high-frequency
artifacts. Then, the first step was to calculate porosity using the
Vp- empirical relationship for normally compacted watersaturated siliciclastic sediments from Erickson and Jarrard (1998).
We use this relationship because it accurately reproduces the Vp of a large number of sediment samples with lithologies ranging
from sand to shale acquired in different geodynamic settings (e.g.
Erickson and Jarrard, 1998; Gettemy and Tobin, 2003). This
equation considers sediment compaction without shortening, and
takes into account the critical porosity (o) and the shale fraction
(fsh) of incoming sediments. Nur et al. (1998) define the o as the
transition from the suspension domain to the consolidate rock
domain. The suspension domain, for high-porosity rocks, corresponds to a media that is unconsolidated and mechanically fluidsupported. In contrast, the consolidated domain, for low-porosity
rocks, corresponds to a media that shows a continuous framesupported matrix. It is important to note that the Vp- relationship
behaves differently in each domain: when N o, velocity and

Fig. 4. Cartoon showing the evolution of porosity () and thickness (h) of a


subducting element from point A (at the trench) to point B (below the margin)
separated by a distance d.

Fig. 5. Estimated porosity obtained from velocity considering a 0 = 0.31 and


fsh = 20, 40, 60 and 80.

porosity are not strongly dependent, but when b o, velocity


strongly depends on porosity and increases significantly with a
small decrease in porosity (Raymer et al., 1980). o is typical for
each kind of porous material. For normal compaction Erickson
and Jarrard's (1998) equation, o is 0.31. fsh is the shale/sandstone content ratio. In this case, the lack of available trench cores
made us to infer fsh for both lines based on the geological setting
and the seismic facies observed in the sections. As we stated in the
previous section, line SIS-72 shows a thick sequence of turbiditic
deposits in the trench, while line SIS-18 shows mass-wasting
material, probably constituted by coarser grains and rock fragments. Based on these observations, we considered a mid fsh,
ranging between 0.30 and 0.50 for the incoming SC material in
line SIS-72, and low fsh between 0.200.30 in SIS-18 (Table 1).
Given that these estimations of fsh values can be considered as
somewhat arbitrary in absence of trench sediment samplings, we
have made a test of the influence of fsh on the obtained porosity.
Fig. 5 illustrate that fsh has little influence in Vp- transformation.
It is significant only for high pressures, but the form of the curve is
identical in the different segments. This indicates that fsh values
slightly higher or lower to the value considered would not affect
the overall sediment behavior, the only visible effect being the
different dewatering fluxes.
In the second step, we calculate the amount of water confined
within the SC assuming that sediment pores are water-saturated.
The volume of water (C) contained in a column of section (S) is
Table 2
Acronyms of parameters and calculated physical properties
Vp

o
i
fsh
C
C
Ct
Plit
Pe
Pf
Phyd
P

b
bsh
bss
P

Seismic P-wave velocity


Porosity
Critical porosity
Initial porosity
Shale fraction
Water content
Volume of water expelled
Volume of water expelled per unit of time (fluid flow)
Lithostatic pressure (equivalent to confining pressure)
Effective stress
Fluid pressure or pore pressure
Hydrostatic pressure
Fluid overpressure or excess pore pressure
Fluid overpressure ratio
Compressibility constant
Compressibility constant for shale
Compressibility constant for sandstone
Crushing strength

460

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

the product of the sediment porosity in the column times the SC


thickness (h) times the section:
C Uhs

Then we estimated the amount of water contained in sediment


pores that is likely to be expelled from the system by compaction
as subduction proceeds. The differential thickness (h) between
a reference column of null porosity (h0) and another of porosity
(h) can be expressed as:
h h0 Dh

where h can be expressed in terms of porosity as:


Dh Uh

Considering Eqs. (2) and (3):


h h0 =1  U

461

points, since we do not compare different elements, but the same


element in different locations.
Water content expressed in terms of time is:
DCt uDC=d

where Ct is the volume of water expelled per unit of time


(fluid flow), u is the plate convergence rate (55 km/m.y.), and d
is distance between A and B.
Finally, we estimated fluid pressures along the SC. The total
pressure supported by one point at the SC is the addition of
lithostatic and tectonic pressures. Assuming that tectonic shortening is negligible within the SC, the total pressure corresponds to
lithostatic pressure (Plit, i.e., confining pressure), which is compensated by the effective stress exerted by the solid fraction of
subducted material (Pe), and the pressure exerted by fluid within
pores (fluid pressure or pore pressure, Pf):
Plit Pe Pf :

Let us then consider an element of SC with 1 km2 section


downgoing into subduction. Its initial state at point A, is characterized by hA, A, and the final one, at its consecutive 100 m
distant point B, by hB, B (Fig. 4). From (4) hB will be:

In our calculations, Plit includes the weights of the water and


overriding-plate sediment and basement columns. To transform
velocity into density for the slope-sediment layer, we use the
Gardner et al.'s (1974) relationship:

hB hA 1  UA =1  UB

qsed 0:23 Vp0:25

Assuming that the pressure and temperature conditions within


the SC, as well as the permeability of the overriding-plate and the
nature of the downgoing sediments remain basically invariable
with time at each point along the SC, sediments transported at a
given depth beneath the margin will always show the physical
properties (Vp, , etc) characteristics of that particular point. The
volume of water that would be potentially expelled by a given
element when moving between two points A and B can be
expressed as the difference in water content between these two
points (C = CA CB), which is:
CA UA hA

6a

CB UB hA 1  UA =1  UB

6b

and
DC hA UA  UB =1  UB :

Note that Eq. (7) expresses C as the difference in fluid


content between an element of SC located in A, where it has a
porosity A and a volume S hA, and the same element in B,
where it would have a porosity B and a volume S hA(1 A)/
(1 B), this is, its thickness in A corrected for the porosity
change between A and B. This means that the estimation of C
does not depend on the thickness variation between the two

10

whereas, for the upper-plate basement, we used the NafeDrake


curve modified by Barton (1986) for the continental rocks:
qbas 1:724 0:168Vp

11

which is valid for Vp b 5.5 km/s.


Effective stress (Pe) was estimated using a modification of
the so-called Athy's Law (Athy, 1930) proposed by Le Pichon
et al. (1993). This modification consists on linking porosity with
Pe instead of depth, which was the variable originally used in
Athy's equation:
U Ui ebPe ;

12

where i is the initial porosity and b the compressibility constant.


We consider as i the average velocity-derived porosity of
the 50 shallowest meters of the incoming sediment column at
the trench, and a compressibility constant, b, varying linearly
between bsh = 0.1 MPa 1 for fsh = 1, and bss = 0.02 MPa 1 for
fsh = 0 (Le Pichon et al., 1993). Interpolated b values obtained
for lines SIS-72 and SIS-18, according to different fsh values are
listed in Table 1. Thus, we calculated Pe using Eq. (12) and then
Pf using Eq. (9). Then, when Pf exceeds the hydrostatic
pressure (Phyd) we got the fluid overpressure (or excess pore
pressure, P) along the SC, P = Pf Phyd.

Fig. 6. Line SIS-72. A) 2D velocity model. B) Average Vp velocity along the SC calculated along the white dashed-line in A. The grey wide band indicates velocity
uncertainties. Thin dashed black line indicates the position of the deformation front. Thick dashed grey lines indicate limits between Zones I, II and III. C) Average
porosity and uncertainties. D) Fluid flow and water content along the subduction channel. E) Confining pressure F) Effective pressure (Pe) for fsh = 0.40, 0 = 0.45,
b = 0.052, and fluid overpressure (P). G) Fluid overpressure radio (). Color scale represents the distance from the deformation front to each calculation point.

462

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

Finally, a useful parameter to evaluate the drainage conditions


in the different segments of the SC is the fluid overpressure ratio,
= (Pf Phyd)/(Plit Phyd), which corresponds to the ratio between fluid overpressure and overburden pressure. When = 1,
P equals the overburden pressure, reflecting undrained
conditions or total fluid retention (the media is impermeable).
Conversely, when = 0, P = 0 indicating optimal drainage
conditions for fluid evacuation (the media is totally permeable).
Table 2 shows a compilation of acronyms of the parameters
considered for calculations.
7. Physical properties of the subduction channel
Based on Vp variations, we can divide the first tens of
kilometers of the SC along lines SIS-72 and SIS-18 in three
zones (referred hereafter to as Zones I, II and III), showing
contrasting Vp-derived physical properties.
7.1. Line SIS-72
Zone I extends over the first ~9 km of the SC, and shows a
rapid Vp increase, from 1.8 km/s at the deformation front to
2.6 km/s, ~9 km landward (Fig. 6A, B). The corresponding Vpderived porosity drops from ~0.50 to ~0.25, and it reaches the
critical porosity at ~4 km from the deformation front (Fig. 6C).
This Vp increase, and reduction denote progressive compaction
and subsequent dewatering of underthrust material as a response
to the increasing load of the overriding-plate. The estimated
amount of fluids expelled in Zone I, calculated by integrating the
fluid flow along the whole segment, is ~12 l * yr 1 * m 2
(Fig. 6D). The gradual landward thickening and consequent
load increase of the frontal prism is reflected by the steady growth
of Plit from ~45 MPa at the trench, to ~70 MPa at 9 km from the
deformation front (~2 km bsf) (Fig. 6E). The porosity reduction
makes Pe to increase similarly to Plit, indicating that a significant
part of the load of the overriding-plate is transferred to the solid
fraction of the underthrust material as fluids are expelled (Fig. 6F).
Fluid overpressure is lower than ~8 MPa in this zone (Fig. 6F)
and, consistently, remains inferior to ~0.5 indicating that fluid
drainage is efficient (Fig. 6G).
Zone II extends between ~9 km and ~25 km from the
deformation front. Here, the average Vp remains uniform around
2.8 km/s (Fig. 6B), as well as the corresponding , with values
around 0.28 (Fig. 6C). Therefore Pe is also uniform along the entire
Zone II (Fig. 6F), despite the continuous increment of Plit to ~110
120 MPa (Fig. 6E). The difference between Pe and Plit is
compensated by increasing Pf, as illustrated by the growth of P
from ~8 MPa to ~40 MPa at 25 km from the deformation front
(~4.5 km bsf) (Fig. 6F). This means that the load of the overridingplate in Zone II, is mainly supported by the pressure of pore fluids
trapped within the SC. This phenomenon agrees with the high
value of ~0.8 (Fig. 6G) that approaches to the fluid retention limit,

463

and suggests the presence of low permeable layers in this segment.


Our calculations indicate that the amount of fluids that is potentially expelled from subducted sediments in Zone II is one order of
magnitude smaller than in Zone I (Fig. 6D).
Zone III extends between ~25 km and ~32 km from the trench.
In this zone, Vp rapidly increases from ~2.8 km/s to ~4.0 km/s
(Fig. 6B), and drops from ~0.25 to less than 0.10 (Fig. 6C). The
calculated Pe grows from ~10 to ~28 MPa (from ~4.5 to 5.2 km
bsf), and the P slightly decreases from ~40 MPa to 35 MPa
regardless of the Plit increase to ~130 MPa (Fig. 5E, F). drops
from ~0.8 to ~0.6 (Fig. 6G), suggesting enhanced fluid expulsion
conditions following the rapid compaction of SC material. The
estimated amount of fluids expelled from the subducted sediments
in this zone is ~6 l * yr 1 * m 2 (Fig. 6D).
7.2. Line SIS-18
In line SIS-18, the analysis of physical properties within the SC
was done from the deformation front to ~16 km landward (~5 km
bsf) (Fig. 7). Although the average velocity is here higher than in
line SIS-72, the global velocity trend evolves similarly along all
zones in both lines. Whereas Vp increases along Zones I and III, it
remains practically constant along Zone II (Fig. 7A, B).
The higher Vp along line SIS-18 implies that must be
lower than in line SIS-72 (Fig. 7C). At the trench, i is ~ 0.30
0.32, indicating less pore-filling fluids entering the SC as
compared with line SIS-72, where i ~ 0.45 (Fig. 5C). In Zone I,
the estimated is higher than in the same zone of SIS-72,
suggesting less efficient drainage conditions (Fig. 7G). The
estimated fluid flow for Zone I is 4 l * yr 1 * m 2 only.
In Zone II, a constant ~ 0.25 is obtained. The estimated
fluid flow along this zone is one order of magnitude smaller than
in Zone I, indicating a low permeability zone. values are
close to 0.5 but increase to 0.75 over a short distance of ~ 3 km,
pointing to strong spatial variations in permeability conditions
and drainage.
Finally, in Zone III, porosity drops to 0.10 and shows a
marked decrease to values lower than 0.5, supporting a notable
increase of Pe arising from rapid compaction and enhanced fluid
expulsion (Fig. 7F).
8. Discussion
The quantification of physical properties of underthrust material
allowed differentiating three zones along the first ~30 km of the
SC. Each zone is directly related to non-linear variations in Vp, ,
and Pf that suggest the presence of discrete steps of the mechanical
behavior of underthrust material as it is buried beneath the margin
with lithostatic pressure increasing from ~0 to 130 MPa.
In the next paragraphs we compare the results obtained in
Zones I, II, and III in lines SIS-18 and SIS-72, and relate them
with deformation process at grain scale in order to discuss the

Fig. 7. Line SIS-18. A) 2D velocity model. B) Average velocity along the SC. The grey wide band indicates velocity uncertainties. Thin dashed black line indicates the
position of the deformation front. Thick dashed grey lines indicate limits between Zones I, II and III. The grey band corresponds to uncertainties. C) Average porosity.
D) Fluid flow and water content along the subduction channel. E) Confining pressure. F) Effective pressure (Pe) for fsh = 0.25 for fsh = 0.25, 0 = 0.30, b = 0.04, and
fluid overpressure (P). G) Fluid overpressure radio (). Colour scale represents the distance from the deformation front to each calculation point.

464

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

evolution of the mechanical processes governing transformations of underthrust material and its influence on the stress state
of the inter-plate contact and the margin deformation.
8.1. Zone I
In both profiles, this zone is characterized by increasing Vp
and reducing , indicating an overall progressive consolidation
of underthrust sediment, in response to the increasing load of
the overriding-plate. At the same time, an efficient drainage of a
permeable media results in a rapid expel of pore fluid, particularly before the critical porosity is reached. Laboratory experiments indicate that at low effective pressures, b20 MPa,
granular deformation is dominated by mechanisms like particle
rotation and frictional slip at grain contacts, with no or very little
grain damage (Karner et al., 2003). These mechanisms induce
granular rearrangements that increase the population of grains in
contact until the system attains an optimal grain packaging,
which primarily depends on its grain size and is likely to occur
near 0.
Pore-fluids may be expelled seaward by lateral migration
along high-permeability stratigraphic layers (Saffer et al., 2000;
Moore and Vrolijk, 1992); upward migration to the inter-plate
contact and further seaward migration through the dcollement
zone (e.g. Moore, 1989; Saffer and Screaton, 2003), or upward
migration across the dcollement and the overriding-plate or the
active thrusts of the frontal prism (e.g. Cloos, 1984; Brown and
Westbrook, 1987; Screaton and Saffer, 2005). In both lines, the
high reflectivity of the dcollement and the landward-dipping
reflections of the small prism in line SIS-72 are probably explained by the presence of fluids, suggesting they act as seaward
fluid escape conduits.
8.2. Zone II
In this zone, drops below the o to 0.250.28 (Figs. 6C
and 7C), indicating that underthrust sediment have attained its
optimal packaging and behave now as a consolidate rock.
Steady values of porosity regardless of the increasing Plit reflect
a strong deceleration of sediment dewatering and consolidation
rate. The growing fluid pressure that compensates the increasing
load of the upper-plate, with a lower influence on effective
pressure and concomitant porosity reduction, explains this
behavior. Grain deformation mechanisms must be similar to
those dominating at the end of Zone I, as effective pressure
conditions remain practically the same. Undrained conditions
along this zone are corroborated by high ~ 0.8 that indicates a
lost of permeability form Zones I to II. This results are
consistent with those obtained with numerical modeling of
Screaton (2006), showing a similar evolution of porosity, excess
pore pressure and for underthrust sediments of a nearly nonaccretionary margin at distances from the trench that correspond
well with those of our Zones I and II.
The low permeability of Zone II could be either related to the
intrinsic nature of the underthrust sediments (e.g. Bryant et al.,
1975; Screaton and Saffer, 2005), to a different physical behavior
of subducted sediments (e.g. Kimura et al., 2007), to the

mechanical characteristics of the dcollement (e.g. Tobin et al.,


2001), or to the overriding-plate basement acting as an impermeable barrier. This last case seems to be less probable as Sage et al.
(2006) showed high fluid over-pressures along the SC in the
central Ecuadorian margin, regardless of a pervasive normalfaulted basement.
8.3. Zone III (and deeper)
This zone is characterized by a drop of porosity to ~ 0.10
0.05, together with an increase of the effective pressure, and the
subsequent decrease in fluid overpressure and that indicates a
rapid expel of over-pressured fluids out from the system. In this
case, the system must be better drained, and the progressively
increasing load of the overriding-plate (Plit from ~ 70 to
110 MPa) is mainly transferred into increasing effective pressure
(up to ~ 30 MPa). Experimental results show that for increasing
Pe between 30 MPa and 100 MPa, compaction also increases
responding mainly to elastic deformation of grains (i.e. Hertzian
deformation) (Terzaghi, 1925).
The over-pressured fluids may either escape seaward along the
dcollement, or, more likely, they may migrate upward through
the overriding-plate. In the latter case, fluids may reach the
landward-dipping faults of the overriding-plate basement, imaged
in the seismic records (Fig. 2A, B), and migrate up to the seafloor.
The presence of fluids would also enhance reflectivity for seismic
imaging. If this is the case, seeps of fluids would be expected at
the seafloor near CDP 11000 in profile SIS-72 (Fig. 2A), as it was
observed offshore the Pacific margin of Costa Rica (Hensen et al.,
2004). It has been also described that fluid over-pressures itself
may enhance the breakage of the permeable barrier (Sibson, 1981)
and thus, the released over-pressured fluids have the potential to
undermine the underside of the overriding plate by means of
hydrofracturation. This mechanism may enhance basal erosion as
observed in other margins (i.e. Muruachi and Ludwig, 1980;
Ranero and von Huene, 2000; von Huene et al., 2004).
Deeper than Zone III, the granular media continues
compaction and it densifies (Karner et al., 2003). The grain
population reaching the failure threshold gradually increases.
This process eventually leads to the macroscopic yield stage
where grain cracking becomes pervasive (i.e., cataclasis),
compactive strain rates accelerate, and the system catastrophically collapses (e.g., Wong et al., 1997). Reported crushing
strengths (P) are quite variable (e.g. Zoback and Byerlee,
1976), and depend on the loading style and it decreases with
increasing grain size and increasing porosity (Zhang et al.,
1990, Wong et al., 1997). One possibility is that the higher
pressures at which the different transformational steps occur in
line SIS-72 in comparison to line SIS-18 could be explained by
a larger grain size of subducted sediments in SIS-18, which
agree in turn with observations suggesting turbidites and
pelagic-hemipelagic deposits of the oceanic crust as potential
source of material feeding the SC in line SIS-72, and coarse
mass-wasting deposits eroded from the margin's front in line
SIS-18.
Additionally, diagenetic processes like pressure solution start
to be important enhancing consolidation and lithification of the

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

underthrust material. These phenomena could be the next


deformation stage, ahead of the 30 first kilometers from the
deformation front, where the aseismicseismic transition zone is
located. An example is the Mugi tectonic Melange (Onishi and
Kimura, 1995; Kitamura et al., 2005) which has recorded the
processes of underthrusting to underplating under temperature
conditions of 120220 C and depth of 67 km depth, similar to
the conditions near the updip limit of the Seismogenic Zone
(Ikesawa et al., 2005; Matsumara et al., in press). In our case, a
simple thermal model based on the plate geometry derived from
PSDM line SIS-72, and assuming a 34 My age for the downgoing
plate and thermal parameters from Marcaillou et al. (2006)
predicts that temperature in Zone III of SIS-72 is close to ~80 C
(Calahorrano, 2005). Such a temperature range, associated with
lithostatic pressure of 110 to 130 MPa, and effective pressure
~30 MPa, corroborates our interpretation of the dominant
deformation processes affecting underthrust sediments up to
67 km depth, and suggest that Zone III is practically preceding
the theoretical emplacement of the first diagenetic and low-grade
metamorphic processes associated to the aseismic/seismic limit
of the subduction megathrust.
9. Conclusion
We have quantified the along-strike and across-strike
variations of physical properties of incoming sediments along
the first ~ 30 km of subduction, based on two MCS lines
acquired north (SIS-18) and south (SIS-72) of the GFZ, off the
erosional southern Ecuadorian margin. The analysis includes
the estimation of PSDM velocity and uncertainty within the SC
and its propagation to other velocity-derived parameters such as
porosity, fluid or effective pressures.
Results indicate a similar evolution of the behaviour of SC
material in both profiles, with increasing confining pressure. We
identified three zones based on the non-linear variations of
physical properties: Zone I (09 km from the deformation front in
SIS-72, 05 km in SIS-18) shows a progressive increase of
seismic velocity and porosity reduction, related to sediment
compaction and effective fluid drainage along the dcollement
and thrust faults of the frontal prism. Zone II (925 km in SIS-72,
512 km in SIS-18) displays remarkably uniform velocity and
porosity regardless of the increasing confining pressure, indicating undrained conditions and thus progressively landward-increasing fluid over-pressures within the SC. Finally, Zone III (25
30 km in SIS-72, 1216 km in SIS-18) is characterized by a
sudden velocity increase and porosity decrease suggesting a
sudden release of over-pressured fluids.
Changing physical properties estimated along Zones I, II and III
are in agreement with experimental results for compaction and
grain deformation at different effective pressure ranges. At low
pressures, the dominant processes are granular flow and re-packing
(Zone I), at mid pressures, compaction stops owing to fluid retention that makes fluid overpressure to support the progressively
increasing margin load (Zone II), whereas at high pressures, once
over-pressured fluids are released, compaction starts to be
governed by Hertzian-like, elastic deformation (Zone III). This
must be the dominant process until the grains attain their crushing

465

strength, leading to pervasive granular cataclasis and, eventually,


to the collapse of the system.
Although lines SIS-72 and SIS-18 are only 60 km apart, their
SC shows different thicknesses, seismic characters and Vp that
are related to a variable nature and volume of sedimentary input.
We suggest that the smaller grain size of the sediments entering
the subduction zone in line SIS-72, as indicated by the seismic
character of trench sediments, may explain the higher pressures
at which the different transformations occur in this line with
respect to SIS-18.
We suggest that the sudden fluid released in Zone III may
induce hydrofracturing favoring basal erosion, and probably
install the adequate conditions for the beginning of earthquake
generation, near the updip limit of the seismogenic zone.
Acknowledgements
This work is part of the Ph. D. research (UPMC) of A.
Calahorrano B., supported by a grant of the Institut de Recherche
pour le Dveloppement-IRD. A. Calahorrano has been suported
by the Juan de la Cierva Program of the Spanish Ministry of
Education and Science during the writing of this paper.
SISTEUR project was funded by the French institutes IRD,
CNRS, IFREMER, UPMC. Computing and seismic processing
facilities were supported by Geosciences Azur and IFMGEOMAR using the Large Scale Facilities of the European
program Improving Human Potential. C. R. Ranero, V. Sallars
and A. Calahorrano B. are members of the Barcelona Center for
Subsurface Imaging (Barcelona CSI) that is supported by the
KALEIDOSCOPE project from REPSOL-YPF. This is a
contribution of the UMR Geosciences Azur 6526, which is
part of the Observatoire de la Cte d'Azur. A. Calahorrano B. is
grateful to co-authors for fruitful discussions that help to
significantly improve the initial manuscript and to an anonymous reviewer for its careful and critical reviews.
References
Agudelo, W., 2005. Imagerie sismique quantitative de la marge convergente
dEquateur-Colombie: application des mthodes tomographiques aux donnes
de sismique rflexion multitrace et rfraction-rflexion grand-angle des
campagnes SISTEUR et SALIERI. Univerit Pierre et Marie Curie, 203 pp.
Athy, L.F., 1930. Density, porosity and compaction sedimentary rocks. Am.
Assoc. Pet. Geol. Bull. 14, 124.
Bangs, N.L., Westbrook, G.K., Ladd, J.W., Buhl, P., 1990. Seismic velocities
from the Barbados Ridge complex: indicator of high pore pressures in an
accretionnary complex. J. Geophys. Res. 95, 87678782.
Bangs, N., Shipley, T.H., Moore, J.C., Moore, G.F., 1999. Fluid accumulation and
channeling along the northern Barbados Ridge dcollement thrust. J. Geophys.
Res. 104, 2039920414.
Bangs, N., Shipley, T.H., Gulik, S., Moore, G.F., Kuromoto, S., 2004. Evolution
of the Nanakai Trough dcollement from the trench into the seismogenic
zone: Inferences from three dimensional seismic reflection imaging. Geology
32 (4), 273276.
Barckhausen, U., Ranero, C.R., von Huene, R., Cande, S.C., Roeser, H.A., 2001.
Revised tectonic boundaries in the Cocos Plate off Costa-Rica: implications
for the segmentation of the convergent margin and for plate tectonic models.
J. Geophys. Res. 106, 1920719220.
Bartolom, R., Contrucci, I., Nouz, H., Thiebot, E., Klingelhoefer, F., 2005.
Using the OBS wide-angle reflexion/refraction velocities to perform pre-

466

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467

stack depth migration image of the single bubble multichannel seismic:


example of the Moroccan margin. J. Appl. Geophys. 57, 107118.
Barton, P.J., 1986. The relationship between seismic velocity and density in the
continental crust- a useful constraint? Geophys. J. R. Astron. Soc. 87, 195208.
Brown, K., Westbrook, G., 1987. The tectonic fabric of the Barbados Ridge
accretionary complex. Mar. Pet. Geol. 4, 7181.
Bryant, W.R., Hottman, W., Trabant, P., 1975. Permeability of unconsolidated
and consolodated marine sediments, Gulf of Mexico. Mar. Geotechnol. 1,
114.
Bolton, A.J., Vannucchi, P., Clennell, M.B., Maltman, A., 2000. Microstructural
and geomechanical constraints on fluid flow at the Costa Rica convergent
margin, Ocean Drilling Program Leg 170. Proc. ODP, Sci. Results, 170:
College Station, TX (Ocean Drilling Program). doi:10.2973/odp.proc.
sr.170.007.2000.
Calahorrano B., A. 2005. Structure de la Marge du Golfe de Guayaquil (Equateur)
et proprits physiques du chenal de subduction, partir des donns de
sismique marine, rflexion et rfraction. Ph.D document, Univerist Pierre et
Marie Curie-Paris 6, 221p.
Cloos, M., 1984. Landward dipping reflectors in accretionary wedges: active
dewatering conduits? Geology 12, 519522.
Cloos, M., Shreve, R.L., 1988. Subduction-channel model of prism accretion,
melange formation, sediment subduction, and subduction erosion at convergent plate margins, 1, Background and description. Pure Appl. Geophys.
128, 455500.
Clowes, R.M., Brandon, M.T., Green, A.G., Yorath, C.J., Sutherland-Brown, A.,
Kanasewich, E.R., Spencer, C.S., 1987. LITHOPROBE southern Vancouver
Island: Cenozoic subduction complex imaged by deep seismic reflections.
Can. J. Earth Sci. 24, 3151.
Cochrane, G.R., Moore, J.C., MacKay, M.E., Moore, G.F., 1994. Velocity and
inferred porosity model of the Oregon accretionary prism from multichannel
seismic reflection data: Implications on sediment dewatering and overpressure. J. Geophys. Res. 99 (B4), 70337043.
Collot, J.-Y., Charvis, P., Gutscher, M.A., Operto, S., et al., 2002. Exploring the
Ecuador-Colombia active margin and interplate seismogenic zone. EOS
Trans., Amer. Geophys. Union 83 (17), 189190.
Collot, J.-Y., Migeon, S., Spence, G., Legonidec, Y., Marcaillou, B., Schneider,
J.L., Michaud, F., Alvarado, A., Lebrun, J.F., Sosson, M., Pazmino, A.,
2005. Seafloor margin map helps in understanding subduction earthquakes.
EOS Trans., Amer. Geophys. Union 86 (46), 463465.
Davis, E.E., Hyndman, R.D., 1989. Accretion and recent deformation of sediments along the northern Cascadia subduction zone. Geol. Soc. Amer. Bull.
101, 11531172.
Davis, D., Suppe, J., Dahlen, F.A., 1983. Mechanics of fold-and-thrust belts and
accretionary wedges. J. Geophys. Res. 88 (B2), 11531172.
Ego, F., Sbrier, M., Lavenu, A., Yepes, H., Egues, A., 1996. Quaternary state of
stress in the Northern Andes and the restraining bend model for the Ecuadorian Andes. Tectonophysics 259, 101116.
Erickson, S.N., Jarrard, R.D., 1998. Velocityporosity relationships for watersaturated siliciclastic sediments. J. Geophys. Res. 103 (B12), 3038530406.
Gardner, G.H.F., Gardner, L.W., Gregory, A.R., 1974. Formation velocity and
density the diagnostic basics for stratigraphic traps. Geophysics 39,
770780.
Gettemy, G.L., Tobin, H.J., 2003. Tectonic signatures in centimeter-scale velocity
porosity relationships of Costa Rica convergent margin. J. Geophys. Res. 108
(B10), 24942505.
Guo, N., Fagin, S., 2002. Becoming effective velocity-model builders and depth
imagers, Part 1 the basics of prestack depth migration. Lead. Edge
12051209.
Handschumacher, D.W., 1976. Post-Eocene plate tectonics of Eastern Pacific.
The Geophysics of the Pacific Basin and Its Margin: American Geophys.
Union Geophys. Mon., 19, pp. 799804.
Hensen, C., Wallmann, K., et al., 2004. Fluid expulsion related to mud extrusion
off Costa Rica a window to the subducting slab. Geology 32 (3), 201204.
Hey, R.N., 1977. Tectonic evolution of the CocosNazca spreading center.
Geol. Soc. Amer. Bull. 88, 14141420.
Hyndman, R.D., Yorath, C.Y., Clowes, R.M., Davis, E.E., 1990. The northern
Cascadia subduction zone at Vancouver Island: seismic structure and
tectonic. Can. J. Earth Sci. 27, 313329.

Hyndman, R.D., Wang, K., Yamano, M., 1995. Thermal constraints on the
seismogenic portion of the southwestern Japan subduction thrust. J.
Geophys. Res. 100, 1537315392.
Ikesawa, E., Kimura, G., Sato, K., Ikehara-Ohmori, K., Kitamura, Y.,
Yamaguchi, A., Ujiie, K., Kato, A., Hashimoto, Y., 2005. Cataclastic
involvement of basement basalts into plate boundary mlange: an asperity
breakage in seismogenic zone? Inference from on-land mlange in the
Shimanto Belt of eastern Shikoku, southwest Japan. Tectonophysics 401,
2172300.
Karner, S.L., Chester, F.M., Kronenberg, A.K., Chester, J.S., 2003. Subcritical
compaction and yielding of granular quartz sand. Tectonophysics 377, 357381.
Kastner, M., Elderfield, H., Martin, J.B., 1991. Fluids in convergent margins:
what do we know about their composition, origin, role in diagenesis and
importance for oceanic chemical fluxes? Philos. Trans. R. Soc. Lond. Ser. A:
Math. Phys. Sci. 335, 243259.
Kimura, G., Kitamura, Y., Hashimoto, Y., Yamagushi, A., Shibata, T., Ujiie, K.,
Okamoto, S., 2007. Transition of accretionary wedge structures around the
up-dip limit of the seismogenic subduction zone. Earth Planet. Sci. Lett. 255,
471484.
Kitamura, Y., Sato, K., Ikesawa, E., Ikehara-Ohmori, K., Kimura, G., Kondo,
H., Ujiie, K., Onishi, C.T., Kawabata, K., Hashimoto, Y., Mukoyoshi, H.,
Masago, H., 2005. Melange and its seismogenic roof dcollement: a plate
boundary fault rock in subduction zone: an example from the Shimanto Belt,
Japan. Tectonics 24.
Lallemand, S.E., Schnurle, P.S., Malavieille, J., 1994. Coulomb theory applied
to accretionary and non accretionary wedges: possible causes for tectonic
erosion and/or frontal accretion. J. Geophys. Res. 99, 12,03312,055.
Le Pichon, X., Henry, P., Lallemant, S., 1993. Accretion and erosion in subduction
zones: the role of fluids. Annu. Rev. Earth Planet. Sci. 21, 307331.
Lines, L., 1993. Ambiguity in analysis of velocity depth. Geophysics 58, 596597.
Lonsdale, P., 1976. Abyssal circulation of the Southern Pacific and some
geological implications. J. Geophys. Res. 81, 11631176.
Lonsdale, P., 1978. Ecuadorian subduction system. AAPG Bull. 62 (12),
24542477.
Lonsdale, P., Klitgord, K.D., 1978. Structure and tectonic history of the eastern
Panama Basin. Geol. Soc. Amer. Bull. 89, 981999.
Mascle, A., Moore, J.C., et al., 1988. Proc. ODP Initial Report 110. Ocean
Drilling Program, College Station TX, p. 63.
Marcaillou, B., Spence, G., Collot, J.-Y., Wang, K., 2006. Thermal regime
from bottom simulating reflectors along the N EcuadorS Colombia
margin: relation to margin segmentation and the XXth century great
subduction earthquakes. J. Geophys. Res. 111, B12407. doi:10.1029/
2005JB004239.
Matsumura, M., Hashimoto, Y., Kimura, G., Ohmori-Ikehara, K., Enjohji, M.,
Ikesawa, E., 2003. Depth of oceanic-crust underplating in a subduction
zone: Inferences from fluid-inclusion analyses of crack-seal veins.
Geology 31, 10051008.
McIntosh, K.D., Sen, M.K., 2000. Geophysical evidence for dewatering and
deformation processes in the ODP Leg 170 area offshore Costa Rica. Earth
Planet. Sci. Lett. 178, 125138.
Moore, J.C., 1989. Tectonics and hydrogeology of accretionnary prisms: role of
the dcollement zone. J. Struct. Geol. 11, 95106.
Moore, G.F., Shipley, T.H., 1988. Behavior of the dcollement at the toe of the
middle America trench. Geol. Rundsch. 77 (1), 275284.
Moore, J.C., Saffer, D.M., 2001. Updip limit of the seismogenic zone beneath
the accretionnary prism of southwest Japan: an effect of diagenetic to lowgrade metamorphic processes and increasing effective stress. Geology 29
(2), 183186.
Moore, J.C., Vrolijk, P., 1992. Fluids in accretionary prisms. Rev. Geophys. 30,
113135.
Moore, J.C., Shipboard Party ODP Leg156, 1995. Abnormal fluid pressures and
fault-zone dilatation in the Barbados accretionary prism: Evidence from
logging while drilling. Geology 23, 605608.
Moore, G.F., Taira, A., Klaus, A., et al., 2001. Proc. ODP, Initial Rep.190. Ocean
Drilling Program, College Station TX, p. 87.
Muruachi, S., Ludwig, W.J., 1980. Crustal structures of the Japan Trench: the
effect of subduction of oceanic crust. Initial Rep. Deep Sea Drill. Proj. 56
and 57, pp. 463469.

A. Calahorrano B. et al. / Earth and Planetary Science Letters 267 (2008) 453467
Nur, A., Mavko, G., et al., 1998. Critical porosity: a key to relating physical
properties to porosity in rocks. Lead. Edge 357362.
Onishi, C., Kimura, G., 1995. Melange fabric and relative convergence in
subduction zone. Tectonics 4, 12731289.
Ranero, C.R., von Huene, R., 2000. Subduction erosion along the Middle
America convergent margin. Nature 404.
Ranero, C.R., Phipps Morgan, J., McIntosh, K., Reichert, C., 2003. Bendingrelated faulting and mantle serpentanization at the Middle America trench.
Nature 425, 367373.
Raymer, L.L., Hunt, E.R., et al., 1980. An improved sonic transit time-toporosity transform. Trans. SPWLA Annu. Loggin Symp. 21st., pp. 113.
Ross, W.S., 1994. The velocitydepth ambiguity in seismic traveltime data.
Geophysics 59, 830843.
Saffer, D.M., Bekins, B.A., 1998. Episodic fluid flow in the Nankai accretionary
complex: timescale, geochemistry, flow rates, and fluid budget. J. Geophys.
Res. 103 (B12), 3035133370.
Saffer, D.M., Screaton, E.J., 2003. Fluid flow at the toe of convergent margins:
interpretation of sharp pore-water geochemical gradients. Earth Planet. Sci.
Lett. 213, 261270.
Saffer, D.M., Silver, E.A., et al., 2000. Inferred pore pressures at the Costa Rica
subduction zone: implications for dewatering processes. Earth Planet. Sci.
Lett. 177, 193207.
Sage, F., Collot, J.-Y., Ranero, C.R., 2006. Interplate patchiness and subduction
erosion mechanisms: evidence from depth-migrated seismic images at the
central Ecuador convergent margin. Geology 34 (12), 9971000. doi:10.1130/
G22790A.1.
Sallars, V., Charvis, P., 2003. Seismic constraints on the geodynamic evolution
of the Galapagos province. Earth Planet. Sci. Lett. 214, 545559.
Sallars, V., Ph. Charvis, E.R., Flueh, J., 2005. Bialas and the SALIERI
Scientific Party. Seismic structure of the Carnegie ridge and the nature of the
Galpagos hotspot. Geophys. J. Int., 161 (3), 763788. doi:10.1111/ j.1365246 X.2005.02592.x.
Scholl, D.W., Marlow, M.S., Cooper, A.K., 1977. Sediment subduction and
offscraping at Pacific margins, in Island arcs, Deep Sea Trenches, and back-arc
basins. In: Talwani, M., Pitman, W.C. (Eds.), Am. Geophy. Union, pp. 199210.
Screaton, E., 2006. Excess pore pressures within subducting sediments: does the
proportion of accreted versus subducted sediments matter? Geophys. Res.
Lett. 33, L10304. doi:10.1029/2006GL025737.
Screaton, E.J., Saffer, D., 2005. Fluid expulsion and overpressure development
during initial subduction at the Costa Rica convergent margin. Earth Planet.
Sci. Lett. 233, 361374.
Shipley, T.H., Stoffa, P.L., Dean, D.F., 1990. Underthrust sediments, fluid
migration paths, and mud volcanoes associated with accretionary wedge off
Costa Rica: middle America trench. J. Geophys. Res. 95 (B6), 87438752.
Shreve, R.L., Cloos, M., 1986. Dynamics of sediment subduction, melange
formation, and prism accretion. J. Geophys. Res. 91 (B10), 10,22910,245.
Sibson, R.H., 1981. Fluid flow accompanying faulting: field evidence and
models. In: Simpson, D., Richards, P.G. (Eds.), Earthquake Prediction: An
International Review. Maurice Ewing Series, 4. American Geophysical
Union, Washington, DC, pp. 593603.

467

Stauffer, P., Bekins, B., 2001. Modeling consolidation and dewatering near the
toe of the northern Barbados accretionary complex. J. Geophys. Res. (ISSN:
0148-0227) 106 (B4). doi:10.1029/2000JB900368.
Taira, A., Hill, I., Firth, J., Berner, U., Brckmann, W., Byrne, T., Chabernaud, T.,
Fisher, A., Foucher, J.P., Gamo, T., Gieskes, J., Hyndman, R.D., Karig, D.E.,
Kastner, M., Kato, Y., Lallemant, S., Lu, R., Maltman, A., Moore, G.F., Moran,
K., Olaffson, G., Owens, W., Pickering, K., Siena, F., Taylor, E., Underwood,
M.B., Wilkinson, C., Yamano, M., Zhang, J., 1992. Sediment deformation and
hydrogeology of the Nankai Through accretionary prism: synthesis of
shipboard results of ODP Leg 131. Earth Planet. Sci. Lett. 109, 431450.
Terzaghi, C., 1925. Principles of soil mechanics: II. Compressive strength of
clay. Eng. News-Rec. 95, 796800.
Thierry, P., Operto, S., Lambar, G., 1999. Fast 2-d ray+born migration/
inversion in complex media. Geophysics 64, 162181.
Tobin, H.J., Vannucchi, P., Meschede, M., 2001. Structure, inferred mechanical
properties, and implications for fluid transport in the dcollement zone,
Costa Rica convergent margin. Geology 29 (10), 907910.
Trenkamp, R., Kellogg, J.N., Freymueller, J.T., Mora, H.P., 2002. Wide plate
margin deformation southern Central America and northwestern South
America, CASA GPS observations. J. South Am. Earth Sci. 15, 157171.
von Huene, R., Klaeschen, D., Gutscher, M., Fruehn, J., 1998. Mass and fluid
flux duing accretion at the Alaskan margin. GSA Bull. 110 (4), 468482.
von Huene, R., Ranero, C.R., Vannucchi, P., 2004. Generic model of subduction
erosion. Geology 32 (10), 913916.
Vrolijk, P., 1990. On the mechanical role of smectite in subduction zones.
Geology 18, 703707.
Westbrook, G.K., Smith, M.J., Peacock, S.M., Poulter, M.J., 1982. Extensive
underthrusting of undeformed sediments beneath the accretionnary complex
of the Lesser Antilles subduction zone. Nature 300, 625628.
Winter, T., Avouac, J.-P., Lavenu, A., 1993. Late Quaternary kinematics of the
Pallatanga strike-slip fault (Central Ecuador) from topographic measurements of displaced morphological features. Geophys. J. Int. 115 (3),
905920.
Witt, C., Bourgois, J., Michaud, F., Ordoez, M., Jimnez, N., Sosson, M., 2006.
Development of the Gulf of Guayaquil (Ecuador) during the Quaternary as
an effect of the North Andean block tectonic escape. Tectonics 25, TC3017.
doi:10.1029/2004TC001723.
Wong, T.-F., David, C., Zhu, W., 1997. The transition from brittle faulting to
cataclastic flow in porous sandstones: mechanical deformation. J. Geophys.
Res. 102, 30093025.
Yilmaz, O., 2001. Seismic Data Analysis: Processing, Inversion and Interpretation of Seismic Data, Vols. 1 & 2. Society of Exploration Geophysicists,
Tulsa Oklahoma. 2027 pp.
Yuan, T., Spence, G.D., Hyndman, R.D., 1994. Seismic velocities and inferred
porosities in the accretionary wedge sediments at the Cascadia margin. J.
Geophys. Res. 99 (B3), 44134427.
Zhang, J., Wong, T.-F., Davis, D.M., 1990. Micromechanics of pressure-induced
graincrushing in porous rocks. J. Geophys.Res. 95, 341352.
Zoback, M.D., Byerlee, J.D., 1976. Effect of high-pressure deformation on
permeability of Ottawa sand. AAPG Bull. 60, 15311542.

Das könnte Ihnen auch gefallen