Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

RESEARCH ARTICLE Tectono-Stratigraphic Insights on the Dynamics of a Complex

10.1029/2023TC007860
Subduction Zone, Northern Peruvian Forearc
Key Points:
J. A. Lajo-Yáñez1 , S. S. Flint1 , M. Huuse1 , and R. L. Brunt1
• T his is the first integrated onshore
to offshore analysis of the northern 1
Department of Earth and Environmental Sciences, University of Manchester, Manchester, UK
Peruvian margin from forearc basin to
the trench and oceanic crust
• The subduction erosion process on the
northwestern Peruvian margin shows Abstract Two main types of subduction are recognized around the world: accretionary and erosive. The
spatial and temporal variability northern Peruvian margin is a well-known example of a margin subjected to subduction erosion, but to date the
• Uplift and collapse periods along-margin variability and temporal changes in subduction process and forearc basin evolution have not been
in the forearc were driven by
subduction-erosion tectonics, which characterized in detail. Interpretation of regional seismic lines and integration of oil-industry wells and seafloor
has been recorded in the basin data captures the erosive nature of subduction underneath the forearc with only a minor accretionary component
stratigraphy to the north. Episodes of uplift driven by plate coupling were followed by normal faulting/extensional collapse
due to plate decoupling. This mechanism explains the dominance of normal faulting across the forearc until
Supporting Information: the Oligocene with a slight reactivation within the Miocene. The subduction history is complex and includes a
Supporting Information may be found in reduction in plate convergence rate related to forearc crustal shortening, represented by large-scale structures
the online version of this article.
including the Peru fault (reactivated) and the Illescas fault-propagation anticlines of the Northwest Peru
transpressional system. This crustal deformation started in the Miocene. Integration with magnetic anomaly
Correspondence to: data indicates that activity of the present-day transpressional system driven by tectonic escape of the Nazca
J. A. Lajo-Yáñez,
Sliver toward the northeast, may explain the seismicity gap in southern Ecuador and northern Peru. An
jorge.lajoyanez@gmail.com
evolutionary model of the northern Peruvian margin shows how subduction zone geodynamics left its erosive
fingerprint in the forearc basin configuration.
Citation:
Lajo-Yáñez, J. A., Flint, S. S.,
Huuse, M., & Brunt, R. L. (2024).
Plain Language Summary The subduction process can be classified as accretionary or erosive.
Tectono-stratigraphic insights on the However, the northern Peruvian margin shows marked variability, being dominantly erosive but with a minor
dynamics of a complex subduction zone, accretionary component. Geophysical data show that deformation is represented by a transpressional system,
northern Peruvian forearc. Tectonics,
43, e2023TC007860. https://doi.
including the Peru fault and Illescas fault-propagation anticlines, driven by tectonic escape of the Nazca Sliver.
org/10.1029/2023TC007860 Forearc basin-fill records cycles of uplift followed by extensional collapse, driven by cycles of plate coupling
and decoupling. Subduction zone geodynamics leaves its fingerprint in the fills of forearc basins, in this case by
Received 27 MAR 2023 basal and frontal crustal tectonic erosion.
Accepted 11 NOV 2023
Corrected 7 FEB 2024

This article was corrected on 7 FEB 2024.


1. Introduction
See the end of the full text for details.
Characterization of the subduction process is crucial to understand the controls on the evolution of forearc basins.
Subduction zone processes have been studied in several examples around the Circum-Pacific Ring of Fire, and can
Author Contributions:
be classified into two main types of overriding plate margin: accretionary and erosive (Clift & Vannucchi, 2004;
Conceptualization: J. A. Lajo-Yáñez,
S. S. Flint
Cloos & Shreve, 1988; Noda, 2016; Scholl et al., 1980; von Huene et al., 2004). The main factors that differ-
Data curation: J. A. Lajo-Yáñez entiate an erosive margin from an accretionary type are a poorly sedimented trench and high plate convergence
Formal analysis: S. S. Flint, M. Huuse, rates (Clift & Vannucchi, 2004). Sediment supply can be controlled by climate and catchment size (Kukowski &
R. L. Brunt
Investigation: J. A. Lajo-Yáñez
Oncken, 2006). In the southeastern Pacific, where the continental South American Plate overrides the oceanic
Methodology: J. A. Lajo-Yáñez Nazca Plate, the subduction margins of Colombia (Collot et al., 2008), Ecuador (Collot et al., 2011), central and
Project Administration: J. A. southern Peru (Clift et al., 2003; Genge et al., 2020; Herbozo et al., 2020; von Huene & Lallemand, 1990), and
Lajo-Yáñez
Resources: J. A. Lajo-Yáñez
northern and central Chile (Contreras-Reyes et al., 2015; von Huene & Scholl, 1991) have been identified as the
subduction erosion type.

The section of the northern Peruvian forearc in the present study bridges the gap between these better studied
© Wiley Periodicals LLC. The Authors.
This is an open access article under
areas to the north and south. This forearc section generated accommodation for the deposition of the Talara Basin
the terms of the Creative Commons through the Late Cretaceous to the Eocene. Stratigraphic and structural aspects of this basin have been studied
Attribution License, which permits use, at an intra-basin scale (Baldry, 1938; Bush et al., 1994; Euribe, 1976; Fildani et al., 2008; Gonzales, 1976;
distribution and reproduction in any
medium, provided the original work is
Hessler & Sharman, 2018; Higley, 2004; Iddings & Olsson, 1928; Marsaglia & Carozzi, 1990; Murany, 1975;
properly cited. Peralta-Cárdenas, 1967; Séranne, 1987; Stainforth, 1954; Travis, 1953; Travis et al., 1976; Weiss, 1955); however

LAJO-YÁÑEZ ET AL. 1 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Supervision: S. S. Flint, M. Huuse, R. elements such as relative sea level, origin of accommodation for the thick basin fill, sedimentary environments
L. Brunt
and tectonic history have remained unresolved.
Validation: J. A. Lajo-Yáñez, S. S. Flint,
M. Huuse, R. L. Brunt
In this work we present a multi-scale analysis, addressing aspects of plate tectonics and basin structure and
Visualization: J. A. Lajo-Yáñez
Writing – original draft: J. A. mega-sequences, based on the integration of oil industry and academic data sets of the northern Peruvian forearc
Lajo-Yáñez which allows us to re-evaluate the models of subduction. The study area comprises a region from 1°S to 9°S in
Writing – review & editing: J. A. Lajo-
northwestern Peru, between the Peru Trench off Chimbote city to the south, and southwestern Ecuador to the
Yáñez, S. S. Flint, M. Huuse, R. L. Brunt
north (Figure 1). This includes the forearc region of the Talara Basin.

The aim of this study is to analyze the basin and subduction zone elements in order to critically test the applica-
tion of the subduction erosion model to the northern Peruvian forearc, especially to the Talara Basin. The results,
defining the specific type of subduction mechanism in the area and its development through geological time, help
us identify possible relationships between the subduction history and Talara Basin evolution.

2. Geological Framework
2.1. Nazca Plate
The geodynamic conditions between the Nazca and South American Plates have varied over time. Pardo-Casas
and Molnar (1987) determined the Farallon/Nazca Plate convergence rate and drift direction relative to the South
American Plate. Convergence was slow from the end of the Cretaceous until the beginning of the Early Eocene
with a drift direction generally to the north, which then changed to the northeast, accompanied by high convergence
rate until an abrupt decrease during the Oligocene. At end-Oligocene, the convergence rate increased again, when
the Farallon Plate split into the Nazca and Cocos Plates (Lonsdale, 2005). Today, the Nazca Plate is converging
with the northwestern Peruvian margin in an east-northeast direction at rates of 5.97, 6.20, and 6.27 cm year −1
in the north, center and south, respectively (Kendrick et al., 2003 cited in Villegas-Lanza et al., 2016a), and the
age of the portion of plate in front of the Peru Trench ranges from 40 to 30 Ma, from north to south, based on the
sea-floor age map from Müller et al. (2008; Figure 1a). Based on earthquake distribution and magnitude, volcanic
activity and outcrop data, the present-day northern and central Peruvian margins overlie a segment of oceanic
crust known as the Peruvian Flat Slab, which has a low-angle subduction plane beneath the South American Plate
(Barazangi & Isacks, 1976, 1979; Ramos & Folguera, 2009).

The structural configuration in the northern region of the Nazca Plate is represented by the Grijalva (1–2 km-high
scarp, 10–20 km wide and 680 km long), Alvarado and Sarmiento (1–2 km high, 10–20 km wide and 400 km long)
Ridges. These features have southwest-northeast orientations and lie in front of the northern Peruvian forearc.
They have been interpreted as ancient transform faults within the Farallon Plate by Mammerickx et al. (1975),
and volcanic ridges by Lonsdale (2005). Some hundreds of kilometers to the south, Huchon and Bourgois (1990)
identified a large northeast-orientated fracture zone called Virú (Figure 1b), which deviates to the north-northeast
close to the Peru Trench, toward the Illescas hills, where it begins to show thrust-faulted structures around the
Trujillo trough. These positive structural elements were interpreted as due to oceanic plate fragmentation induced
by subduction, and linked to the south to the opening of the Mendaña Fracture Zone perpendicular to its trend,
where new oceanic crust has been created since 3.5 Ma. Prince et al. (1974), Kulm et al. (1982), and Prince and
Kulm (1975) considered the ridges to the west of the Peru Trench off Chimbote city to be the result of imbricate
thrusting due to oceanic crust rupture.

2.2. Subduction Erosion on the Peruvian Pacific Margin

The concepts of sediment subduction and subduction erosion at convergent margins appeared from the 1960s
(Coats, 1962; Hilde, 1983; H. Miller, 1970; J. Murauchi, 1971; Scholl et al., 1980) as a different process to
subduction accretion (Dickinson & Seely, 1979; Seely et al., 1974). Four main differentiating features were iden-
tified: unbalanced volumes in sediment budget, the lack of pelagic oceanic deposits in outcropping subduction
complexes, high plate convergence rates, and the poor development or absence of an accretionary prism (Clift &
Vannucchi, 2004; Scholl, 1987; Scholl et al., 1980; von Huene & Scholl, 1991).

Erosive margins are the result of subduction erosion and represent more than half of the present-day subduction
margins around the world (Clift & Vannucchi, 2004; von Huene et al., 2004; von Huene & Scholl, 1991). However,
subduction erosion is also present, to a lesser or greater degree, even in accretionary margins (Stern, 2011; von

LAJO-YÁÑEZ ET AL. 2 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 1. (a) Sea-floor age map (Müller et al., 2008) showing the location of the study area within the context of the
oceanic Nazca Plate subducting beneath the continental South American Plate. (b) Bathymetry and topography from Global
Multi-Resolution Topography (GMRT) synthesis showing the distribution of Ribiana-Petroperú 2D seismic surveys and well
data over the Talara forearc basin. Current plate-convergence rate in cm year −1 (yellow arrows) by Kendrick et al. (2003, cited
in Villegas-Lanza et al., 2016a). R, ridge; FZ, fracture zone. Coastal Range outcrops: (1) Lobos de Afuera island, (2) Lobos
de Tierra island, (3) Illescas hills, (4) Paita hills, (5) La Brea hills, (6) Amotape-Tahuín hills.

Huene & Lallemand, 1990), with small frontal prisms (von Huene & Scholl, 1991). In the process of subduc-
tion erosion, missing crustal material derived from the overriding plate is assimilated in the subduction chan-
nel, through shear with the subducting plate (Scholl & von Huene, 2007). This type of subduction implies
tectonic erosion that can be frontal, over the lower slope of the forearc, and basal, at the base of the upper plate
(Stern, 2011; von Huene & Lallemand, 1990; von Huene & Scholl, 1991). As a consequence, tectonic erosion
causes thinning and loss of continental plate margin material by oceanic crust abrasion (Langseth et al., 1981; S.
Murauchi & Ludwig, 1980) and a simultaneous landward retreat of the trench (Scholl et al., 1980; Scholl & von
Huene, 2007, 2010).

Erosive margin models have been built across upper plate—trench—subducting plate profiles from Central Amer-
ica to central Chile, and western Pacific oceanic-continental plate boundaries (Clift & Vannucchi, 2004; Noda &
Miyakawa, 2017; Scholl & von Huene, 2007). An important characteristic of erosive margins along this sector
is the trench fill thickness, which is <1 km, except for Colombia and northern Ecuador (Scholl et al., 2015). The
entire Peru Trench fill thickness is <1 km, which is consistent with an arid coast with low sediment supply, and
is an indicator of the erosive nature of the Peruvian margin. In the south of the study area, at the latitudes of
Bayóvar bay and Trujillo city, the subduction style was originally interpreted as an accretionary complex of more
than 15 km length (Bourgois et al., 1988; von Huene et al., 1989; von Huene & Miller, 1988). However, toward
the north, in the Gulf of Guayaquil, in southwestern Ecuador, the margin shows development of an incipient

LAJO-YÁÑEZ ET AL. 3 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

accretionary prism (Collot et al., 2011). The central and southern Peruvian margins are being affected by the
Nazca Ridge, an important bathymetric high in the oceanic crust that has migrated southwards along the margin,
where the subducting portion has reconfigured the subduction zone and the tectonic evolution of the forearc
through time (Contreras-Reyes et al., 2019; Hampel, 2002; von Huene & Lallemand, 1990).

2.3. Northern Peruvian Forearc and the Talara Basin

The northern Peruvian forearc is located on the Pacific convergent margin (Moberly et al., 1982). To the east, the
area is bounded by Paleozoic low-grade metamorphic basement of the Coastal Range, intruded by granitic rocks
of the Triassic magmatic arc (Bellido et al., 2009; Sánchez et al., 2006) (Figure 1b).

The Late Cretaceous to Late Eocene Talara Basin lies in the forearc along with the Oligocene-Miocene Tumbes
basin and younger units. The Talara Basin-fill comprises an 8-km thick succession of mainly siliciclastic sedi-
ments (Travis et al., 1976) and is separated by angular unconformities from the Early Cretaceous and Paleo-
zoic metasedimentary basement (Higley, 2004). The boundaries of the Talara Basin are poorly defined (Fildani
et al., 2008), except for the eastern limit with the Coastal Range. Toward the north, Upper Cretaceous and Eocene
units are found in southwestern Ecuador, to the south of the Chongón-Colonche mountains, with a good biostrati-
graphic correlation (Coxall, 2000; Frizzell, 1945; Galloway & Morrey, 1929; Hofker, 1956; Stainforth, 1948;
Thalmann, 1946; Vaughan, 1937); therefore, there is a possibility that those units make up the northern bound-
ary of the basin, passing through the Gulf of Guayaquil. In the same way, the southern sedimentary limit has
been defined to the south of the Illescas hills as part of the Paleozoic mountains (Higley, 2004), although the
Early Cenozoic portion of the Talara Basin may be correlated to the Trujillo Basin farther south (Thornburg &
Kulm, 1981).

3. Data Sets and Methods


The main database for this study is seven regional seismic reflection lines that cover the subduction zone between
the Nazca and South American Plates. These lines were acquired in 1993 by Ribiana Inc. in agreement with the
state-owned company Petroperú S.A. This work uses the latest reprocessing performed by Spectrum (now TGS)
in 2014, based on post-stack time migration and converted to depth domain using the processing velocity. The
reason for displaying the seismic lines in depth is to avoid the effect of deep-water bathymetry in time domain that
could mislead the seismic interpretation due to visual distortion of the subduction zone geometry. Oil-industry
well data, intersecting the western end of seismic lines, were important as a control on mapping the boundaries
of different forearc basin sequences. These subsurface data sets, including local seismic surveys, were provided
by the Peruvian government institution Perupetro S.A. The onshore outcropping lithologies were covered by
surface geological maps at a scale of 1:100,000 from the Geological, Mining and Metallurgical Institute of Peru
(INGEMMET), and Lajo-Yáñez et al. (2022). In addition, bathymetry and topography data were obtained from
Global Multi-Resolution Topography synthesis and Geersen (2019). Bathymetric surveys incorporate seafloor
geological mapping, sampling and seismic surveys from research vessels in previous campaigns around the Peru
Trench: Nazca Project in 1972 (Kulm et al., 1982; Prince et al., 1974; Prince & Kulm, 1975), SEAPERC in
1986 (Bourgois et al., 1988; Huchon & Bourgois, 1990), NAUTIPERC in 1991 (Bourgois et al., 1993; Sosson
et al., 1994) and GEOPECO in 2000 (Krabbenhöft et al., 2004). Other data used are the seafloor magnetic anom-
alies map from the National Oceanic and Atmospheric Administration (Maus et al., 2009). Additional seismolog-
ical and GPS velocity field data along the continental margin of Peru (Villegas-Lanza et al., 2016a) and Ecuador
(Nocquet et al., 2014) were integrated.

This study started with bathymetry interpretation carried out to analyze the surficial structural features on the
seafloor around the Peru Trench, integrated with geological mapping from vessels. Structural and stratigraphic
interpretation of west-east seismic lines 93-01, 93-10, 93-16, 93-19, 93-23 and 93-25 characterized the northern
Peruvian forearc basins at different latitudes and these lines were intersected by a south-north margin-parallel
seismic line 93-20. Interpretation of faults and other structural elements through geological time enabled the
structural restoration to inform the stratigraphic analysis. Integration with the seafloor study allowed calibration
of shallow seismic stratigraphic and structural interpretations. Tectonic analysis of the plate boundary consisted
of identifying and characterizing each element of the subduction system to determine the mechanism specific to
the study area. An important point in this analysis was the timing of any changes in subduction process recorded

LAJO-YÁÑEZ ET AL. 4 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 2. High-resolution integrated bathymetric and topographic maps showing the main surficial structural features in the oceanic crust, the Peru Trench, and the
South American Plate margin (source: Global Multi-Resolution Topography [GMRT] synthesis and Geersen (2019)). (a) HSV (Hue, Saturation, Value) color model,
(b) Slope shading and (c) Interpretation maps of the different morpho-structural elements potentially involved in the northern Peruvian and southern Ecuadorian forearc
configuration. Seismic lines (red) and well data (yellow circles) used for subsurface interpretation, where the Illescas faulted-anticlines (light-green dashed lines) and
Peru fault (cyan dashed lines) are the proposed structures in this work. R, ridge; FZ, fracture zone; MTDs, massive transport deposits.

by forearc basin evolution. Finally, an interpretation of the seafloor magnetic anomalies in front of the Peruvian
and Ecuadorian margins of the South American Plate, delineated the geometry and structural configuration of the
northernmost segment of the Nazca Plate.

4. Seafloor Structural Configuration Around the Northern Peru Trench


High-resolution bathymetry mapping by ship-based multibeam acoustic data (Figures 2a and 2b) was linked
directly or indirectly to subsurface structural elements imaged in seismic lines. This information complemented
the subsurface interpreted geology and vice versa, especially around the northern Peru Trench where it provided
map views of the surface boundary of the South American Plate.

LAJO-YÁÑEZ ET AL. 5 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

4.1. Nazca Plate


4.1.1. Oceanic Plate Bend Faulting

Slope shading of bathymetry (Figure 2b) highlights normal faults which can be easily seen on the present-day
seafloor, identified as bend faults (Geersen et al., 2022; Kita & Ferrand, 2018; Ranero et al., 2005). A key obser-
vation of bend faults is their distribution within the Nazca Plate and relationship with adjacent structures. These
faults surround the western margin of the trench, facing the Gulf of Guayaquil, Talara and Trujillo sector of the
forearc (blue thin lines in Figure 2c). Bend faulting in the southern sector preserves a general north-northwest
alignment, parallel to the seafloor magnetic anomalies map used in this work (Maus et al., 2009). This alignment
suggests a spreading fabric source for the faults that then are reactivated as they approach the plate bending
sector. From the area between lines 93-16 and 93-10 to the north, the trench axis direction changes northwards
and bend faulting re-aligns with the new trench axis direction, suggesting the formation of new bend faults beside
the reactivated ones.

4.1.2. Thrust- and Strike-Slip Structures

Mammerickx et al. (1975) identified parallel ridges on the seafloor, aligned to the northeast, in front of the north-
ern Peruvian forearc and Gulf of Guayaquil, known as the Sarmiento, Alvarado, and Grijalva Ridges (Figure 2c),
and classified them as fracture zones related to transform faults of the Farallon Plate. A fourth one identified in
this work is named the Yáñez Ridge. However, Lonsdale (2005) argued that they do not cross the oceanic crust
magnetic anomalies orthogonally, as in the case of the Mendaña fracture zone, and proposed an origin as volcanic
ridges built over eruptive fractures not related to transform faults.

To the south (∼8°S), around the Trujillo trough, a zone of seafloor structures related to thrust faults was docu-
mented by Huchon and Bourgois (1990) and Prince and Kulm (1975), herein named the Trujillo Ridges, which
are connected to the Virú Fracture Zone. While the thrusts mostly strike northwards and dip eastwards, toward
the Peru Trench (light-orange polygons with red jagged lines in Figure 2c), the Virú Fracture Zone has a north-
east trajectory (dark-orange polygons in Figure 2c). In a wrench fault kinematic context, these two structural
features could mark possible intraplate dextral transpressional deformation of the Nazca Plate. To the southeast
of the Trujillo Ridges (∼9°S), another area of ridges off Chimbote city aligned to the trench western margin was
identified by Bourgois et al. (1988), Kulm et al. (1982), and Prince and Kulm (1975). These positive geomor-
phological structures are herein termed the Chimbote Ridges (southeasternmost orange polygons with red jagged
lines in Figure 2c). Based on interpretation of single- and multi-channel seismic, previous authors interpreted the
structures as westward-directed imbricate thrusts related to oceanic crust rupture. The ridges can be imaged in
detail through ship-based bathymetry data compiled in Figure 2a. The eastern flank of these ridges is flat and dips
steeply toward the trench, while the western side shows a fault-like scarp. Thus, the Chimbote Ridges could also
belong to the transpressional system.

4.2. Northern Peru Trench

Bathymetry augmented by HSV (Hue, Saturation, Value) color model (Figure 2a) shows the Peru Trench as
a deep and elongated flat surface bounded by the South American Plate margin and the outcropping oceanic
crust, which define the inner and outer trench walls, respectively. The trench bathymetry represents the youngest
undeformed sedimentary layer of the trench fill at depth ranges between 6 and 4.5 km below sea level (kmbsl),
in front of the northern Peruvian margin. It gets narrower and shallower to the north, being less than 4 kmbsl,
as it is affected by the Carnegie Ridge (Michaud et al., 2009). The eastern margins of the Alvarado and Yáñez
Ridges cross and deform the trench, and from that point to the north the trench top becomes shallower. Seismic
data shows a trench fill with an average thickness of 0.5 km.

4.3. Northern Peru and Southern Ecuador Slope and Shelf

In order to determine the slope along the Ecuadorian—Peruvian forearc, the 200-m sub-sea contour was extracted
from the bathymetric map since it approximates the shelf edge (light-gray polygon in Figure 2c). Thus, in map
view, the slope widens to the north in the Gulf of Guayaquil and to the south of the Illescas hills. The Peru
Bank, identified in the northern sector, is raised as an important elongated morpho-structural element that has a
southwest-northeast axis direction and a similar flat shallow water relief as the shelf.

LAJO-YÁÑEZ ET AL. 6 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 3. (a) Uninterpreted and interpreted west-east seismic line 93-01, crossing the northern part of the northern Peruvian forearc, and also the Peru Bank. As this
line does not cover the trench, the western section of line SIS-72 was projected onto this line due to their proximity (∼21 km). Seismic profile only was taken from
Collot et al. (2011) and interpreted in this work. (b) Zoomed-in view of the upper section of the Peru fault showing the syn-sedimentary activity during the Miocene.
(c) Spectral decomposition on a 3D survey by Borda and Bianchi (2020), approximately located by the blue dashed circle in basemap, ∼4 km to the south of line
93-01, the displacement of frequency anomalies indicates a possible strike-slip reactivation of the Peru growth fault during the Miocene, that formed a half-graben in
the Oligocene. Wireline logs used are gamma ray (green), spontaneous potential (red), and deep resistivity (black); values increase to the right. OCB, Oceanic crust
basement; SC, Subduction channel; Pz+, Paleozoic basement and older units; E, Eocene; O, Oligocene; M, Miocene; P, Pliocene.

5. Geological Configuration of the Northern Peruvian Forearc and the Talara Basin
West-east and south-north geological transects were built based on the longest and most representative seismic
lines across the northern Peruvian forearc at different latitudes, integrated with well data, research vessel surveys
and outcrop mapping. Predominant structural styles are described in conjunction with the main sedimentary
mega-sequences.

5.1. Northern Zone (3°S–4°S)

The seismic line 93-01 (Figure 3a) shows a transverse section of the forearc northern region. Upper
Cretaceous-Paleocene formations are absent in this sector in contrast to biostratigraphically-constrained
Late Cretaceous deposits in southwestern Ecuador (Coxall, 2000; Frizzell, 1945; Galloway & Morrey, 1929;
Hofker, 1956; Stainforth, 1948; Thalmann, 1946; Vaughan, 1937). This suggests no sedimentary continuation
or later erosion between Peruvian and Ecuadorian forearc basins at that time. In the north, the stratigraphy
comprises Eocene strata of the Talara Basin, overlain by Oligocene, Miocene and Pliocene strata of the Tumbes
Basin, which is calibrated by well Corvina-40-X-1 to the east (Figure 3a). The Eocene mega-sequence does not
reach the trench inner wall, likely due to non-deposition, possibly augmented by erosion, suggesting an uplifted
area during that period, which is overlain by Miocene-Pliocene strata. The Coastal Range is not imaged by this
seismic line since it is located further to the east.

The structural features seen at this latitude are predominantly normal faults, mostly concentrated in the slope,
accompanied by folding and few associated short-displacement reverse faults. Integration with geological surface
maps and oil-industry seismic and well data shows a southwest-northeast trend of the normal faults around the

LAJO-YÁÑEZ ET AL. 7 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

shelf. It also captures a northwest-directed tilting and thickening of the Early Eocene mega-sequence, suggesting
accommodation was higher in the northwest of the Talara Basin at this time.
Oligocene sedimentation filled the northwestern depression. However, local variability is related to growth fault
activity on a structure named here as the Peru fault (cyan thick line in Figure 3). Miocene sequences exhibit
divergent seismic facies toward the fault, confirming syn-sedimentary behavior at that time (seismic reflectors
between green dashed lines in Figure 3), until they onlap the footwall block. To the west of the half-graben,
Eocene and Miocene mega-sequences are conformable, which means that this sector was stable and without
accommodation during the half-graben formation. With a listric geometry, the Peru fault is interpreted as a criti-
cal deep structural element deforming the continental crust.
The upper 3 km of the Peru fault plane slightly overturns eastwards, and bounds a deformed footwall anticline.
Reflectors onlap this area (green dashed line in Figure 3b) and are truncated by a conspicuous unconformity (red
dashed line in Figure 3b). Short-displacement inversion of the half-graben was identified in the uppermost 1.5 km
of the fault (cyan dashed line in Figure 3b), forming a subtle harpoon structure. In this section, spectral decom-
position performed by Borda and Bianchi (2020) on a seismic cube, 4 km to the south of line 93-01, through
Miocene sequences, indicates later strike-slip behavior of the Peru fault. This seismic attribute shows a strati-
graphic feature, possibly related to a fan with a sedimentation direction to the SE, whose frequency anomalies
show a dextral displacement (Figure 3c). These observations suggest that compressional stress extended across
the west of the Peru fault and reactivated it with dextral strike-slip kinematics producing an oblique inversion,
still in the Miocene. Folding and uplift of the entire sedimentary fill identified below the Peru Bank (submarine
plateau at the center of Figure 3), including the underlying Paleozoic basement, is consistent with subsequent
compressional stress until the Pliocene.
Finally, another group of large listric faults defines a horst close to the continental margin edge (left-hand side
of Figure 3a). Pliocene strata, onlapping the erosive top of the Miocene mega-sequence on the horst, suggest an
Early Pliocene age of formation, likely related to a period of subsidence. Several normal faults affect mostly the
Oligocene and younger strata along the slope. The normal fault density in the forearc is higher from the west of
the Peru fault until the trench, and may represent one of the last deformation phases in the area as a stress release
period following the compression.

5.2. Central Zone (4°S–5.4°S)

Interpretation of line 93-10 (Figure 4) demonstrates a near-constant thickness of stratified units within the Upper
Cretaceous-Paleocene and Eocene mega-sequences overlying the Paleozoic basement. These units extend from
the Coastal Range westward to the trench inner wall. The Eocene mega-sequence overlies nearly parallel Upper
Cretaceous-Paleocene strata, generally with a subtle unconformity, which indicates no abrupt tectonic deforma-
tion during sedimentation through this period. Eocene strata extend across all of the forearc, which suggests a
major area of accommodation during this period, as seen in the northern zone. The Oligocene mega-sequence is
almost completely restricted to the deepest half of the margin slope, and overlies Eocene strata with a sub-parallel
contact. The Miocene mega-sequence overlies the Eocene and Oligocene units with an angular unconformity,
denoting a Pre-Miocene structural rearrangement. Pliocene strata also have discordant contact with older units.
The only structural style characterizing this central zone is normal faults, which formed across the forearc, most
of them dipping to the west. Most of these faults record extensional strain from the Eocene, with some reactiva-
tion before the Pliocene. As a result of this crustal deformation, major subsidence was produced by deeper normal
faulting that terminates close to the subduction zone, resulting in a hanging-wall syncline. Resultant accommo-
dation was filled by Oligocene, Miocene and Pliocene mega-sequences. The central zone is the best example of
slope collapse in the study area.
Local seismic surveys, well logs and outcrops in the Negritos oilfield (right-hand side of Figure 4a) allow
integration of data from line 93-10 to the La Brea hills (local sector of Coastal Range, including the Triassic
magmatic arc). The structural and stratigraphic configuration is of rotated blocks, bounded by listric normal
faults, and comprising the Eocene and Upper Cretaceous-Paleocene mega-sequences overlying the Paleozoic
meta-sedimentary basement. Some faults are syn-sedimentary and were activated in the Early-Middle Eocene,
controlling the deposition of fluvio-deltaic systems (Lajo-Yáñez et al., 2022), which are overlain by Upper
Eocene deep-water deposits. Biostratigraphic and core data from deep wells indicate Upper Cretaceous systems
developed in environments ranging from transgressive shelf to deep-water.

LAJO-YÁÑEZ ET AL. 8 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 4. (a) Uninterpreted and interpreted west-east seismic line 93-10, located in the central part of the northern Peruvian forearc from the Peru Trench to the
Coastal Range. This is a section with extensive well control for interpretation (Negritos oilfield), which shows a larger development of the Talara Basin Cretaceous to
Eocene mega-sequences than to the north and south. The main structural characteristic is the high density of normal faulting verging to the west, giving a configuration
of a collapsed forearc. (b) Zoomed-in view of the main unconformities between Eocene, Oligocene, Miocene and Pliocene mega-sequences. Wireline logs used are
gamma ray (green), spontaneous potential (red), and deep resistivity (black); values increase to the right. OCB, Oceanic crust basement; SC, Subduction channel; Pz+,
Paleozoic basement and older units; K-P, Upper Cretaceous and Paleocene; E, Eocene; O, Oligocene; M, Miocene; P, Pliocene.

5.3. Southern Zone (5.4°S–7°S)

Similar to the central zone, lines 93-16 and 93-19 (Figures 5 and 6) show the western boundary of Upper
Cretaceous-Paleocene sediments at the trench inner wall. The Eocene mega-sequence shows a dramatic thickness
decrease seaward and is overlain by Miocene, Pliocene, and Pleistocene deposits, separated by unconformities.
As in the central zone, Upper Cretaceous-Paleocene and Eocene mega-sequences are sub-parallel, confirming
no major tectonic deformation between them. Both mega-sequences onlap the western flank of the subcropping
Coastal Range; however, calibration with nearby wells SP1-1X and SP2-1X-ST indicates that only the Eocene
along with Oligocene and Miocene mega-sequences continue further eastward, covering the large structure.

The structural style is partly extensional, represented by a high density of mostly-westward dipping normal faults
concentrated in the slope and at the plate edge. Antithetic normal faults appear in the southernmost lines. The
other structural component is compressional, marked by a kilometer-scale group of large fault-propagation anti-
clines, herein named Illescas since they formed in front of the Illescas Paleozoic high, resulting in shortening of
the continental crust. These structures are documented in lines 93-16, 93-19, 93-23, and 93-25 (Figures 5–8) and
show apparent west-dipping reverse faults, suggesting a general eastward direction of the compression, concord-
ant with the Nazca Plate drift direction. The folding appears to affect not only the continental crust, but the
oceanic crust. The angular unconformity of the Miocene over the folded older sequences at the top of the anticline
zone suggests that the structural deformation took place shortly before, or at the beginning of the Miocene.
Seafloor dives NP26 and NP34 performed by the Nautile submersible (Sosson et al., 1994) in the Chiclayo
canyon, sampled Paleozoic black schists and quartzites, which correlate with the basement drilled by deep wells
in the oilfields and outcrops in the Coastal Range. Seismic line 93-25 (Figure 8) crosses the Chiclayo canyon and

LAJO-YÁÑEZ ET AL. 9 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 5. (a) Uninterpreted and interpreted west-east seismic line 93-16, representing the south of the northern Peruvian forearc, passing through the Nazca Plate
and the Peru Trench to the Bayóvar bay. An incipient Illescas fault-propagation anticline is forming (light-green line) and deforming the continental crust. Nautile
submersible survey helped interpretation of the shallow stratigraphy on the slope. The Cretaceous of the Talara Basin is restricted to the center and west. (b) Zoomed-in
view of the mass-transport deposits (MTDs) on the lower slope. OCB, Oceanic crust basement; SC, Subduction channel; Pz+, Paleozoic basement and older units; K-P,
Upper Cretaceous and Paleocene; E, Eocene; O, Oligocene; M, Miocene; P, Pliocene.

records a large-scale uplifted Paleozoic structure, ∼12 km from dive NP34, that appears to be part of the Illescas
fault-propagation anticlines. Some of the normal faults cut the Illescas fault-propagation anticlines, and the over-
lying unconformable Miocene strata, which shows them to have been the last to form.
Twenty five northward seafloor dives from this submersible in front of the Bayóvar bay (Figure 5a), and in
combination with a hydrosweep survey from the research vessel Sonne, identified Pleistocene debris-avalanche
deposits on the slope (Bourgois et al., 1993). The distribution of these deposits was controlled by a group of three
main curved collapse scarps that were mapped with bathymetry collected by the Jean Charcot cruise (Bourgois
et al., 1988; von Huene et al., 1989). Integration with seismic lines 93-16 and 93-19 (Figures 5 and 6, respec-
tively), crossing the mapped area, shows discontinuous seismic reflections with internal low amplitudes bounded
by an erosional base located in the lower slope. These deposits are distributed seaward, adjacent to the middle
slope, which is delimited by the upper and middle scarps related to normal faulting. The location of this slope
segment just above the Illescas fault-propagation anticlines is consistent with an unstable slope formed by uplift.
The lowermost scarp, originally interpreted to be related to the subduction zone, is well displayed in line 93-16 as
a listric fault cutting the plate wedge tip and ending up as thrusts in the oceanic deposits that overlie the oceanic
crust basement, producing a gravity-driven slide. This third scarp is at ∼5 km up dip from the identified subduc-
tion entry in this study. The interpretation associates these structures to an unstable submarine slope conducive to
development of mass-transport deposits (MTDs) that flowed into the trench.

5.4. Lateral Variability in the Forearc

The south-north line 93-20 (Figure 9) provides the critical seismic-stratigraphic tie between the main west-east
geological transects, and gives spatial context to the stratigraphic and structural features, recording the variable
distribution of the mega-sequences along the northern Peruvian forearc.

LAJO-YÁÑEZ ET AL. 10 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 6. (a) Uninterpreted and interpreted west-east seismic line 93-19, further south in the northern Peruvian forearc, covering the section between the Peru Trench
and the Paleozoic Coastal Range represented by the Illescas hills, where the Triassic magmatic arc is currently outcropping. (b) Zoomed-in view of the mass-transport
deposits (MTDs) on the lower slope. Wireline logs used are gamma ray (green) and deep resistivity (black); values increase to the right. OCB, Oceanic crust basement;
SC, Subduction channel; Pz+, Paleozoic basement and older units; K-P, Upper Cretaceous and Paleocene; E, Eocene; O, Oligocene; M, Miocene; P, Pliocene; BSR,
Bottom-simulating reflector.

As the basal mega-sequence of the Talara Basin, the Upper Cretaceous-Paleocene strata have a near constant and
considerable thickness of about 2.5 km from the south to the central zone and then thin toward the north, pinching out
before the intersection with line 93-01. Characteristic benthic foraminifera from oil-industry well data suggest initial
outer-shelf deposition (100–200 m) in the Late Cretaceous that changed to bathyal-to-abyssal (200–>2,000 m), and
ended in shelf-to-marginal-marine (200–10 m) in the Paleocene. In an opposite sense, the Eocene mega-sequence
appears thicker in the northern half of the forearc; however, west-east lines support the evidence for an angular uncon-
formity below the Miocene and younger mega-sequences, which means that Eocene strata could have originally
been more regional. Within this mega-sequence, local seismic surveys record a thicker Lower Eocene sequence to
the north, whose benthic foraminifera indicate deposition in a bathyal-to-middle-shelf environment (2,000–100 m),
which is absent in the southern area. Middle and particularly Upper Eocene deposits covered the whole basin, charac-
terized by shelf-to-marginal-marine (200–10 m) and bathyal-to-abyssal (200–>2,000 m) environments, respectively.
The Oligocene mega-sequence, deposited in a middle-shelf-to-marginal-marine environment (100–10 m), presents
along-margin discontinuity, which suggests part truncation by the unconformity and is onlapped by Miocene and
Pliocene strata, representative of outer-shelf environment (200–100 m). Accommodation history in the northern
Peruvian forearc can therefore be summarized as follows: (a) high sea level in Late Cretaceous, probably coinciding
with global highstand (Haq, 2014; Kominz et al., 2008; K. G. Miller et al., 2005; Simmons et al., 2020), (b) northern
subsidence in Early Eocene, (c) relative sea-level rise in Middle and Late Eocene due to continental margin subsid-
ence produced by local normal faulting within the basin, and (d) subsequent normal faulting from the Oligocene.

LAJO-YÁÑEZ ET AL. 11 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 7. (a) Uninterpreted and interpreted southwest-northeast seismic line 93-23, to the southwest of the Illescas hills, in the northern Peruvian forearc. The
Illescas fault-propagation anticlines pushed the continental crust against the Triassic intrusive massif, which prevented passage of the subduction channel underneath.
(b) Zoomed-in view of the anticline related to the continental crust shortening. OCB, Oceanic crust basement; Pz+, Paleozoic basement and older units; K-P, Upper
Cretaceous and Paleocene; E, Eocene; O, Oligocene; M, Miocene; P, Pliocene; BSR, Bottom-simulating reflector.

High-density normal faulting has resulted in a collapse configuration to the forearc, mostly in the center and south,
which was the main mechanism for producing forearc accommodation for the Talara and younger basins since the
Eocene. In addition, the two largest forearc-scale structural features associated with crustal shortening, are identi-
fied north and south of line 93-20. They are the Peru growth fault (thick cyan lines in Figure 10) and the Illescas
fault-propagation anticlines (thick light-green lines in Figure 10), respectively. The compressional and transpres-
sional (reactivation) deformation related to these large structures suggests a Pre- or Early-Miocene initiation. The
relationship between those fault groups is interpreted as sequential stresses; reverse faults responded to compres-
sional stress due to the acceleration in Nazca Plate drift and once the stress started to diminish, the margin of the
South American Plate experienced subsequent extensional collapse in the Oligocene. These events are thought to
have continued until the Quaternary and be associated with the stress build-up and release intervals due to plate
coupling and decoupling, respectively, identified by Bourgois et al. (2007), DeVries (1988), and Pedoja et al. (2006).

6. Subduction Elements and Mechanisms in the Northern Peruvian Margin


6.1. Trench Fill

Zoomed-in views of the trench from seismic lines SIS-72, 93-10 and 93-16 (Figures 11a–11c, respectively), show
that the sedimentary fill thickness is ∼0.6 km in the northern and central zones, and ∼0.4 km in the southern
zone. The trench fill also extends landward beyond the trench, underneath the continental margin. Contorted
reflectors suggest that part of the sediments that entered the subduction zone were deformed by friction at the
base of the South American Plate. In addition, seismic data also indicate the shallower bathymetry of the trench
in the northern zone, ∼4.7 km deep, compared with the central and southern zones, ∼5.2 km deep. The segment
of seismic line SIS-72 was taken from Collot et al. (2011) and re-interpreted in this study.

LAJO-YÁÑEZ ET AL. 12 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 8. (a) Uninterpreted and interpreted southwest-northeast seismic line 93-25, to the southwest of the Lobos de Tierra island, in the northern Peruvian forearc,
crossing the Chiclayo canyon (left-hand side), where the Nautile submersible (Sosson et al., 1994) sampled part of the seafloor made up of Paleozoic black schists
and quartzites lithologically similar to the basement drilled in the Talara oilfields and outcropping in the Coastal Range. The sub-sea outcrop implies a large-scale
uplifted structure and supports the fault-propagation anticlines from seismic interpretation. (b) Zoomed-in view of an Illescas fault-propagation anticline. OCB,
Oceanic crust basement; Pz+, Paleozoic basement and older units; K-P, Upper Cretaceous and Paleocene; E, Eocene; O, Oligocene; M, Miocene; P, Pliocene; BSR,
Bottom-simulating reflector.

Lines SIS-72 and 93-10, in the northern and central zones, respectively (Figures 11a and 11b), image the trench
fill with moderate-amplitude, subparallel and continuous seismic reflections. South of the seismic surveys,
turbidite successions were cored in local basins around the Trujillo Ridges, during the Nazca Project campaign
(Prince et al., 1974), where a pilot core of ∼5 m length and shorter piston cores recovered an intercalation of clay
and turbidites of Pleistocene—Holocene age. The silty base is composed of quartz, feldspars, rock fragments,
and organic matter, which indicates a continental source. Seismic data from this area and around the Mendaña
fracture zone acquired also in the SEAPERC campaign (Huchon & Bourgois, 1990), show similar seismic facies
to those found in lines SIS-72 and 93-10, which suggests the occurrence of turbidite deposits in the northern
and central zones of the northern Peru Trench. By contrast, in the south, seismic facies of the trench fill in line
93-16 (Figure 11c), are discontinuous with low-to-moderate amplitudes in a chaotic pattern, which suggests the
development of MTDs, characteristic of an unstable submarine slope (Posamentier & Martinsen, 2011), as seen
in Ecuador to the north, off the study area (Ratzov et al., 2010). These trench deposits are in front of the MTDs
identified in the middle and lower slopes at the same latitude (Bourgois et al., 1993), and complete the longitudi-
nal section of the MTD complex through line 93-16.

In central (line 93-10) and southern (line 93-16) zones, a subgroup of normal faults (white lines in Figure 11) cut
the oceanic crust to the base of the trench fill, and do not rupture the seafloor, either because they do not have a
displacement great enough or they have been buried by younger oceanic strata. In addition, thicker oceanic depos-
its are preserved in some grabens indicating syn-sedimentary fault movement and in other cases with upward
continuation of the fault with shorter displacement, suggesting reactivation periods. These features demonstrate
that the fault density is much more than what is captured by bathymetry data alone.

LAJO-YÁÑEZ ET AL. 13 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 9. (a) Uninterpreted and interpreted south-north seismic line 93-20, along the northern Peruvian forearc, showing the Peru growth fault, to the north, and the
Illescas fault-propagation anticline, to the south. (b) Zoomed-in view of the Illescas fault-propagation anticlines affecting the continental crust and Paleozoic basement.
(c) Zoomed-in view of the Peru growth fault and the anticlines developed in the footwall as a result of its strike-slip reactivation. For location of the line see Figure 2.
Pz+, Paleozoic basement and older units; K-P, Upper Cretaceous and Paleocene; E, Eocene; O, Oligocene; M, Miocene; P, Pliocene; BSR, Bottom-simulating reflector.

6.2. Continental Margin Wedge

The degree of margin accretion has been represented in conceptual models (Clift & Vannucchi, 2004; Cloos &
Shreve, 1988; Noda, 2016; von Huene et al., 2004; von Huene & Scholl, 1991). In the north, near the Peru-Ecuador
border, part of line SIS-72, shows a poorly-developed accretionary prism with a duplex structure within the Pale-
ozoic basement (yellow lines in Figure 11a). In the central zone, based on line 93-10, a small accretionary prism
is identified with two duplex structures in the margin wedge (yellow lines in Figure 11b), involving the Paleozoic
basement and the Upper Cretaceous-Paleocene mega-sequence. This is the sector that shows the highest degree
of accretion although it is still incipient. The southern zone, represented by line 93-16, is where no accretion has
been produced, so that the South American Plate edge abuts the trench with no frontal thrust (Figure 11c). These
characteristics at the three different latitudes show a spatial variability of the margin wedge accretion along the
northern Peruvian margin.

There is a proportional relationship between the thickness of trench sediments (≥1 km) entering the subduction
zone and the amount of accretion (Clift & Vannucchi, 2004; Scholl et al., 2015; Scholl & von Huene, 2007;
von Huene & Scholl, 1991). However, the study area shows that not only the subducting trench fill is involved
in accretion, but also oceanic crust deposits (blue and cyan transparent layers in Figure 11). Thus, in the north-
ern zone, the very small accretionary prism, 5 km wide, is consistent with the thickness of subducting oceanic
crust deposits plus trench sediments less than 1 km thick. In the central zone, incipient accretion is greater than in
the north, and correlates with the thickest oceanic crust deposits along with the trench fill. This material reaches
a thickness of 2.2 km at the plate boundary and shows a wrinkled appearance, likely related to friction with the
overriding plate during subduction. In the southern zone, where there is no accretion, oceanic deposits plus trench
fill are thinner, averaging ∼0.7 km. This relationship seen in the three zones depicts the importance of sediment
subduction in the subduction process.

6.3. Subduction Channel

The subduction channel is a layer of viscous material sandwiched between the base of the overriding plate and
the downgoing slab and acts as a lubricant in the subduction process (Cloos & Shreve, 1988). When oceanic plate
deposits, trench fill, and lower slope sediments enter the subduction zone in a subduction erosion environment,

LAJO-YÁÑEZ ET AL. 14 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 10. West-east geological profiles from seismic lines along the northern Peruvian forearc. Most of the lines intersect with south-north line 93-20 (red dashed
line). The Talara Basin extends from the Coastal Range westward to the Peru Trench. High-density normal faulting has resulted in a forearc collapse configuration.
The two main structural elements identified in the forearc are the Peru growth fault (thick dark line) to the north and the Illescas fault-propagation anticlines (thick
light-green line) to the south, which are part of the crustal shortening that locally uplifted the forearc.

LAJO-YÁÑEZ ET AL. 15 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 11. Trench zoomed-in views of seismic lines (a) SIS-72 in the north, (b) 93-10 in the center and (c) 93-16 in the south. The close-ups in lines SIS-72 and
93-10 show the only minor development of the accretionary prism, with less than 10 km of duplex formation starting from the trench. The surrounding low degree of
deformation is marked by short-displacement reverse faults (red lines). Line 93-16 shows no accretion to the south, and mass-transport deposits are identified in the
slope and trench fill, represented by chaotic seismic facies.

LAJO-YÁÑEZ ET AL. 16 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

they become part of the subduction channel (Scholl & von Huene, 2010) as viscous material with high pore
fluid pressure along the subduction zone (Shreve & Cloos, 1986; Vannucchi et al., 2012) that hydrofractures
the overriding plate, dislodging and dragging fragments due to active interface mega-thrusting (Kukowski &
Oncken, 2006; von Huene et al., 2004).

In the study area, seismic data allow tentative interpretation over the first ∼20 km of the subduction interface
from the trench, and suggest that the subduction channel underlies the mega-thrust plane with variable thickness
and extent. It was possible to delineate the subduction channel as a single and undifferentiated layer beneath the
South American Plate margin (pale yellow layer in all the seismic lines). The comparative analysis in Figure 10
involves the variation in thickness of the subduction channel. Northern and central areas show the thickest section
with ∼1 and ∼1.3 km, respectively; while it is much thinner in the south, reaching ∼0.3 km, which is consistent
with the total thickness of subducting oceanic crust deposits and trench sediments. This relationship also reflects
how important is the process of sediment subduction in subduction erosion.

6.4. Deformation at Subduction Entry

A set of high-angle reverse faults with short displacement (red lines in Figure 11) is affecting the South American
Plate margin wedge, the trench fill, and the oceanic crust deposits at the subduction entry; mostly in the northern
and central zones. These faults are continuous across the subduction plane and are consistent with a later tectonic
event, unrelated to offscraping. Compressional stress is interpreted to have produced this brittle deformation in
response to the difficulty in the advance of the Nazca Plate at subduction entry. Only a few of these reverse faults
are developed a little further landward from the subduction entry. In the south, a conspicuous listric normal fault
(black curved lines in the center of Figure 11c), described in previous sections, affects the subduction entry,
cutting the continental margin wedge and dipping into the oceanic crust deposits as a result of a gravity-driven
slide, which is consistent with extensional stress and more efficient subduction erosion.

7. Discussion
7.1. The Nazca Sliver Escape

The seafloor magnetic anomaly map from Maus et al. (2009) shows structural deformation in the northern part
of the Nazca Plate (Figure 12a), herein named the Talara Segment. In the south of this segment, a right-lateral
shift of the magnetic anomalies, aligned to the Mendaña fracture zone, clearly delineates a dextral displacement
of the Virú fracture zone (red bold line in Figure 12b). Therefore, this large structure has been interpreted as a
shear zone that appears to be linked to the reverse-fault structures of the Trujillo Ridges that fit with a change in
trajectory of the strike-slip kinematics to the north-northeast (orange jagged lines in Figure 12b), confirming the
observations of Lonsdale (2005). The similarity of alignment and geometry between the Trujillo and Chimbote
Ridges further south suggests the same causative strain, therefore, we assign these ridges to the same system. In
the north, the Grijalva Ridge, interpreted as a shear zone, is overprinted by a left-lateral displacement of magnetic
anomalies. This structure bounds the oceanic crust portion to the north, delimited also by the southern shear
deformation, which led to the interpretation of the Nazca Sliver (center of Figures 12b and 12c).

The seafloor magnetic anomalies map also shows two areas with different anomaly alignments to the north
and west of the Nazca Plate (Figure 12a). Based on integration with the seafloor ages from Müller et al. (2008)
and Farallon/Nazca Plate-drift reconstruction from Pardo-Casas and Molnar (1987), it is interpreted that in the
Oligocene (∼26 Ma), when the Farallon Plate split into the Nazca and Cocos Plates, creation of a microplate was
initiated and, at the same time, a second microplate formed from the East Pacific Rise (mid-ocean ridge). These
two new sections of oceanic crust expanded and pushed the Talara Segment of the Nazca Plate, one to the south
and the other to the east. We interpret that the resultant stress expelled the Nazca Sliver in a northeast direction.

7.2. The Northwest Peruvian Transpressional System

The geometry and kinematics of the Nazca Sliver correlate with the major structures of the northern Peruvian
margin in a transpressional model. In the south, the northeast projection of the Virú-Trujillo dextral-shear struc-
ture follows the Illescas fault-propagation anticlines (orange and light-green jagged lines in Figure 12c, respec-
tively), deviating the structural alignment to the north-northwest. Continuing to the north, the other element

LAJO-YÁÑEZ ET AL. 17 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Ki-v

Figure 12. (a) 2-arc-minute resolution magnetic anomaly grid (Maus et al., 2009) of oceanic crust off Peru and Ecuador continental margins, and (b) zoomed-in view.
(c) Interpreted present-day configuration of oceanic microplates, formed since Farallon Plate split, which pushed the Talara Segment of the Nazca Plate, producing the
Nazca Sliver. The two main structures that deformed the continental crust are the Peru fault (cyan line) to the north, and the Virú-Illescas transpressional system (red,
orange- and light-green-jagged lines) to the south. Seafloor ages adapted from Müller et al. (2008). FZ, fracture zone; NAS, North Andean Sliver; SZ, shear zone.

interpreted as part of the transpressional system is the Peru growth fault (cyan line in Figure 12c), which was
incorporated into the system when it underwent dextral strike-slip reactivation. Based on the angular uncon-
formity of the Miocene strata over the deformed Cretaceous, Eocene and Oligocene mega-sequences, the main
compressional strain of the sliver was produced probably in the Middle Miocene, as a result of an increase in
plate-convergence rate and a change in movement direction of the Nazca Plate.

To the northwest of the Peru fault, continental crust is folded and thickened, uplifting the area that was partially
eroded, and giving rise to the shallow-marine plateau of the Peru Bank. This crustal shortening is seen until the
Pliocene mega-sequence, which indicates that the last major compression by the Nazca Sliver occurred in the
Peru fault footwall during the Pliocene-Pleistocene transition. The Yáñez, Alvarado, and Sarmiento Ridges are
located within the Nazca Sliver, where the intersection of the first two with the localized deviations identified
in the bend-faulting alignment around the uplifted area of the trench (between latitudes of seismic lines 93-01
and 93-10) would indicate a dextral strike-slip kinematic behavior. Therefore, we interpret the ridges as minor
shear zones. Further northeast, projection of the Northwest Peruvian transpressional system coincides with the
deep and active Puná-Pallatanga fault analyzed by Tamay et al. (2021) (upper right side of Figure 12c). Earth-
quake focal mechanisms determined its dextral strike-slip behavior from the Gulf of Guayaquil to the southern

LAJO-YÁÑEZ ET AL. 18 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Western Cordillera (Ecuadorian Northern Andes). The shared trajectory and kinematics with the Peru fault
suggest a transference of deformation from the Nazca Sliver to the North Andean Sliver (NAS), in northwestern
Ecuador. Some small normal, reverse, and even reactivated faulting identified in the youngest strata suggest a
compressional-extensional strain related to the Nazca Sliver to date.

Using the GPS velocity field map by Nocquet et al. (2014), shown as orange and red arrows in Figure 12c,
three seismological sectors are differentiated. The NAS has a clear motion toward the east of 15 mm year −1.
However, the middle sector, involving southern Ecuador and northern Peru, has a minor motion to the southeast
of about 5 mm year −1, coinciding with the seismicity gap in the study area and the position in front of the Nazca
Sliver and the transpressional system. To the south, plate motion resembles that of the NAS; therefore, variable
along-margin locking is interpreted as related to the stress deviation of the Nazca Sliver to the northeast.

7.3. Seismological Expressions of the Transpressional System Activity

Along convergent margins, earthquakes occur as ruptures that propagate along the mega-thrust in a subduction
complex (McCann et al., 1979). The Andean plate margin has experienced many earthquakes during historical
times but there are some segments referred to as seismicity gaps. This is the case in northern Peru and south-
ern Ecuador, where no Mw > 8 subduction-related earthquake is known to have occurred (Nocquet et al., 2014;
Villegas-Lanza et al., 2016a). The major seismological activity on the northern Peruvian margin is represented
by two mega-thrust earthquakes recorded in 1953 (Espinoza, 1992) and 1959 (Ioualalen et al., 2014), with Mw 7.3
and Mw 7.5, and at focal depths of 33 km and >20 km, respectively (rupture areas represented by red ellipses in
Figure 13a). The two epicenters are aligned to the Peru fault azimuth. Other two significant mega-thrust events
are the tsunami earthquakes that occurred in the south of the study area (Nocquet et al., 2014; Villegas-Lanza
et al., 2016a; rupture areas represented by blue ellipses in Figure 13a); one in 1960 with Mw 7.6 and at a focal
depth between 5 and 25 km (Pelayo & Wiens, 1990), located within the area of the Illescas fault-propagation anti-
clines, and the other in 1996 with Mw 7.5 and around 7–10 km deep (Ihmlé et al., 1998), in front of the Chimbote
Ridges, close to the trench. The focal mechanism suggests a low-angle thrusting for the two tsunami earthquakes.

Duperret et al. (1995) defined the two major areas of earthquakes at a depth greater than 6 km recorded by the
world seismic network, from 1981 to 1992 on the northern Peruvian margin (Figure 13b), which coincided with
the area of the Peru fault and the Illescas fault-propagation anticlines. The focal mechanism method determined
reverse and strike-slip fault kinematics in the north, and reverse in the south. These data confirm the relationship
between two main areas of earthquakes in the study area and the transpressional system of the northern Peruvian
forearc.

Daudt et al. (2009) reported an earthquake of Mw 4.7 on 22 April 2007 (Figure 13c), with an epicenter located
in the Peru fault area, altering oil seepages in the Talara region, and increasing temporarily oil production from
its oilfields. From February to September 2009, four periods of earthquakes were recorded to the west of the
Illescas hills by the National Seismic Network of Peru (Figure 13c). Villegas-Lanza et al. (2016b) identified
a reverse kinematic stress followed by a 5-month period of relative quiescence, and then a final event with
normal-kinematic stress. This seismically active area is located exactly over the Illescas fault-propagation anti-
clines interpreted in this work, thus, the first and the third seismic events, which occurred at 8–12 km depth,
confirm the existence and activity of these large structures.

7.4. Tectonic Configuration of the Talara Basin

A chronostratigraphic chart, shown in Figure 14, was constructed to analyze the tectonic factors involved in the
evolution of the Talara Basin within the geodynamics of the northern Peruvian forearc. The chronostratigraphic
logs of the Talara Basin built from northern, central, and southern seismic interpretations, and calibrated with
well data, are integrated with the oxygen-isotope-derived eustatic sea-level history in the Pacific region (Kominz
et al., 2008; K. G. Miller et al., 2020) and the plate-convergence rate of the eastern Pacific margin through the
Mesozoic and Cenozoic (Larson & Pitman, 1972; Pardo-Casas & Molnar, 1987).

Throughout the Ypresian, Oligocene and Miocene, the long-term eustatic sea-level curve is in phase with
plate convergence rate. Periods of concomitant sea-level rise and increase in plate convergence rate may have
been amplified by an increase in mid-ocean ridge spreading (Flemming & Roberts, 1973; Mörner, 1980; Vail

LAJO-YÁÑEZ ET AL. 19 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 13. The transpressional system interpreted in the northern Peruvian forearc includes the Peru fault to the north
(thick cyan line) and the Illescas fault-propagation anticline (thick green jagged lines). (a) Rupture areas of all the largest
earthquakes (>7.3 Mw) registered in the forearc; two occurred in the north in 1953 and 1959 (red ellipses), and two tsunami
earthquakes in the south in 1960 and 1996 (light blue ellipses). (b) Focal mechanisms from earthquakes recorded between
1981 and 1992, showing reverse and strike-slip kinematics to the north, and reverse to the south. (c) Earthquake registered in
2007 to the north, and earthquake monitoring through 2009 with focal mechanism of the four largest to the south. Note that
all the seismological expressions compiled in all the different periods are concentrated in two main areas coinciding with the
Peru fault to the north and the Illescas fault-propagation anticline to the south. Geological features are described in Figure 2c.
FZ, fracture zone; SZ, shear zone.

et al., 1984). In the same way, the greatest sea-level fall periods match with plate convergence rate decrease. This
close correlation between both curves is consistent with a tectonic component to relative sea level behavior from
the Eocene onwards. Data integration allowed the construction of a subduction-erosion rate curve throughout
the Talara and younger basins since the plate-convergence rate specifically represents the Farallon/Nazca Plate
motion.

Based on regional stratigraphic studies (Euribe, 1976; Gonzales, 1976) and benthic foraminifera documentation
(Cushman & Stone, 1949; Frizzell, 1943; Stainforth, 1954; Vegas, 1970; Weiss, 1955), the Talara Basin can be
subdivided in a lower succession, deposited from the Campanian to the Danian, of marine shelf deposits over-
lain by deep-water sediments, passing upwards into marginal marine facies. The upper basin fill, of Ypresian to

LAJO-YÁÑEZ ET AL. 20 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 14. Chrono-tectonostratigraphic chart of the northern Peruvian forearc showing the relationship of Talara Basin development with the plate convergence rate
and eustatic sea-level curves. The basin was formed during a period of subduction erosion coincident with an abrupt increase in the convergence rate of the Nazca and
South American Plates. During this period, instability in the forearc also increased due to a high incidence of uplift-collapse events preserved in the stratigraphic record
as episodes of extensional fault-controlled sedimentation. Eustatic sea level started to track convergence rate, possibly due to an increase of mid-ocean ridge spreading.

Priabonian deposits, shows an alternation of shallow- and deep-water sequences. The transition between these
two sections coincides with the increase in plate-convergence rate in the Ypresian. The lower stratigraphy is
consistent with a gradually subsiding forearc with a long-term eustatic sea-level fall and no evidence of specific
tectonic activity. Seismic interpretation shows that the upper section is characterized by fault-controlled sedi-
mentation, where uplift-collapse events took place, resulting in subsidence acceleration mostly in the central
and northern areas. Variability in geological settings throughout this period reflects the high degree of tecto-
nism, which is related to the abrupt increase in plate convergence rate from 5 to 15 cm year −1, with a rate of
12 cm year −1 through part of the Middle Eocene to the end of the Late Eocene. This high rate through almost all
the Eocene represents the Farallon Plate drift acceleration, interpreted as the period with the maximum subduc-
tion erosion rate. Therefore, the geodynamic transition in the Eocene marks an important geological change in the
Talara Basin stratigraphy.

7.5. 3-D Tectono-Stratigraphic Evolution of the Northern Peruvian Forearc

Forearc, subduction entry and seafloor characterization, and plate-tectonic data have been integrated to propose
a six-stage 3-D evolutionary model for the northern Peru forearc (Figure 15):
1. Initial Triassic magmatic arc development with minor forearc subsidence, which did not allow any basin
development.
2. Development of the marginal-rift Lancones-Alamor Basin related to subduction due to slab roll-back (Winter
et al., 2010) in the inner forearc from Early Cretaceous magmatic arc emplacement until the Late Cretaceous.
3. Late Cretaceous Redondo Formation extended from the Lancones Basin to the outer forearc due to subsidence
triggered by initiation of subduction erosion. A significant increase in subsidence rate in the Early Eocene,

LAJO-YÁÑEZ ET AL. 21 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Figure 15. 3-D geodynamic evolution of the northern Peruvian forearc, in a scenario of a trench and abandoned magmatic arc interaction. (a) No forearc
accommodation during the Triassic magmatic arc formation. (b) Volcano-sedimentary Lancones-Alamor Basin in the inner forearc during Albian aulacogen. (c)
Eocene outer forearc collapse due to major margin consumption, during major subduction-erosion regime. (d) Outer forearc crustal shortening (transpressional system:
PF, Peru fault; IA, Illescas anticlines) produced by the Nazca Sliver northeastward escape. (e) Miocene sedimentation over the deformed forearc through an angular
unconformity after a period of crustal shortening. (f) Intermittent transpressive system activity at the end of the Pliocene, folding the Peru fault footwall.

due to increase in plate convergence rate and subduction erosion, produced westward depocentre migration to
the outer forearc for the Talara Basin (Late Cretaceous—Paleocene and Eocene mega-sequences), bounded to
the east by the Coastal Range (abandoned Triassic magmatic arc).
4. An abrupt reduction in plate convergence rate occurred in the Eocene-Oligocene transition, resulting in major
forearc collapse to the north and south of the northern Peruvian forearc, which coincides with the end of
deposition in the Talara Basin and initiation of the Tumbes and Sechura basins (Oligocene mega-sequence).
As part of this collapse, development of the Peru fault started to form the large-scale half-graben in the north-
west, which provided accommodation for the basal Early Miocene mega-sequence. A subsequent increase in
plate-convergence rate set up the transpressional regime and triggered escape of the Nazca Sliver through the
Northwest Peruvian Transpressional System (Illescas fault-propagation anticlines and strike-slip-reactivated
Peru fault), resulting in crustal shortening across the outer forearc since the Middle Miocene.
5. The Miocene mega-sequence onlapped, through angular unconformity, the outer forearc, which was deformed
by the transpressional system.
6. Transpressional activity continued through the Pliocene, folding and uplifting the Peru fault footwall area,
with subsequent erosion resulting in the Peru Bank as a current offshore geomorphological element. Since
then, an alternation of compressional stress and extensional release periods led to episodes of forearc uplift
and subsidence, respectively.

7.6. Evidence for Long-Term Subduction Erosion

Critical indicators of long-term subduction erosion at the northern Peruvian margin since the Early Eocene include
forearc subsidence along with gravitational normal faulting, due to thinning of the overriding plate edge by basal
tectonic erosion, and minor or no accretionary prism development. These characteristics are well represented

LAJO-YÁÑEZ ET AL. 22 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

in the main subduction-erosion models (Clift & Vannucchi, 2004; Cloos & Shreve, 1988; Noda, 2016; Scholl
et al., 1980; von Huene et al., 2004), which are important in understanding the spatial and temporal distribution
of accommodation in forearc basins and fault activation through uplift-collapse events.

Additional evidence for long-term subduction erosion is the magmatic arc retreat (Kukowski & Oncken, 2006), in
the context of a non-variable dip angle for the subducted slab. The model addresses the distance between a trench
and an abandoned magmatic arc as related to continental margin truncation (Scholl & von Huene, 2010). This
setting can play an important role in the configuration over time of the forearc basin, showing how the frontal and
basal erosion of the continental margin leads to cannibalization of parts of the forearc basin and basement, with
area reduction through geological time (Moore, 2001), producing landward trench migration. As the margin front
is consumed, the active arc retreats leaving older emplaced intrusive massifs closer to the trench. In the northern
Peruvian forearc, the shortest distance between the outcropping Triassic magmatic arc (Coastal Range) and the
current trench is ∼85 km. In the subsurface, this distance is much shorter at ∼35 km from the front of the Triassic
intrusive body to the subduction zone, which highlights the intense subduction-erosion activity.

8. Conclusions
The northern Peruvian forearc represents a continental margin configured by subduction erosion with marked
along-margin variability. In the south, there is no accretion, while the central and northern areas exhibit a
poorly-developed accretionary prism. The margin is characterized by clear truncation (margin consumption) and
subsidence by basal material loss, considered the main mechanism of forearc basin accommodation.

Second-order discontinuities identified along the forearc include a nonconformity separating the Paleozoic meta-
sedimentary basement from overlying sedimentary sequences and an angular unconformity between the Talara
and younger basins. The Cretaceous—Paleocene section of the Talara Basin was controlled by glacio-eustasy,
while in the Eocene, sedimentation was strongly influenced by episodes of uplift followed by extensional
collapse, moderated by tectono-eustatic sea-level changes. Absence of a basin-scale unconformity between these
mega-sequences suggests a continuous long-term subduction-erosion process overprinted by gravity-driven
normal faulting. Uplift episodes are related to short periods of Nazca Plate advance during the Oligocene and
Miocene.

The Nazca Sliver is a portion of oceanic crust identified to the north of the Nazca Plate (Talara Segment). It is
interpreted to have formed and escaped by pushing of microplates of the East Pacific Rise development from
the Oligocene (∼26 Ma, when the Farallon Plate split into the Nazca and Cocos Plates) and an acceleration in
mid-ocean ridge spreading rate in the Early Miocene (∼20 Ma). During this period, crustal shortening across the
forearc took place, producing the northwest Peruvian transpressional system, represented by the configuration
of the strike-slip-reactivated Peru fault and Illescas fault-propagation anticlines. Decrease in the rate of plate
convergence since the Late Miocene (∼12 Ma) implies that the rate of north-eastward escape of the Nazca Sliver
has also been reduced in northern Peru since then.

The northern Peruvian margin provides a model for understanding spatial and temporal variability in structural
development and forearc basin responses to subduction zone processes over ∼80 My. The Peruvian model may
have application in better understanding of forearc tectono-stratigraphic relationships in more ancient convergent
margin complexes.

Conflict of Interest
The authors declare no conflicts of interest relevant to this study.

Data Availability Statement


Data sets used for this research are available in these in-text data citation references: Müller et al. (2008), Age, spread-
ing rates and spreading symmetry of the world's ocean crust map (https://www.ngdc.noaa.gov/mgg/ocean_age/
ocean_age_2008.html). Maus et al. (2009), EMAG2: A 2–arc min resolution Earth Magnetic Anomaly Grid (https://
geomag.colorado.edu/emag2-earth-magnetic-anomaly-grid-2-arc-minute-resolution.html). Ryan et al. (2009),
Global Multi-Resolution Topography synthesis (https://www.gmrt.org/GMRTMapTool/). Geersen (2019),

LAJO-YÁÑEZ ET AL. 23 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Collated bathymetric data from convergent margins that experienced tsunami earthquakes. PANGAEA (https://
doi.org/10.1594/PANGAEA.899049).

Acknowledgments References
This work was carried out as part of the
first author's PhD program at the Univer- Baldry, R. A. (1938). Slip-planes and breccia zones in the Tertiary rocks of Peru. Quarterly Journal of the Geological Society of London, 376(4),
sity of Manchester, United Kingdom, 347–363. https://doi.org/10.1144/GSL.JGS.1938.094.01-04.12
which was sponsored by the Peruvian Barazangi, M., & Isacks, B. L. (1976). Spatial distribution of earthquakes and subduction of the Nazca plate beneath South America. Geology,
National Programme of Scholarships and 4(11), 686–692. https://doi.org/10.1130/0091-7613(1976)4<686:SDOEAS>2.0.CO;2
Educational Loans (PRONABEC). The Barazangi, M., & Isacks, B. L. (1979). Subduction of the Nazca plate beneath Peru: Evidence from spatial distribution of earthquakes. Geophys-
authors thank Perupetro S.A. for all the ical Journal International, 57(3), 537–555. https://doi.org/10.1111/j.1365-246X.1979.tb06778.x
subsurface data supplied for this study. Bellido, F., Valverde, P., Jaimes, F., Carlotto, V., & Díaz-Martínez, E. (2009). Datación y caracterización geoquímica de los granitoides
IHS is thanked for the provision of King- peralumínicos de los cerros de Amotape y de los Macizos de Illescas y Paita (Noroeste de Perú). Boletín de la Sociedad Geológica del Perú,
dom software. A special thanks to Prof. 103, 197–213.
Peter Clift for insights that contributed Borda, E., & Bianchi, C. (2020). The role of 2D/3D seismic data in evaluating the hydrocarbon prospectivity of Peruvian offshore basins. Pacific
to this work. The authors are grateful Basins Virtual Research Symposium. American Association of Petroleum Geologists, Latin America & Caribbean Region. Retrieved from
to Serge Lallemand, Jacob Geersen https://store.aapg.org/detail.aspx?id=VRSK-58113
and two anonymous reviewers for their Bourgois, J., Bigot-Cormier, F., Bourles, D., Braucher, R., Dauteuil, O., Witt, C., & Michaud, F. (2007). Tectonic record of strain buildup and
constructive reviews with comments and abrupt coseismic stress release across the northwestern Peru coastal plain, shelf, and continental slope during the past 200 kyr. Journal of
suggestions that substantially improved Geophysical Research, 112(B4), 1–22. https://doi.org/10.1029/2006JB004491
this paper. Tectonics Editor in Chief Bourgois, J., Lagabrielle, Y., Wever, P. D., & Suess, E. (1993). Tectonic history of the northern Peru convergent margin during the past 400 ka.
Taylor Schildgen and Associate Editor Geology, 21(6), 531–534. https://doi.org/10.1130/0091-7613(1993)021<0531:THOTNP>2.3.CO;2
Paola Vannucchi are thanked for the Bourgois, J., Pautot, G., Bandy, W., Boinet, T., Chotin, P., Huchon, P., et al. (1988). Seabeam and seismic reflection imaging of the tectonic
evaluation and editorial handling of the regime of the Andean continental margin off Peru (4°S to 10°S). Earth and Planetary Science Letters, 87(1–2), 111–126. https://doi.
manuscript. org/10.1016/0012-821X(88)90068-4
Bush, V. A., Vinogradov, L. D., & Titov, A. I. (1994). Tectonic breccia and thrust tectonics of the Tertiary deposits in northwestern Peru. Geotec-
tonics, 28(2), 159–168.
Clift, P. D., Pecher, I., Kukowski, N., & Hampel, A. (2003). Tectonic erosion of the Peruvian forearc, Lima Basin, by subduction and Nazca Ridge
collision. Tectonics, 22(3), 1–16. https://doi.org/10.1029/2002TC001386
Clift, P. D., & Vannucchi, P. (2004). Controls on tectonic accretion versus erosion in subduction zones: Implications for the origin and recycling
of the continental crust. Reviews of Geophysics, 42(2), 1–31. https://doi.org/10.1029/2003RG000127
Cloos, M., & Shreve, R. L. (1988). Subduction-channel model of prism accretion, melange formation, sediment subduction, and subduc-
tion erosion at convergent plate margins: 1. Background and description. Pure and Applied Geophysics, 128(3–4), 455–500. https://doi.
org/10.1007/BF00874548
Coats, R. R. (1962). Magma type and crustal structure in the Aleutian arc. In G. A. MacDonald & H. Kuno (Eds.), The crust of the Pacific Basin,
Geophysical Monograph Series 6 (pp. 92–109). American Geophysical Union. https://doi.org/10.1029/GM006p0092
Collot, J. Y., Agudelo, W., Ribodetti, A., & Marcaillou, B. (2008). Origin of a crustal splay fault and its relation to the seismogenic zone and
underplating at the erosional north Ecuador–south Colombia oceanic margin. Journal of Geophysical Research, 113(B12), 1–19. https://doi.
org/10.1029/2008JB005691
Collot, J. Y., Ribodetti, A., Agudelo, W., & Sage, F. (2011). The South Ecuador subduction channel: Evidence for a dynamic mega-shear zone
from 2D fine-scale seismic reflection imaging and implications for material transfer. Journal of Geophysical Research, 116(B11), 1–20.
https://doi.org/10.1029/2011JB008429
Contreras-Reyes, E., Muños-Linford, P., Cortés-Rivas, V., Bello-González, J. P., Ruiz, J. A., & Krabbenhoeft, A. (2019). Structure of the colli-
sion zone between the Nazca Ridge and the Peruvian convergent margin: Geodynamic and seismotectonic implications. Tectonics, 38(9),
3416–3435. https://doi.org/10.1029/2019TC005637
Contreras-Reyes, E., Ruiz, J. A., Becerra, J., Kopp, H., Reichert, C., Maksymowicz, A., & Arriagada, C. (2015). Structure and tectonics of the
central Chilean margin (31°–33°S): Implications for subduction erosion and shallow crustal seismicity. Geophysical Journal International,
203(2), 776–791. https://doi.org/10.1093/gji/ggv309
Coxall, H. K. (2000). Hantkeninid planktonic foraminifera and Eocene palaeoceanographic change (PhD thesis). University of Bristol.
Cushman, J. A., & Stone, B. (1949). Foraminifera from the Eocene Chacra formation of Peru. Contribution From the Cushman Laboratory for
Foraminiferal Research, 25(3), 49–58.
Daudt, J. A., Benedicto, A., & Pozo, E. (2009). Migração de petróleo induzida por sismicidade: Observações de campo e de subsuperfície na
Bacia de Talara (noroeste do Peru). Boletim de Geociências da Petrobras, 17(2), 371–374.
DeVries, T. J. (1988). The geology of late Cenozoic marine terraces (tablazos) in northwestern Peru. Journal of South American Earth Sciences,
1(2), 121–136. https://doi.org/10.1016/0895-9811(88)90030-2
Dickinson, W. R., & Seely, D. R. (1979). Structure and stratigraphy of forearc regions. American Association of Petroleum Geologists Bulletin,
63(1), 2–31. https://doi.org/10.1306/C1EA55AD-16C9-11D7-8645000102C1865D
Duperret, A., Bourgois, J., Lagabrielle, Y., & Suess, E. (1995). Slope instabilities at an active continental margin: Large-scale poly-
phase submarine slides along the northern Peruvian margin, between 5 S and 6 S. Marine Geology, 122(4), 303–328. https://doi.
org/10.1016/0025-3227(94)00125-5
Espinoza, J. (1992). Terremotos tsunamigénicos en el Ecuador. Acta Oceanográfica del Pacífico, INOCAR, 7(1), 21–28. Retrieved from http://
hdl.handle.net/1834/2181
Euribe, A. (1976). Glossary of stratigraphic units of Northwestern Peru. Internal biostratigraphic report 76-01. Belco Petroleum Corporation of
Peru.
Fildani, A., Hessler, A. M., & Graham, S. A. (2008). Trench-forearc interactions reflected in the sedimentary fill of Talara Basin, northwest Peru.
Basin Research, 20(3), 305–331. https://doi.org/10.1111/j.1365-2117.2007.00346.x
Flemming, N. C., & Roberts, D. G. (1973). Tectono-eustatic changes in sea level and seafloor spreading. Nature, 243, 19–22. https://doi.
org/10.1038/243019a0
Frizzell, D. L. (1943). Upper Cretaceous foraminifera from northwestern Peru. Journal of Paleontology, 17(4), 331–353. Retrieved from https://
www.jstor.org/stable/1298991
Frizzell, D. L. (1945). Larger foraminifera of northwestern Peru and western Ecuador. Internal report. International Petroleum Company.

LAJO-YÁÑEZ ET AL. 24 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Galloway, J. J., & Morrey, M. (1929). A lower Tertiary foraminiferal fauna from Manta, Ecuador. Bulletin of American Paleontology, XV(55),
1–56.
Geersen, J. (2019). Sediment-starved trenches and rough subducting plates are conductive to tsunami earthquakes [Dataset]. PANGAEA.
Tectonophysics, 762, 28–44. https://doi.org/10.1594/PANGAEA.899049
Geersen, J., Sippl, C., & Harmon, N. (2022). Impact of bending-related faulting and oceanic-plate topography on slab hydration and
intermediate-depth seismicity. Geosphere, 18(2), 562–584. https://doi.org/10.1130/GES02367.1
Genge, M. C., Witt, C., Chanier, F., Reynaud, J. Y., & Calderon, Y. (2020). Outer forearc high control in an erosional subduction regime: The case
of the central Peruvian forearc (6–10°S). Tectonophysics, 789, 1–15. https://doi.org/10.1016/j.tecto.2020.228546
Gonzales, G. (1976). Ciclos de Sedimentación en el Eoceno de la Cuenca Talara. Boletín de la Sociedad Geológica del Perú, 51, 73–80.
Hampel, A. (2002). The migration history of the Nazca Ridge along the Peruvian active margin: A re-evaluation. Earth and Planetary Science
Letters, 203(2), 665–679. https://doi.org/10.1016/S0012-821X(02)00859-2
Haq, B. U. (2014). Cretaceous eustasy revisited. Global and Planetary Change, 113, 44–58. https://doi.org/10.1016/j.gloplacha.2013.12.007
Herbozo, G., Kukowski, N., Clift, P. D., Pecher, I., & Bolaños, R. (2020). Cenozoic increase in subduction erosion during plate convergence
variability along the convergent margin off Trujillo, Peru. Tectonophysics, 790, 1–19. https://doi.org/10.1016/j.tecto.2020.228557
Hessler, A. M., & Sharman, G. R. (2018). Subduction zones and their hydrocarbon systems. Geosphere, 14(5), 2044–2067. https://doi.
org/10.1130/GES01656.1
Higley, D. K. (2004). The Talara Basin province of northwestern Peru, Cretaceous-tertiary total petroleum system. US Geological Survey Bulle-
tin 2206-A. US Department of the Interior. Retrieved from https://pubs.usgs.gov/bul/2206/A/b2206-a.pdf
Hilde, T. W. (1983). Sediment subduction versus accretion around the Pacific. Tectonophysics, 99(2–4), 381–397. https://doi.
org/10.1016/0040-1951(83)90114-2
Hofker, J. (1956). Tertiary foraminifera of coastal Ecuador: Part II, Additional notes on the Eocene species. Journal of Paleontology, 30(4),
891–958. Retrieved from http://www.jstor.org/stable/1300429
Huchon, P., & Bourgois, J. (1990). Subduction-induced fragmentation of the Nazca Plate off Peru: Mendana Fracture Zone and Trujillo Trough
revisited. Journal of Geophysical Research, 95(B6), 8419–8436. https://doi.org/10.1029/JB095iB06p08419
Iddings, A., & Olsson, A. A. (1928). Geology of northwest Peru. American Association of Petroleum Geologists Bulletin, 12(1), 1–39. https://doi.
org/10.1306/3D9327D7-16B1-11D7-8645000102C1865D
Ihmlé, P. F., Gomez, J. M., Heinrich, P., & Guibourg, S. (1998). The 1996 Peru tsunamigenic earthquake: Broadband source process. Geophysical
Research Letters, 25(14), 2691–2694. https://doi.org/10.1029/98GL01987
Ioualalen, M., Monfret, T., Béthoux, N., Chlieh, M., Ponce-Adams, G., Collot, J. Y., et al. (2014). Tsunami mapping in the Gulf of Guayaquil,
Ecuador, due to local seismicity. Marine Geophysical Researches, 35(4), 361–378. https://doi.org/10.1007/s11001-014-9225-9
Kita, S., & Ferrand, T. P. (2018). Physical mechanisms of oceanic mantle earthquakes: Comparison of natural and experimental events. Scientific
Reports, 8(1), 1–11. https://doi.org/10.1038/s41598-018-35290-x
Kominz, M. A., Browning, J. V., Miller, K. G., Sugarman, P. J., Mizintseva, S., & Scotese, C. R. (2008). Late Cretaceous to Miocene sea-level
estimates from the New Jersey and Delaware coastal plain coreholes: An error analysis. Basin Research, 20(2), 211–226. https://doi.
org/10.1111/j.1365-2117.2008.00354.x
Krabbenhöft, A., Bialas, J., Kopp, H., Kukowski, N., & Hübscher, C. (2004). Crustal structure of the Peruvian continental margin from wide-angle
seismic studies. Geophysical Journal International, 159(2), 749–764. https://doi.org/10.1111/j.1365-246X.2004.02425.x
Kukowski, N., & Oncken, O. (2006). Chapter 10. Subduction erosion – The “normal” mode of fore-arc material transfer along the Chilean
Margin? In O. Oncken, G. Chong, G. Franz, P. Giese, H. Götze, V. Ramos, et al. (Eds.), The Andes – Active subduction orogeny, Frontiers in
Earth Sciences (pp. 217–236). Springer. https://doi.org/10.1007/978-3-540-48684-8_10
Kulm, L. D., Thornburg, T. M., Schrader, H. J., & Resig, J. M. (1982). Cenozoic structure, stratigraphy and tectonics of the central Peru forearc.
In J. K. Leggett (Ed.), Trench-forearc geology: Sedimentation and tectonics on modern and ancient active plate margins, Geological Society,
London, Special Publications (Vol. 10, pp. 151–169). https://doi.org/10.1144/GSL.SP.1982.010.01.10
Lajo-Yáñez, J. A., Flint, S. S., Brunt, R. L., Huuse, M., Searle, S. R. A., & Sheppard, J. M. (2022). Disentangling tectonic and eustatic controls on
forearc basin stratigraphy, Talara Basin, Peru. Sedimentary Geology, 442, 1–25. https://doi.org/10.1016/j.sedgeo.2022.106277
Langseth, M. G., von Huene, R., Nasu, N., & Okada, H. (1981). Subsidence of the Japan Trench forearc region of northern Honshu. Oceanologica
Acta SP, 173–179. Retrieved from https://archimer.ifremer.fr/doc/00246/35686/
Larson, R. L., & Pitman, W. C., III. (1972). World-wide correlation of Mesozoic magnetic anomalies, and its implications. Geological Society of
America Bulletin, 83(12), 3645–3662. https://doi.org/10.1130/0016-7606(1972)83[3645:WCOMMA]2.0.CO;2
Lonsdale, P. (2005). Creation of the Cocos and Nazca plates by fission of the Farallon plate. Tectonophysics, 404(3–4), 237–264. https://doi.
org/10.1016/j.tecto.2005.05.011
Mammerickx, J., Anderson, R. N., Menard, H. W., & Smith, S. M. (1975). Morphology and tectonic evolution of the East-Central Pacific.
Geological Society of America Bulletin, 86(1), 111–118. https://doi.org/10.1130/0016-7606(1975)86<111:mateot>2.0.CO;2
Marsaglia, K. M., & Carozzi, A. V. (1990). Depositional environment, sand provenance, and diagenesis of the Basal Salina Formation (lower
Eocene), northwestern Peru. Journal of South American Earth Sciences, 3(4), 253–267. https://doi.org/10.1016/0895-9811(90)90007-N
Maus, S., Barckhausen, U., Berkenbosch, H., Bournas, N., Brozena, J., Childers, V., et al. (2009). EMAG2: A 2–arc min resolution Earth
Magnetic Anomaly Grid compiled from satellite, airborne, and marine magnetic measurements [Dataset]. Geochemistry, Geophysics, Geosys-
tems, 10(8), 1–12. https://doi.org/10.1029/2009GC002471
McCann, W. R., Nishenko, S. P., Sykes, L. R., & Krause, J. (1979). Seismic gaps and plate tectonics: Seismic potential for major boundaries. In
M. Wyss (Ed.), Earthquake prediction and seismicity patterns (pp. 1082–1147). Birkhäuser. https://doi.org/10.1007/BF00876211
Michaud, F., Witt, C., & Royer, J. Y. (2009). Influence of the subduction of the Carnegie volcanic ridge on Ecuadorian geology: Reality and
fiction. In S. M. Kay, V. A. Ramos, & W. R. Dickinson (Eds.), Backbone of the Americas: Shallow subduction, plateau uplift, and ridge and
terrane collision, The Geological Society of America Memoir 204 (pp. 217–228). https://doi.org/10.1130/2009.1204(10)
Miller, H. (1970). Das Problem des hypothetischen „Pazifischen Kontinentes“ gesehen von der chilenischen Pazifikküste. Geologische Rund-
schau, 59(3), 927–938. https://doi.org/10.1007/BF02042277
Miller, K. G., Browning, J. V., Schmelz, W. J., Kopp, R. E., Mountain, G. S., & Wright, J. D. (2020). Cenozoic sea-level and cryospheric evolution
from deep-sea geochemical and continental margin records. Science Advances, 6(20), 1–15. https://doi.org/10.1126/sciadv.aaz1346
Miller, K. G., Kominz, M. A., Browning, J. V., Wright, J. D., Mountain, G. S., Katz, M. E., et al. (2005). The Phanerozoic record of global
sea-level change. Science, 310(5752), 1293–1298. https://doi.org/10.1126/science.1116412
Moberly, R., Shepherd, G. L., & Coulbourn, W. T. (1982). Forearc and other basins, continental margin of northern and southern Peru and adja-
cent Ecuador and Chile. Geological Society, London, Special Publications, 10(1), 171–189. https://doi.org/10.1144/GSL.SP.1982.010.01.11

LAJO-YÁÑEZ ET AL. 25 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Moore, J. C. (2001). Accretionary prisms. In J. H. Steele (Ed.), Encyclopedia of ocean science (2nd ed., pp. 31–37). Academic Press. https://doi.
org/10.1016/B978-0-12-813081-0.00465-1
Mörner, N. A. (1980). Relative sea-level, tectono-eustasy, geoidal-eustasy and geodynamics during the Cretaceous. Cretaceous Research, 1(4),
329–340. https://doi.org/10.1016/0195-6671(80)90042-7
Müller, R. D., Sdrolias, M., Gaina, C., & Roest, W. R. (2008). Age, spreading rates and spreading symmetry of the world's ocean crust [Dataset].
Geochemistry, Geophysics, Geosystems, 9(4), 1–19. https://doi.org/10.1029/2007GC001743
Murany, E. (1975). Depositional model: Pariñas formation. Internal report. Belco Petroleum Corporation of Peru.
Murauchi, J. (1971). The renewal of island arcs and the tectonics of marginal seas. In M. Uda (Ed.), The Ocean world. Proceedings of joint
oceanographic assembly (pp. 303–305).
Murauchi, S., & Ludwig, W. J. (1980). Crustal structure of the Japan trench: The effect of subduction of ocean crust. In Initial reports of the deep-
sea drilling project, LVI-LVII, Part 1(10) (pp. 463–469).
Nocquet, J. M., Villegas-Lanza, J. C., Chlieh, M., Mothes, P. A., Rolandone, F., Jarrin, P., et al. (2014). Motion of continental slivers and creeping
subduction in the northern Andes. Nature Geoscience, 7(4), 287–291. https://doi.org/10.1038/ngeo2099
Noda, A. (2016). Forearc basins: Types, geometries, and relationships to subduction zone dynamics. Geological Society of America Bulletin,
128(5–6), 879–895. https://doi.org/10.1130/B31345.1
Noda, A., & Miyakawa, A. (2017). Chapter 1. Deposition and deformation of modern accretionary-type forearc basins: Linking basin formation
and accretionary wedge growth. In Y. Itoh (Ed.), Evolutionary models of convergent margins: Origin of their diversity (pp. 3–27). IntechOpen.
https://doi.org/10.5772/67559
Pardo-Casas, F., & Molnar, P. (1987). Relative motion of the Nazca (Farallon) and South American plates since Late Cretaceous time. Tectonics,
6(3), 233–248. https://doi.org/10.1029/TC006i003p00233
Pedoja, K., Ortlieb, L., Dumont, J. F., Lamothe, M., Ghaleb, B., Auclair, M., & Labrousse, B. (2006). Quaternary coastal uplift along the Talara
Arc (Ecuador, Northern Peru) from new marine terrace data. Marine Geology, 228(1–4), 73–91. https://doi.org/10.1016/j.margeo.2006.01.004
Pelayo, A. M., & Wiens, D. A. (1990). The November 20, 1960 Peru tsunami earthquake: Source mechanism of a slow event. Geophysical
Research Letters, 17(6), 661–664. https://doi.org/10.1029/GL017i006p00661
Peralta-Cárdenas, G. (1967). Structural analysis of northwestern Peru, South America (PhD thesis). Cornell University.
Posamentier, H. W., & Martinsen, O. J. (2011). The character and genesis of submarine mass-transport deposits: Insights from outcrop and 3D
seismic data. In R. C. Shipp, P. Weimer, & H. W. Posamentier (Eds.), Mass-transport deposits in deepwater settings, Society for Sedimentary
Geology Special Publication 96 (pp. 7–38). https://doi.org/10.2110/sepmsp.096.007
Prince, R. A., & Kulm, L. D. (1975). Crustal rupture and the initiation of imbricate thrusting in the Peru-Chile Trench. Geological Society of
America Bulletin, 86(12), 1639–1653. https://doi.org/10.1130/0016-7606(1975)86<1639:CRATIO>2.0.CO;2
Prince, R. A., Resig, J. M., Kulm, L. D., & Moore, T. C., Jr. (1974). Uplifted turbidite basins on the seaward wall of the Peru Trench. Geology,
2(12), 607–611. https://doi.org/10.1130/0091-7613(1974)2<607:UTBOTS>2.0.CO;2
Ramos, V. A., & Folguera, A. (2009). Andean flat-slab subduction through time. Geological Society, London, Special Publications, 327(1),
31–54. https://doi.org/10.1144/SP327.3
Ranero, C. R., Villaseñor, A., Phipps Morgan, J., & Weinrebe, W. (2005). Relationship between bend-faulting at trenches and intermediate-depth
seismicity. Geochemistry, Geophysics, Geosystems, 6(12), 1–25. https://doi.org/10.1029/2005GC000997
Ratzov, G., Collot, J. Y., Sosson, M., & Migeon, S. (2010). Mass-transport deposits in the northern Ecuador subduction trench: Result of frontal
erosion over multiple seismic cycles. Earth and Planetary Science Letters, 296(1–2), 89–102. https://doi.org/10.1016/j.epsl.2010.04.048
Ryan, W. B., Carbotte, S. M., Coplan, J. O., O'Hara, S., Melkonian, A., Arko, R., et al. (2009). Global multi-resolution topography synthesis
[Dataset]. Geochemistry, Geophysics, Geosystems, 10(3), 1–9. https://doi.org/10.1029/2008GC002332
Sánchez, J., Palacios, O., Feininger, T., Carlotto, V., & Quispesivana, L. (2006). Puesta en evidencia de granitoides Triásicos en los
Amotapes-Tahuín: Deflexión de Huancabamba. In XIII Congreso Peruano de Geología, Resúmenes Extendidos: Lima, Perú (pp. 312–315).
Sociedad Geológica del Perú.
Scholl, D. W. (1987). Plate tectonics the predictor: The history of wonderments about subduction erosion and sediment subduction – A search for
the missing. In T. W. C. Hilde & R. H. Carlson (Eds.), Silver anniversary celebration of plate tectonics: Geodynamics symposium (pp. 54–56).
Texas A&M University, Geodynamics Research Institute.
Scholl, D. W., Kirby, S. H., von Huene, R., Ryan, H., Wells, R. E., & Geist, E. L. (2015). Great (≥Mw 8.0) megathrust earthquakes and the subduc-
tion of excess sediment and bathymetrically smooth seafloor. Geosphere, 11(2), 236–265. https://doi.org/10.1130/GES01079.1
Scholl, D. W., & von Huene, R. (2007). Crustal recycling at modern subduction zones applied to the past-issues of growth and preservation
of continental basement crust mantle geochemistry, and supercontinent reconstruction. In R. D. Hatcher, M. P. Carlson, J. H. McBride, &
J. R. Martínez-Catalán (Eds.), 4-D framework of continental crust, The Geological Society of America Memoir 200 (pp. 9–32). https://doi.
org/10.1130/2007.1200(02)
Scholl, D. W., & von Huene, R. (2010). Subduction zone recycling processes and the rock record of crustal suture zones. Canadian Journal of
Earth Sciences, 47(5), 633–654. https://doi.org/10.1139/e09-061
Scholl, D. W., von Huene, R., Vallier, T. L., & Howell, D. G. (1980). Sedimentary masses and concepts about tectonic processes at underthrust
ocean margins. Geology, 8(12), 564–568. https://doi.org/10.1130/0091-7613(1980)8<564:SMACAT>2.0.CO;2
Seely, D. R., Vail, P. R., & Walton, G. G. (1974). Trench slope model. In C. A. Burk & C. L. Drake (Eds.), The geology of continental margins
(pp. 249–260). Springer Science+Business Media. https://doi.org/10.1007/978-3-662-01141-6_18
Séranne, M. (1987). Evolution Tectono-Sedimentaire du Basin de Talara (nord-ouest du Perou). Bulletin de l'Institut Français d'Études Andines,
XVI(3–4), 103–125. Retrieved from https://www.persee.fr/doc/bifea_0303-7495_1987_num_16_3_952
Shreve, R. L., & Cloos, M. (1986). Dynamics of sediment subduction, melange formation, and prism accretion. Journal of Geophysical Research,
91(B10), 10229–10245. https://doi.org/10.1029/JB091iB10p10229
Simmons, M. D., Miller, K. G., Ray, D. C., Davies, A., van Buchem, F. S. P., & Gréselle, B. (2020). Chapter 13. Phanerozoic eustasy. In F.
M. Gradstein, J. G. Ogg, M. D. Schmitz, & G. B. Ogg (Eds.), Geologic time scale 2020 (pp. 357–400). Elsevier. https://doi.org/10.1016/
B978-0-12-824360-2.00013-9
Sosson, M., Bourgois, J., & de Lépinay, B. M. (1994). SeaBeam and deep-sea submersible Nautile surveys in the Chiclayo canyon off Peru (7 S):
Subsidence and subduction-erosion of an Andean-type convergent margin since Pliocene times. Marine Geology, 118(3–4), 237–256. https://
doi.org/10.1016/0025-3227(94)90086-8
Stainforth, R. M. (1948). Applied micropaleontology in coastal Ecuador. Journal of Paleontology, 22(2), 113–151. Retrieved from http://www.
jstor.org/stable/1299388
Stainforth, R. M. (1954). A revised summary of the stratigraphic and paleontology of northwest Peru. Internal report 121-21A. International
Petroleum Company.

LAJO-YÁÑEZ ET AL. 26 of 27
19449194, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023TC007860 by Cochrane Peru, Wiley Online Library on [26/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Tectonics 10.1029/2023TC007860

Stern, C. R. (2011). Subduction erosion: Rates, mechanisms, and its role in arc magmatism and the evolution of the continental crust and mantle.
Gondwana Research, 20(2–3), 284–308. https://doi.org/10.1016/j.gr.2011.03.006
Tamay, J., Galindo-Zaldivar, J., Soto, J., & Gil, A. J. (2021). GNSS constraints to active tectonic deformations of the South American Continental
margin in Ecuador. Sensors, 21(12), 1–17. https://doi.org/10.3390/s21124003
Thalmann, H. E. (1946). Micropaleontology of Upper Cretaceous and Paleocene in western Ecuador. American Association of Petroleum Geolo-
gists Bulletin, 30(3), 337–347. https://doi.org/10.1306/3D933802-16B1-11D7-8645000102C1865D
Thornburg, T. M., & Kulm, L. D. (1981). Sedimentary basins of the Peru continental margin: Structure, stratigraphy, and Cenozoic tectonics from
6°S to 16°S latitude. In L. D. Kulm, J. Dymond, E. J. Dasch, & D. M. Hussong (Eds.), Nazca plate: Crustal formation and Andean conver-
gence, The Geological Society of America Memoir 154 (pp. 393–422). https://doi.org/10.1130/MEM154-p393
Travis, R. B. (1953). La Brea-Parinas oil field, Northwestern Peru. American Association of Petroleum Geologists Bulletin, 37(9), 2093–2118.
https://doi.org/10.1306/5CEADD73-16BB-11D7-8645000102C1865D
Travis, R. B., Gonzales, G., & Pardo, A. (1976). Hydrocarbon potential of coastal basins of Peru. American Association of Petroleum Geologists
Memoir, 25, 331–338. https://doi.org/10.1306/83D918CB-16C7-11D7-8645000102C1865D
Vail, P. R., Hardenbol, J., & Todd, R. G. (1984). Jurassic unconformities, chronostratigraphy, and sea-level changes from seismic stratigraphy
and biostratigraphy. In J. S. Schlee (Ed.), Interregional unconformities and hydrocarbon accumulation, American Association of Petroleum
Geologists Memoir 36 (pp. 129–144). https://doi.org/10.5724/gcs.84.03.0347
Vannucchi, P., Sage, F., Phipps Morgan, J., Remitti, F., & Collot, J. Y. (2012). Toward a dynamic concept of the subduction channel at erosive
convergent margins with implications for interplate material transfer. Geochemistry, Geophysics, Geosystems, 13(1), 1–24. https://doi.
org/10.1029/2011GC003846
Vaughan, T. W. (1937). The Tertiary larger foraminifera of southwest Ecuador. In G. Sheppard (Ed.), The geology of south-western Ecuador
(pp. 150–175). Thomas Murby and Co.
Vegas, F. (1970). Ilustraciones de la Microfauna de la Formación Lutita Talara. Internal report 300-19. Petróleos del Perú.
Villegas-Lanza, J. C., Chlieh, M., Cavalié, O., Tavera, H., Baby, P., Chire-Chira, J., & Nocquet, J. M. (2016a). Active tectonics of Peru: Hetero-
geneous interseismic coupling along the Nazca megathrust, rigid motion of the Peruvian Sliver, and Subandean shortening accommodation.
Journal of Geophysical Research: Solid Earth, 121(10), 7371–7394. https://doi.org/10.1002/2016JB013080
Villegas-Lanza, J. C., Nocquet, J. M., Rolandone, F., Vallée, M., Tavera, H., Bondoux, F., et al. (2016b). A mixed seismic–aseismic stress release
episode in the Andean subduction zone. Nature Geoscience, 9(2), 150–154. https://doi.org/10.1038/ngeo2620
von Huene, R., Bourgois, J., Miller, J., & Pautot, G. (1989). A large tsunamogenic landslide and debris flow along the Peru trench. Journal of
Geophysical Research, 94(B2), 1703–1714. https://doi.org/10.1029/JB094iB02p01703
von Huene, R., & Lallemand, S. (1990). Tectonic erosion along the Japan and Peru convergent margins. Geological Society of America Bulletin,
102(6), 704–720. https://doi.org/10.1130/0016-7606(1990)102<0704:TEATJA>2.3.CO;2
von Huene, R., & Miller, J. (1988). Migrated multichannel seismic-reflection records across the Peru continental margin. Proceedings of the
Ocean Drilling Program, 109–124. Initial report 112. https://doi.org/10.2973/odp.proc.ir.112.107.1988
von Huene, R., Ranero, C. R., & Vannucchi, P. (2004). Generic model of subduction erosion. Geology, 32(10), 913–916. https://doi.org/10.1130/
G20563.1
von Huene, R., & Scholl, D. W. (1991). Observations at convergent margins concerning sediment subduction, subduction erosion, and the growth
of continental crust. Reviews of Geophysics, 29(3), 279–316. https://doi.org/10.1029/91RG00969
Weiss, L. (1955). Foraminifera from the Paleocene Pale Greda Formation of Peru. Journal of Paleontology, 29(1), 1–21. Retrieved from http://
www.jstor.org/stable/1300125
Winter, L. S., Tosdal, R. M., Mortensen, J. K., & Franklin, J. M. (2010). Volcanic stratigraphy and geochronology of the Cretaceous Lancones
Basin, Northwestern Peru: Position and timing of giant VMS deposits. Economic Geology, 105(4), 713–742. https://doi.org/10.2113/
gsecongeo.105.4.713

Erratum
The originally published version of the article contained an error in Figure 14. The figure contained partial gray-
scale on the right-hand side. The figure has been replaced to show the proper colors. This may be considered the
authoritative version of record.

LAJO-YÁÑEZ ET AL. 27 of 27

You might also like