Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Journal of South American Earth Sciences 127 (2023) 104376

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


journal homepage: www.elsevier.com/locate/jsames

The effects of positive tectonic inversion structures on the formation of


thrust systems on the western Domeyko Cordillera, northern Chile:
Implications for the orogenic development of the outer Andean forearc
Cristopher López a, *, Renato Cisternas a, Sebastián Espinoza a, Rodrigo González a, c,
Fernando Martínez b, Rodrigo Riquelme a, Daniela Montenegro a, Jorge Morales a
a
Departamento de Ciencias Geológicas, Universidad Católica del Norte, Avenida Angamos, 0610 Antofagasta, Chile
b
Carrera de Geología, Facultad de Ingeniería, Universidad Andres Bello, Campus República, República 220, Santiago, Chile
c
Millennium Institute on Volcanic Risk Research - Ckelar Volcanoes, Antofagasta, Chile

A R T I C L E I N F O A B S T R A C T

Keywords: This work aims to propose a novel structural setting for the outer Andean forearc. Specifically, we propose a new
Domeyko cordillera deformation mechanism related to thrusting belt systems induced by positive tectonic inversion processes in the
Thrust systems western flank of the Domeyko Cordillera in northern Chile. Thrust belt systems have been rarely understood for
Positive tectonic inversion structures
the outer Andean forearc; in fact, the mention of thrusting is scarce and structural interpretations are restricted
Tectonosequences
only to strike-slip faulting and positive tectonic inversion processes. In this study, we have constructed novel
Outer andean forearc
2-D seismic reflection line balanced cross-sections from new structural fieldwork and combined them with 2-D reflection seismic line re-
interpretations, thus offering new ideas about the evolution and development of the forearc of the Central
Andes. Based on this, we propose integrated kinematic models that describe the relationships between inherited
extensional structures, positive tectonic inversion structures, and the formation of thrust systems. The proposed
balanced cross-sections have been restored to a pre-shortening condition. In this way, two structural styles are
proposed for the western flank of the Domeyko Cordillera. The first corresponds to west-directed positive tec­
tonic inversion structures, which expose and extrude the Middle Jurassic syn-rift deposit of the ancient Mesozoic
rift basin. The second consists of east-directed thin-skinned thrust systems, mainly affecting the Upper Cretaceous
to Eocene volcano-sedimentary synorogenic sequences. The interaction between inversion structures and thrust
system structural styles would generate a west-directed intracutaneous wedge. Finally, the minimum shortening
obtained from our balanced cross-section restorations is ca. 11% for the Río Loa and the San Salvador Canyons
segment and ca. 56% for the Quebrada Mala segment. The formation of thrusting structures is controlled by a
major detachment fault activated from a regional-scale angular unconformity, which we have named the Que­
brada Mala detachment fault. An important part of the shortening was taken up and accommodated by but­
tressing structures and their associated intraformational detachment-thrust levels.

1. Introduction pure thrusting may also induce further orogenic growth through a rapid
shortening-and-uplift, accumulating extensive synorogenic sequences
Orogenic chains induced by positive tectonic inversion show that (Amilibia et al., 2008; Bonini et al., 2012; Martínez et al., 2020b; López
basement-involved structures are a first-order deformation mechanism et al., 2022). Nevertheless, shortening-related deformational styles,
generating crustal shortening and thickening (e.g., Williams et al., 1989; focused on synorogenic deposits, often obliterate much of the structures
McClay and Buchanan, 1992; Hayward and Graham, 1996; Bailey et al., that build orogens, thus hiding a large amount of tectonic shortening. In
2002; among others). Different studies have shown that inverted normal this context, many structural styles (e.g., passive roof duplex, intracu­
faults are able to uplift the infill of extensional rift-basin and taneous wedge, antiformal stack duplex, among others) also play a
deep-crustal blocks (Bonini et al., 2012; Héja et al., 2022). Similarly, fundamental role in the orogenic development of many mountain ranges

* Corresponding author.
E-mail address: cristopher.lopez@ucn.cl (C. López).

https://doi.org/10.1016/j.jsames.2023.104376
Received 28 February 2023; Received in revised form 24 April 2023; Accepted 25 April 2023
Available online 9 May 2023
0895-9811/© 2023 Elsevier Ltd. All rights reserved.
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

worldwide (Bonini, 2001; Couzens-Schultz et al., 2003; Brocher et al., reactivation of inverted normal faults. Recent studies have documented
2004; Scisciani et al., 2019). The interplay that a basement-involved the presence of pure thrust faulting mainly on the eastern slope of the
fault may have with a synorogenic sequence-involved thrust fault may Domeyko Cordillera or its transitional limit with the Preandean basins
generate gaps in structural interpretation on the evolution of an (e.g., the Salar de Atacama, the Punta Negra, and the Pedernales basins).
orogenic system. Specifically, the relationship between a partially However, the areal extension and thickness of the late Cenozoic gravels
inverted normal fault, a buttressing structure (Butler, 1989; Scisciani, make it difficult to formulate reliable structural interpretations on the
2009; Calamita et al., 2018), and/or thrust systems make it difficult western flank of the Domeyko Cordillera. The main structural evidence
when proposing new structural models to estimate the amount of on this side of the Domeyko Cordillera is mainly associated with
shortening (Fig. 1). Buttressing structures by themselves do not strike-slip faulting. In this study, we show evidence to indicate that the
accommodate or generate crustal shortening. In turn, normal inverted deformation concentrated on the western flank of the Domeyko
faults do not often generate total inversion along their slip fault-planes. Cordillera is related to a thin-skinned thrust system induced by large
However, the interaction that these two tectonic inversion-related basement-involved, partially reverse-reactivated normal faults. These
structures may have with the formation of newly-formed thrust faults faults would transfer deformation towards the Upper Cretaceous and
(pure thrust faulting or thin-skinned tectonics) may generate complex early Cenozoic sedimentary and volcanic synorogenic sequence through
structural styles mainly in a synorogenic sequence (Boyer and Elliott, thin-skinned thrust systems.
1982; Bonini, 2001; Calamita et al., 2001; Couzens-Schultz et al., 2003;
Mount et al., 2011; von Hagke and Malz, 2017; Tavarnelli et al., 2021; 2. Regional geological framework
Pace et al., 2022b).
The Central Andes is the second largest orogenic chain worldwide, The outer forearc may be considered as the region that extends from
and a typical example of a subduction-related orogen where continuous the Coastal Cordillera to the transitional limit between the Central
subduction of oceanic plates (e.g., the Farallon and Nazca plates) occurs Depression and the Domeyko Cordillera (Fig. 2). This region is charac­
beneath the South American continental margin (e.g., Dewey and Bird, terized by a thick extensive cover of late Cenozoic gravels that hide a
1970; Isacks, 1988; Allmendinger et al., 1997). Its eastern flank is a large large part of stratigraphic record and the distribution of deep structures.
morpho-tectonic province characterized by a hybrid fold-and-thrust belt The regional-scale structural systems have only been defined along the
system (Baby et al., 1992; Rojas Vera et al., 2019). While, the western eastern flank of the Coastal Cordillera and on the western flank of the
flank of the Central Andes is composed by north-south oriented Domeyko Cordillera. In the Coastal Cordillera, the first-order structures
morpho-tectonic features, which in turn, compose the Andean forearc of are part of the Atacama Fault System (Arabasz, 1971; Scheuber and
northern Chile. From west to east, these are: the Coastal Cordillera, the González, 1999), whereas in the Domeyko Cordillera the first-order
Central Depression, the Domeyko Cordillera, and the Preandean structures are part of the Domeyko Fault System (Maksaev and Zen­
Depression. The highest morpho-tectonic expression corresponds to the tilli, 1999; Mpodozis et al., 1993). The kinematic activity of the Atacama
Domeyko Cordillera, which has been interpreted as a tectonic block Fault System has been associated with a listric extensional system
controlled by a positive tectonic inversion. The most credited models developed during the early Jurassic (Grocott et al., 1994; Taylor et al.,
propose that the shortening has been induced by the partial or total 1998). Then, during the late Jurassic-early Cretaceous time interval, its

Fig. 1. Model of inversion structures and their relationships with thin-skinned thrust systems. The model illustrates positive tectonic inversion structures and how
they may influence the formation of thrust systems (Model integrated from Boyer and Elliott, 1982; Williams et al., 1989; McClay and Buchanan, 1992; Hayward and
Graham, 1996; Bailey et al., 2002).

2
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 2. DEM image of the Andean forearc of the Central Andes. The figure shows the distribution of the major morpho-tectonic units recognized in northern Chile.
The black box corresponds to the study area in the western flank of the Domeyko Cordillera.

kinematics would have changed to sinistral-slip-strike (Pardo-Casas and Salvador Canyons and to the east of the Quebrada Mala and the Que­
Molnar, 1987; Hervé, 1987; Scheuber and Andriessen, 1990; Jaillard brada Perdida creeks (Fig. 3). Complementary, along this stratigraphic
et al., 1990). On the other hand, the Domeyko Fault System has mainly contact, a series of minor Cretaceous and Paleocene intrusives bodies are
been defined as a basement-involved strike-slip set (Mpodozis et al., distributed and emplaced throughout both units (Marinovic et al., 1996;
1993; Amilibia et al., 2008; Mpodozis and Cornejo, 2012) active from Basso, 2004; Duhart et al., 2018).
the middle Eocene-early Oligocene time interval onwards (Maksaev and The Jurassic successions correspond to the syn-rift fill of ancient rift
Zentilli, 1999; Nalpas et al., 2005; Arriagada et al., 2008). basin systems associated with the Tarapacá back-arc extensional basin
Specifically on the western flank of the Domeyko Cordillera, one of (Vicente, 2006), which was created from the western Gondwana
the most notorious geological features corresponds to a NNE-oriented break-up (Uyeda and Kanamori, 1979; Mpodozis and Ramos, 1990,
lithological contact between two of the regional stratigraphic units. 2008; Aguirre-Urreta, 1993; Ardill, 1996; Ardill et al., 1998; Ramos,
This limit is exposed for more than 160 km between ca. 22◦ 20′ to 23◦ 50’ 2010; del Rey et al., 2016). These units consist of nearly 3000 m of
Lat. S., and may also be considered as the transitional morpho-tectonic carbonate and siliciclastic marine sequences (Jensen et al., 1976;
contact between the Central Depression and the Domeyko Cordillera Gröschke et al., 1988; Mpodozis and Ramos, 1990; Prinz et al., 1994;
(Figs. 2 and 3). To the west, the first unit is represented by the Upper Marinovic and García, 1999). Additionally, the basal successions mainly
Cretaceous volcanic and clastic deposits belonging to the Cerro Cortina, consist of Sinemurian to Kimmeridgian fossiliferous limestones defined
the Quebrada San Cristobal, the Paradero del Desierto beds and the as the Caracoles Group (García, 1967; Ramírez and Gardeweg, 1982;
Quebrada Mala Fm. (Montaño, 1976; Marinovic and García, 1999; Marinovic and Lahsen, 1984; Marinovic and García, 1999). This suc­
Medina et al., 2012), which overlies through a marked angular uncon­ cession is unconformably covered by Upper Jurassic to Lower Creta­
formity the Jurassic-Lower Cretaceous marine and continental deposits ceous transitional marine-continental siliciclastic sediments assigned to
composed by the Caracoles Gr., the Cerro Campamento, the Llanura the Cerritos Bayos Formation which is composed of limestones, mud­
Colorada, and the Cerritos Bayos Fms. (García, 1967; Montaño, 1976; stones, sandstones, and shales (Biese, 1961; García, 1967; Baeza, 1979;
Marinovic and García, 1999; Duhart et al., 2018). These rocks are Montaño, 1976). This last stratified unit may be correlated with the
recognized along the westernmost part of the Domeyko Cordillera, and Llanura Colorada Formation (Muñoz, 1989). Frequently, these Jurassic
their most prominent outcrops are exposed in the Río Loa and the San successions show important lateral thickness variations, and they are

3
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 3. Generalized regional-geological map of the western flank of the Domeyko Cordillera. The red box corresponds to the Río Loa and San Salvador Canyons
segment. The yellow line represent location of the Z1020 seismic line, and the trace of the balanced cross-section drawn along this segment. The blue box corresponds
to Quebrada Mala segment. The green line represents the location of the balanced cross-section.

wedge shaped like those typical of syn rift-deposits with increasing dip These rocks include more than 2000 m of dacitic and rhyolitic tuff,
angles in the older strata at depth (Withjack et al., 2002). basaltic, andesitic, and dacitic lavas, sandstone, and conglomerate,
The Upper Cretaceous successions consist of continental volcanic and mostly exposed between the Central Depression and the Domeyko
sedimentary rocks deposited under conditions of intensive volcanism Cordillera and unconformably overlying the Jurassic syn-rift succes­
and continental sedimentation within synorogenic foreland basins sions. Furthermore, deposition of the Upper Cretaceous sediments has
(Cornejo et al., 2003; Amilibia, 2002, 2008; Mpodozis et al., 2005). been interpreted as resulting from a drastic change in tectonic

4
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

environment during the late Cretaceous (Scheuber et al., 1994), from an bedding orientation measurements, fault kinematic indicators and fold
extensional regime (Uyeda and Kanamori, 1979; Aguirre-Urreta, 1993; attitudes recognized along the western flank of the Domeyko Cordillera,
Ramos, 2010; del Rey et al., 2016, 2019; Coloma et al., 2017; Martínez and then projected in the geological cross-sections. Also, to validate our
et al., 2020a) to compressive regime (Cobbold et al., 2007; Amilibia balanced cross-sections and to determine the initial position of
et al., 2008; Ramos, 2010) in the western margin of South America. contractional structures, as well as the syn-rift -and -post-rift deposits
Along the Central Andes, these tectonic variations and geodynamic involved within compressional structures, we have restored a 2-D
changes are associated with the Peruvian tectonic phase (Steinmann, balanced cross-section to its pre-thrusting template. This restoration
1929; Mpodozis and Ramos, 1990; Mpodozis et al., 2005; Arriagada consisted in establishing the position of syn-rift and post-rift strati­
et al., 2008; Bascuñán et al., 2015). This period was accompanied by graphic units before the shortening assuming the Upper Cretaceous as
faulting and folding of the Jurassic-Lower Cretaceous marine and tran­ the onset of contraction. The amount of the shortening was calculated
sitional deposits, erosion of the continental border and using the line-length conservation method (Dahlstrom, 1969; Wood­
eastward-migration of the magmatic arc to its current position in the ward et al., 1989). This allowed us to validate the geometry and slip
Central Depression and the Domeyko Cordillera (Mpodozis et al., 2005; involved in the main faults and folds recognized on the surface and
Bascuñán et al., 2015). interpreted in the subsurface.
The Upper Cretaceous stratigraphic units are followed by a section of
approximately 500 m of lavas, tuffs, and conglomerates assigned to the 4. Results
Paleocene Cinchado Fm. (Montaño, 1976), and the Eocene Icanche Fm.
(Duhart et al., 2018). The Paleocene rocks yield K–Ar ages ranging be­ 4.1. Surface and subsurface structures
tween ca. 61-52 Ma (Marinovic and García, 1999), while the Eocene
rocks yield U–Pb ages ranging between ca. 53-50 Ma (Duhart et al., 4.1.1. The San Salvador and Río Loa Canyons segments
2018). The first-order structures recognized at the Loa River and the San
Finally, the late Cenozoic stratigraphic units consist of nearly ca. 300 Salvador Canyons segment (red box in Fig. 3) consist of preserved
m of continental gravel, sandstone, and lacustrine limestone that extend extension-related depositional geometries, inherited normal faults, and
almost completely over the Central Depression and also infill inter­ positive tectonic inversion structures (Fig. 4). The sequences form part
montane basins (e.g., Calama Basin), named the Atacama Gravels of a regional-scale, west-dipping Jurassic-Lower Cretaceous monoclinal
(Mortimer, 1973), the Loa Group (May, 1997; May et al., 2005; Blanco, (Fig. 4a), that even towards the east, extend until the Permian-Triassic
2008), and the Batea Fm. (Naranjo and Paskoff, 1985). These conti­ deposits (López et al., 2022), recording a continuous accumulation
nental deposits are derived from an important erosion process that associated with the rift-basin fill. From east to west, this sequence is
occurred following the Eocene-Oligocene uplift of the Domeyko composed of the Upper-Middle Jurassic calcareous-marine sequences
Cordillera (Mpodozis et al., 2005; Charrier et al., 2009; Sanchez et al., (Cerro Campamento and Moctezuma Fms., respectively), the Upper
2017; Riquelme et al., 2017). Jurassic-Lower Cretaceous marine-continental transitional deposits
(Cerritos Bayos Fm.), and the Lower Cretaceous continental deposits
3. Methodology (San Salvador Fm.) (Figs. 4 and 6). The positive tectonic inversion
structures affect the Lower-Middle Jurassic deposits through
3.1. Surface and subsurface structures interpretations reverse-reactivation of inherited normal faults (Fig. 4). Specifically,
inversion structures are composed of inversion anticline-and-syncline,
Two geological cross-sections located in two specific areas along the and mainly east-verging overturned folds associated with partially
western flank of the Domeyko Cordillera have been constructed. The inverted syn-rift stratigraphic wedges, which usually show important
fieldwork structural data were collected in sectors that stand out for the thickness-variations (Fig. 4b). Other minor inversion structures consist
exposure of good outcrops of stratigraphic units and structures. These data of west-directed intraformational detachment-thrust faults decoupled
also consider the recognition of depositional geometries associated with along evaporitic levels, putting in contact the different Jurassic units
syn-rift, post-rift and synorogenic deposits (tectonosequences). The two that filled the Mesozoic rift basin (Fig. 4c). A prominent structure
specific areas correspond to the San Salvador and the Río Loa Canyons and recognized along the Río Loa Canyon corresponds to the Chintoraste
the Quebrada Mala creek segments (Fig. 2). The structural data consists of fault (Fig. 4d) (Duhart et al., 2018). The Chintoraste fault is a
geological-structural maps compiled by satellite image, fieldwork data, NE-oriented and subvertical-dipping (80–85◦ W) inverted normal fault
acquisition of drone photographs (DJI Mini 2 and DJI Air 2 S) which were that separates the Middle Jurassic marine deposits to the west (hang­
integrated with stratigraphic and structural data obtained from the ing-wall; the Cerro Campamento Fm.), from the Lower Cretaceous
regional chartography (Marinovic and García, 1999; Tomlinson et al., continental deposits to the east (footwall; San Salvador Fm.). A noto­
2018; Duhart et al., 2018). Additionally, stratigraphical contacts, rious characteristic of this fault corresponds to its prominent damage
strike-and-dip measurements, stratigraphic-thickness, the positions of zone; hydrothermal-alteration related mineralization, disharmonic
faults and folds represent the main structural dataset presented in this folding, and minor faulting are well exposed recognized features along
work. This information was integrated and combined with a the fault zone (Fig. 4d). Thus, this fault may separate two distinctive
time-migrated 2-D seismic profile re-interpretation (obtained and modi­ structural styles. In its hanging-wall block, the beds show chaotic and
fied from López et al., 2019), across the San Salvador and the Río Loa disharmonic anticline- and syncline-folds related to a ‘buttressing-type
Canyons, showing a continuity between various regional-scale structures inversion assemblage’. Associated with this same structure, a series of
exposed on the surface and subsurface structures hidden under the west­ fault-propagation folds and recumbent folds may be recognized on its
ern flank of the Domeyko Cordillera. hanging-wall (Fig. 4d and e). In the footwall, a prominent thrust system
involves the Lower Jurassic-Upper Cretaceous sequences and the Eocene
3.2. Balanced cross-section and 2-D pre-shortening restoration deposits (Fig. 5). These structures consist of an east-directed thin-­
skinned thrust system, which is hidden below the Mio-Pliocene se­
Two EW-trending balanced cross-sections were constructed inte­ quences. There are three main thrust structures recognized in this
grating a 2-D seismic profile re-interpretation and structural fieldwork segment: (1) a detachment fold, (2) an active-roof duplex, and (3) an
data on the analyzed cross-sections of the San Salvador and the Río Loa east-directed imbricate thrust system (Fig. 5). The (1) detachment fold
Canyons and the Quebrada Mala areas. These cross-sections are (Fig. 5a) is a prominent structure, which is well exposed in the Loa River
composed by robust structural information consistent with the main Canyon. Here, it is possible to observe evaporite basal-levels of the
exposed structures and its associated-stratigraphic deposit, as well as Lower Cretaceous sequence, forming part of an east-directed

5
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 4. Inversion structures exposed in the San Salvador – Río Loa Canyons segments. (a) West-dipping Lower Jurassic monoclinal associated with the rift-basin fill.
(b) East-verging overturned fold showing a wedge-shape depositional geometry. (c) Folds associated with partially inverted syn-rift stratigraphic wedges and
intraformational detachment-thrust faults. (d) Buttressing structure associated with a partially inverted normal fault. (e) Fault-propagation fold.

detachment fault, generating an asymmetric anticline of ca. 40 thrust system and the Eocene synorogenic strata show that the defor­
m-amplitude. In turn, this structure generates a west-directed passive mation triggered during the first stage of positive tectonic inversion by
roof thrust affecting the Eocene sequences through its basal-bedding reverse-reactivation of the Chintoraste fault continued during the late
planes. Both structures may be described as an east-directed intracuta­ Cretaceous and early Cenozoic.
neous wedge or passive-roof duplex. The (2) active-roof duplex (Fig. 5b) To describe the structures hidden below the western Domeyko
corresponds to a west-directed antiformal stack duplex system (sensu Cordillera, we have reinterpreted a segment of the Z1020 seismic profile
Boyer and Elliott, 1982) affecting the Upper Jurassic – Lower Cretaceous (Fig. 7) previously analyzed in López et al. (2019). This new analysis is
deposits and the Eocene sequences. The (3) trailing imbricate fan thrust based on novel fieldwork data that integrates the recognition of new
systems (sensu Boyer and Elliott, 1982) (Fig. 5c) is immediately recog­ structural styles and new published U–Pb ages for the units exposed
nized to the east of the Chintoraste fault. In this way, these three along the Río Loa and the San Salvador Canyons (Duhart et al., 2018).
structural styles define a thin-skinned thrusts system, affecting the The velocities of the different tectonosequences used to depth-covert the
Upper Jurassic - Lower Cretaceous and Eocene sequences. Finally, a TWT seismic profile are presented by López et al. (2019). The new
series of synorogenic deposits with a growth strata geometry are analyzed Z1020 seismic line reaches a length of 31.5 km and is located to
observed in the Eocene sequences. These sequences form a wedge sha­ the westernmost part of the Calama Basin, between the previously
ped, thin towards the west, and are linked to the uplift of the mentioned canyons (Fig. 3). The description of the tectonosequences
hanging-wall block of the Chintoraste fault. On the surface, its trace is identified in this seismic profile is supported by their good outcrops
covered by the Mio-Pliocene deposits (Fig. 5e). Both the east-directed (Fig. 6), which allows us to extrapolate these to recognized

6
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 5. Thrusting structures exposed in the Río Loa Canyon. (a) Detachment-thrust fault and passive roof thrust fault. (b) Intracutaneous wedge system. (c) Details of
Intracutaneous wedge system showing west-directed antiformal stack duplex. (d) Trailing imbricate fan thrust systems (e) Growth strata geometry affecting the
Eocene sequence. Note that the recognized tectono-sequences in the Río Loa Canyon have been subdivided for a better visualization of how they are affected.

tectonosequences to depth. The first tectonosequence that may be different truncated reflectors is associated with the ‘Chintoraste fault’
interpreted corresponds to the Middle Jurassic syn-rift deposits (Cerro previously recognized and described. To the eastern most part,
Campamento Fm.). The most prominent structural feature of this tec­ diffuse-and-semicontinuous reflectors may be recognized above the
tonosequence consists of anticline- and syncline-folded reflectors, which Middle Jurassic tectonosequences. Good outcrops on the surface allow
are associate with a harpoon structure-type or inversion for an association of these reflectors with the Eocene and Mio-Pliocene
anticline-and-syncline. To the west, these reflectors are truncated by a deposits.
‘partially inverted east-dipping normal fault’. From the lower segments
of this inverted fault, a short-cut fault affects continuous, parallel, and 4.1.2. The Quebrada Mala segment
medium-amplitude reflectors identified between ca. 0.0 and 1.0 s To show the main highlighted structural styles exposed along the
(two-way travel-time), which are associated with the Lower Cretaceous Quebrada Mala segment (blue box in Fig. 3), we have constructed the
post-rift deposits (San Salvador Fm.). To the east of the seismic profile, EW-trending AA’ geological cross-section (Fig. 8), that extends for 13.4
the Middle Jurassic reflectors appear chaotic and very diffused; how­ km. This profile includes two structural domains. The eastern domain
ever, the correlation of the units exposed on the surface allows us to state consists of prevalently ‘west-verging positive tectonic inversion struc­
that, between 0.0 and 1.0 s (two-way travel-time), all these reflectors tures’ affecting the Middle and Upper Jurassic marine-calcareous (Car­
correspond to the Middle Jurassic marine deposits. Highly deformed acoles Gr.) and the Upper Jurassic-Lower Cretaceous transitional
subvertical-dipping sequences exposed on surface allow us to recognize deposits (Llanura Colorada Fm). The western domain consists of a thin-
a series of structures generated by intraformational detachments faults. skinned east-directed thrust system deforming the Upper Cretaceous
Finally, to the east, a structure that is recognized by putting in contact volcanic and sedimentary rocks (Quebrada Mala Fm.). Along this cross-

7
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 6. Tectonosequences. Pictures of outcrops showing contact relationship between mainly tectonosequences recognized in the Z1020 seismic line and exposed
along the San Salvador and Río Loa Canyons. The location of each picture is indicated in Fig. 7.

section, a marked regional-scale NNE-oriented angular unconformity structures, by-pass faults, buttressing structures (Fig. 9d), and back-
represents the most prominent stratigraphic feature, reaching a length of thrust faults may be recognized as part of the hanging wall of the
ca. 18 km, being widely recognized on the western flank of the Domeyko inverted normal fault. The folds show a NNE-trending axis, by an east-
Cordillera. This stratigraphic limit separates the underlying Middle- dipping axial surface, and by a close-interlimb angle (<10◦ ). The (3)
Upper Jurassic sequences from the overlying Upper Cretaceous se­ detachment faults are strongly controlled by the evaporitic level within
quences (Figs. 3 and 8), and apparently, controlled the emplacement and the syn-rift deposits. Generally, these levels correspond to gypsum beds,
distribution of some Upper Cretaceous-Paleocene hypabyssal intrusive which also, were able to generate a diapiric structure (Fig. 9b and c).
rocks. The ‘thin-skinned east-directed thrust system domain’ only affects
The ‘west-verging positive tectonic inversion structures domain’ is the Upper Cretaceous volcanic and sedimentary sequences (the Que­
represented by (1) short-cut faults, (2) reverse-reactivated normal faults, brada Mala Fm.) (Fig. 8). Its deformational style is strongly dependent
(3) detachment faults, and (4) some minor inversion structures (Fig. 9). on the angular unconformity, which were reused as an “east-directed
The (1) short-cut faults correspond to NNE-trending, moderately dip­ detachment-thrust fault” (Fig. 9). We consider pertinent to define this
ping, and west-directed thrust faults, that propagated near-parallel to structure as ‘the Quebrada Mala detachment fault’, since there are
the bedding-planes, affecting only the Middle Jurassic sequences various rigorous data and arguments based on fieldwork observations to
(Fig. 9a). The (2) reverse-reactivated normal faults strongly controlled support an important shortening-related shear zone which is localized
the depositional geometries of the Middle Jurassic sediments. These along the angular unconformity. In this way, four structural character­
sequences show thickness-variations outlining syn-depositional, wedge- istics associated with the Quebrada Mala detachment fault are described
shaped accumulations, thus allowing association with syn-rift deposits (Fig. 10). First, the stratigraphic-discontinuity, as an angular uncon­
(Fig. 8b and 9a). Also, as for (4) minor tectonic inversion-related formity, presents a marked angular break associated with a flat-ramp-

8
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 7. Z1020 seismic line. (a) Uninterpreted grey-scale seismic line. (b) Interpreted grey-scale seismic line showing the tectonosequences and the main structures.
See Fig. 3 for location.

flat geometry. Nevertheless, the possible occurrence of inverted normal surface parallel to the angular unconformity. Fourth, an intraforma­
faults, under the angular unconformity, could generate this kind of tional detachment-thrust fault, generated from the overlying Upper
geometric variation (or could lead to fault-propagation folding) (Figs. 8, Cretaceous volcanic and siliciclastic sequences and located along the
10a and 10d). Second, foliated and breached basal conglomerates level, angular unconformity, is recognized (Fig. 10d and e). On the other hand,
over the angular unconformity, may be recognized. These breached the structures that form part of the ‘thin-skinned east-directed thrust
rocks standout for having minor shear zone focused on the sandy matrix system domain’ (Fig. 11) result from the superposition of two opposite-
of the basal conglomerate, since both the matrix and the clasts are ori­ directed antiformal stack duplexes (sensu Boyer and Elliott, 1982)
ented parallel to the detachment plane (Fig. 10b). Third, S–C fabric (Fig. 8). Firstly, immediately to the west of the Quebrada Mala
characterizes a shear zone with an E-W-trending shortening axis and C detachment fault, the Upper Cretaceous sequences consist of an

9
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 8. (a) Generalized structural map and (b) AA’ cross-section of the Quebrada Mala segment.

east-dipping beds. This imbricate system is associate with west-directed and the extrusion of the Mesozoic syn-rift basin along the western flank
antiformal stack duplexes, which work as a propagated thrusting system of the Domeyko Cordillera. In the westernmost part of this balanced
from older basement rocks to younger cover sequences. The reactivation cross-section, a regional-scale east-dipping inverted normal fault
of inverted normal fault or even, structures associated with short-cut defined as the Guacate fault (Duhart et al., 2018) involves and exposes
faults transferred shortening as a west-directed thrusting system. large inversion anticline-and-syncline folds widely recognized in the
Then, a second east-verging antiformal stack duplex is the most prom­ Z1020 seismic line (Fig. 7) and at the surface (Fig. 3) (i.g., the Cerritos
inent structural feature. The fact of proposing a system duplex is based Bayos syncline). To the west and associated with the Guacate fault, a
on two main structural characteristics. The first is that the sequence west-directed short-cut fault generates the major positive
behaves like an east-dipping monocline sequence with angles ranging topographic-break on the western flank of the outer Andean forearc
from 60◦ to 80◦ (Fig. 11a and b), in which there are some (transitional boundary between the Central Depression and the
intraformational-shear planes associated with detachment-thrust faults Domeyko Cordillera). Similarly, in the easternmost part of the balanced
(Fig. 11c). The second is that there are slight bedding-dip fluctuations on cross-section, a regional-scale west-dipping inverted normal fault,
each side of the shear plane. The greater angle-dips are associated with named the Chintoraste fault (Duhart et al., 2018) is interpreted, which is
the intraformational detachment level or floor thrusts. This character­ characterized also to form a buttressing structure on its hanging-wall.
istic may be recognized from subvertical-reverse slip fault striations Both inverted normal faults controlled the accumulation of syn-rift de­
(Fig. 11a). posits during the Mesozoic syn-extensional stage. This depositional ge­
ometry is characterized by the good exposure of westward-thickening
4.2. Balanced cross-section and 2-D pre-shortening restoration syn-tectonic wedges, which are recognized along the San Salvador River
and the Río Loa Canyons within the Middle Jurassic sequences (Cerro
4.2.1. The Río Loa and the San Salvador Canyons balanced cross-section Campamento Fm.) (Figs. 4 and 6). Overlying to the syn-rift deposits, we
The Río Loa and the San Salvador Canyons balanced cross-section have described and interpreted a post-rift sequence correlated with the
(Fig. 12a and b) was constrained at depth by the interpretation of the Upper Jurassic and Lower Cretaceous stratigraphic units (the Cerritos
Z1020 seismic line (see Fig. 7). This profile mainly shows a structural Bayos and the San Salvador Fms.). This interpretation is based on the
style associated with first-order positive tectonic inversion structures. parallel reflectors observed in the Z1020 seismic line (Fig. 7) and in the
Two opposite-dipping inverted normal faults controlled the deformation absence of thickness-variations in their deposits (Fig. 4b and 6b). The

10
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 9. Inversion structures recognized in the Eastern domain of the Quebrada Mala segment. (a) Inverted normal faults, short-cut fault and minor inversion
structures. (b) and (c) Evaporitic level as detachment-thrust fault and diapiric body. (d) Fault-propagation fold and back-thrust fault.

combined effects between these two opposite-dipping inverted normal east-directed thrusting- or inversion-faults and east-verging folds. The
faults provoked the formation of opposite-directed thrust faults, such as amount of shortening by detachment-related faults and folds would be
back-thrust faults, detachment faults, and pop-up structures, among conditioned by the positive reverse-reactivation of the Chintoraste fault,
others, thus partly extruding the Middle Jurassic syn-rift deposits and thus concentrating part of shortening through east-directed thin-skinned
entirely affecting the Upper Jurassic-Lower Cretaceous post-rift se­ thrusting. Thus, considering the fieldwork data and the Z1F020 seismic
quences. Finally, to the east of this balanced cross-section, east-directed line interpretations, the structural interpretation proposed in this
thrusting- or inversion-faults are highlighted structures. This kind of balanced-cross section, suggest that both the Guacate and Chintoraste
structural setting would be controlled by the west-dipping bedding of faults correspond to a basement-involved inverted normal fault. A total
Mesozoic syn-rift deposits and, mainly, by the evaporitic levels of ca. 11% of minimum shortening was estimated (or ca. 3200 m lateral
composed by gypsum beds. These levels were able to generate different contraction). We cannot rule out that an important part of the short­
east-directed detachment faults and related east-verging folds. In fact, ening was taken up and accommodated by buttressing structures and
much of the shortening is accommodated by intraformational shear their associated intraformational detachment-thrust level. However, this
planes near parallel to the beddings developed within the Middle is often difficult to determine and quantify because of the complex ge­
Jurassic sequences (Cerro Campamento Fm.). ometry of buttressing structures, that only very rarely provide reliable
The pre-shortening restoration (Fig. 12c) shows that most of the constraints for a correct restoration of balanced cross-sections to their
shortening was accommodated by three types of contractional struc­ pre-thrusting templates (Yakovlev et al., 2023).
tures: (1) inversion anticline-and-syncline folds, (2) partially inverted
normal faults related to the Guacate and the Chintoraste faults, and (3)

11
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 10. The Quebrada Mala detachment-thrust fault.


(a) Transversal section showing the stratigraphic
contact through an angular unconformity between
the underlying Upper Jurassic-Lower Cretaceous Lla­
nura Colorada Fm. and the Upper Cretaceous Que­
brada Mala Fm. (b) Foliated and breached basal
conglomerate levels belonging to the Quebrada Mala
Fm. (c) S–C fabric shear zone parallel to the angular
unconformity. (d) and (e) Generalized view of
detachment-thrust fault showing angular variations
associated with minor thrust structures.

4.2.2. The Quebrada Mala balanced cross-section post-rift deposits in the footwall (see next section). A last inversion
The Quebrada Mala balanced cross-section (Fig. 13a and b) shows structure corresponds to west-directed short-cut faults, which are asso­
that the main deformational mechanisms correspond to west-verging ciated with the main east-dipping inverted normal fault. This fault could
positive inversion structures and thin-skinned east-directed thrust sys­ explain the positive topography, and therefore, the westward-inclined
tems. In addition, the depositional geometry of the involved tectono­ slope recognized on the western flank of the Domeyko Cordillera.
sequences have been generalized and interpreted. These short-cut faults could correspond to the Culebrón and the Sar­
First, as for positive tectonic inversion structures, an east-dipping gento Aldea faults defined by Basso (2004) to the south of this study
inverted normal fault defined as the Sierra del Buitre fault (Marinovic area. Another description proposed in the balanced cross-sections
and García, 1999) standouts. This fault corresponds to an ancient correspond to the depositional geometry of the Upper Jurassic-Lower
east-dipping normal fault that controlled the accumulation of the Cretaceous sequences, which has been interpreted as a post-rift depo­
syn-rift deposits through half-graben rift basins. Then, this structure was sit. Similarly, to the Río Loa and the San Salvador Canyons balanced
partially reactivated as a reverse fault causing the inversion of rift basins cross-section, the Upper Jurassic to Lower Cretaceous stratigraphic units
and subsequent extrusion of the syn-rift fill. We interpreted that both the may be part of the sag-type deposits, and its accumulation would occur
Triassic and Jurassic syn-rift sequences were deformed as inversion after and concordantly to the Jurassic syn-rift deposition. Finally, an
anticlines-and-synclines. Even if the Triassic syn-rift sequences are not important proposed structural feature consists of a thin-skinned, east-­
exposed in the Quebrada Mala segment, we have correlated and directed thrust system. In this balanced cross-section, the interpreted
extrapolated these units recognized to the east of our study area (e.g., thrusting is controlled by an east-directed detachment-thrust fault
Las Lomas Beds, Marinovic and García, 1999). As second-order struc­ developed along the angular unconformity that separates the underlying
tures, short-cut and by-pass faults accommodated an important part of Middle-Upper Jurassic syn-rift and post-rift sequences from the over­
the positive inversion-related shortening. The by-pass fault (i.e., fault lying Upper Cretaceous deposits. Specifically, formed imbricate faults
defined as the Quebrada Perdida fault in Marinovic and García (1999)) allow us to propose that this kind of thrust system may be defined as an
put in contact with the syn-rift deposits in the hanging-wall over the active roof duplex. The influence of mechanically weak gypsum layers,

12
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 11. The Quebrada Mala thrusting structures. (a) and (b) Some evidence recognized in east-dipping monocline sequences with angle ranging from 60◦ to 80◦ to
the west, which are part of an antiformal stack duplex system. (c) Intraformational-shear planes associated with detachment-thrust fault.

within the underlying post-rift sequence, controlled the formation of 5. Discussion


overlying thrust systems.
Similarly, to the Río Loa and the San Salvador Canyons segment, the 5.1. Shortening distribution along inversion tectonic structures
pre-shortening restoration (Fig. 13c) shows that the highest shortening
is mainly controlled by west-directed short-cut and by-pass faults and, to The western flank of the Domeyko Cordillera (between 22◦ 00′ -
a lesser extent, by inversion anticlines-and-synclines folds. These faults 23◦ 50′ Lat. S) shows first-order tectonosequences related to transitional
are associated with the partial inversion of the Sierra del Buitre fault. events that occurred between the final phase of an extensional regime
The shortening generated from the east-directed thin-skinned thrust and the initial phase of a compressional regime; these two events are
system is accommodated by the Quebrada Mala detachment fault. The included in a positive inversion tectonic process. The result and inter­
amount of shortening accommodated by anticlinal stack and imbricate action between kind-of-sequence (i.e., syn-rift, post-rift, or synorogenic
fan duplexes is greater than the shortening taken up by an inversion sequences) and deformation mechanisms triggered the formation of
structure. However, the thrusting system is defined as a thin-skinned thin-skinned thrust systems. Thus, several and complex shortening-
shortening-system, thus only affecting the Upper Cretaceous cover related structures controlled the relationship and distribution between
sequence. Instead, the inversion structures correspond to basement- the Lower-Middle Jurassic syn-rift sequences, the Upper Jurassic-Lower
involved faults, and therefore may generate greater shortening and Cretaceous post-rift sequences, and the Upper Cretaceous-early Ceno­
uplift. The combined effect between west-directed short-cut and inver­ zoic synorogenic sequences (Figs. 12 and 13). Equally, the emplacement
ted normal faults with the detachment-thrust fault would provoke a of the Upper Cretaceous-Paleocene hypabyssal intrusives may be linked
west-verging wedge structure. In such structural setting, the Quebrada with this kind of structural setting. Thus, these are two structural styles
Mala detachment fault is defined as an east-directed passive-roof thrust. recognized on the western flank of the Domeyko Cordillera, which are
A total of ca. 56% of minimum shortening was estimated equivalent to directly related to the development of the tectonosequences. The first
ca. 6000 m of lateral contraction. However, this calculated amount of structures correspond to inherited- and reverse-reactivated normal
shortening may be an overestimated value if only the deformation for faults, and their associated positive tectonic inversion structures, that
the Upper Cretaceous sequences is considered, since when considering controlled the positive tectonic inversion of the Mesozoic rift basins and
only positive tectonic inversion structures, associated with the Mesozoic the extrusion of the Upper-Middle Jurassic syn-rift deposits. The second
rift basin, the minimum shortening is ca. 4,3% or ca. 450 m lateral structures consist of east-directed thin-skinned thrust systems, affecting
contraction. Moreover, we cannot rule out that an important part of the mainly the Upper Cretaceous to Eocene sequences.
shortening was taken up by the formation of a west-verging wedge The east-dipping inverted normal faults accommodated a huge
structure. shortening through inversion anticlines-and-syncline folds deforming
the Middle-Upper Jurassic syn-rift deposits. To a lesser extent, but­
tressing structures, short-cut faults, and by-pass faults accommodated a
minor part of this shortening. Generally, inversion structures have a
westward vergence; however, second-order structures such as back-

13
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 12. The Río Loa and San Salvador Canyons segment on the western flank of the Domeyko Cordillera in northern Chile (see location in Fig. 3). (a) Three-
dimensional view of the structure of the Río Loa and San Salvador Canyons segment showing the correlation of the surface structures with those interpreted to
depth. (b) Balanced cross-section showing the structure interpreted for the Río Loa and San Salvador Canyons segment based on the integration of fieldwork data and
seismic line interpretations. (c) Pre-shortening restoration at the top of the Mesozoic rift-related deposits, indicating the initial position of the main half-graben
structures in the western flank of the Domeyko Cordillera, which controlled the shortening distribution during the Andean forearc orogenic growth.

thrust faults, fault-propagation folds, and detachment-thrust faults Upper Cretaceous cover sequences conditioned the nucleation of the
associated with buttressing structures occasionally show an eastward Quebrada Mala detachment-thrust fault. Similarly, in the Río Loa
vergence. Furthermore, the west-directed short-cut faults may be Canyon, the thrust structures are also conditioned by detachment levels
interpreted as relief-controlled thrust faults since these faults mark the that occur on these post-rift-deposits (San Salvador Fm). Many of these
main topographic break between the Central Depression and the sedimentary accumulations, in other Andean provinces (e.g., the Salta
Domeyko Cordillera. West-verging tectonic inversion structures have rift system; the Neuquén Basin; the Aconcagua and Malargüe Fold and
been recognized in other areas along the western flank of the Domeyko Thrust Belt, among others) have been defined as sag-type deposits
Cordillera (Amilibia et al., 2008; Fuentes et al., 2018; López et al., 2019, (Ramos et al., 1996; Cristallini et al., 1997; Giambiagi et al., 2005, 2008;
2022; Martinez et al., 2021a, 2022b), thus playing a fundamental role among others). Deposition was followed by thermal subsidence-related
during the Upper Cretaceous orogenic growth of the forearc of the accumulation and occurred immediately after tectonic subsidence by
western Central Andes. On the other hand, the well-exposed outcrops Mesozoic extensional faulting; hiding normal faults that control the
and balanced cross-section presented in this study, support that the half-graben rift basins. Similar interpretations are proposed in other
Upper Jurassic-Lower Cretaceous sequences (the Cerritos Bayos and the basins and sequences worldwide, which have a tectonic control marked
Llanura Colorada Fms., and the San Salvador Beds) correspond to a on their deposits. For example, the Lower Cretaceous Abu Gabra For­
post-rift sequence. For example, the reflectors recognized in the Z1020 mation in the Fula Sub-basin Muglad Basin, in southern Sudan (Wu
seismic line (Fig. 7) do not show important variations in their thickness, et al., 2015), in the Sichuan Basin Southwest China (Liu et al., 2021), the
allowing us to propose that the transitional marine-continental strati­ Santos and Campos Basins, both in offshore Brazil (de Paula Faria et al.,
graphic unit corresponds to a post-rift sequence, where there is no 2017; Strugale et al., 2021), the Lower Cretaceous Half Grabens of the
relevant tectonic subsidence due to normal faulting. Complementary, in Southeastern Continental Margin of Brazil (López-Gamundi and Barra­
the Quebrada Mala creek segment, the angular unconformity plane be­ gan, 2012) in the Northern Apennines of Italy (Scisciani et al., 2019;
tween the equivalent post-rift deposits (Llanura Colorada Fm.) and the Tavarnelli et al., 2019), among others. Therefore, the post-rift deposits

14
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Fig. 13. The Quebrada Mala segment on the western flank of the Domeyko Cordillera in northern Chile (see location in Fig. 3). (a) Three-dimensional view of the
structure of the Quebrada Mala segment showing the correlation of the surface structures with those interpreted to depth. (b) Balanced cross-section showing the
structure interpreted for The Quebrada Mala segment based on the integration of fieldwork data. (c) Pre-shortening restoration at the top of the Mesozoic rift-related
deposits, indicating the initial position of the main half-graben structures in the western flank of the Domeyko Cordillera, which controlled the shortening distri­
bution during the Andean forearc orogenic growth.

exposed on the western flank of the Domeyko Cordillera represent the fault that activated a regional-scale angular unconformity along which
transition from an extensional to a compressional regime. The pure the overlying Upper Cretaceous sequences were detached from the un­
thrusting structures are strongly controlled by the rheological conditions derlying Upper Jurassic-Lower Cretaceous sequences. This structure is
of these sequences. named “the Quebrada Mala detachment-thrust fault”. Similarly, in the
The occurrence of east-directed thrust systems was conditioned by Río Loa Canyon, these structures are not only restricted to the Eocene
two stratigraphic- and structural-settings. The first illustrates that thrust sequences, but also to several duplex systems, and detachment folds
faults only affect the Upper Cretaceous-early Eocene synorogenic se­ affect the Upper Jurassic–Lower Cretaceous sequences. In both
quences, while the second shows much of these systems occur in a cross-sections, the thrust systems occur in a position above the footwall
footwall block of pre-existing Mesozoic half-graben basins, nucleating of normal faults or half-graben basins. Moreover, the Eocene sequences
from a footwall-position of those inherited normal faults. The first and their depositional geometries show a synorogenic accumulation
setting is based on the good exposure of the Upper Cretaceous-Eocene (growth strata) related to the uplift of inversion structures (i.g., the
stratigraphic units outcropping mainly in the Quebrada Mala area and Chintoraste fault), thus indicating a continuous deformation and syn­
around the Río Loa Canyon. In both sectors, these units are defined as orogenic accumulation from the Upper Cretaceous to the early Eocene
part of synorogenic sequences-involved thrust systems, where the main time interval. In both segments, the Upper Jurassic-Lower Cretaceous
structural styles correspond to overlapping antiformal stack duplexes post-rift sequences play a first-order role in the generation of
and imbricate fan systems (sensu Boyer and Elliott, 1982). The second east-directed thin-skinned thrusting, a ‘passive roof thrust’, and the
setting is proposed from our balanced cross-section. For example, in the subsequent formation of a possible ‘intracutaneous wedge’. However,
Quebrada Mala segment, we have shown that the formation of this interpretation represents a first approximation of the deep structure
thin-skinned thrust systems is controlled by a major detachment-thrust under the western flank of the Domeyko Cordillera and it needs to be

15
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

tested incorporating new field and geophysical data. Previous studies geometries associated with synorogenic sequences or growth strata
have shown the importance of rheologically weak horizon in the (Mpodozis et al., 2005; Arriagada et al., 2006; López et al., 2020), and
deformation mechanisms of fold-thrust belts (e.g., Butler, 1982, 1989; thermochronology ages (Henriquez et al., 2018; Martínez et al., 2022a)
Bonini, 2001; Couzens-Schultz et al., 2003; Yan et al., 2016; among support that the eastern border of the Domeyko Cordillera is the result of
others). a prominent uplift- and shortening-related structure activated since the
Deformation mechanisms associated with thrust systems and the late Cretaceous Peruvian tectonic phase (Steinmann, 1929; Mpodozis
relationship that these have with positive tectonic inversion structures et al., 2005; Bascuñán et al., 2015). However, our descriptions and in­
represent highlighted structural features widely recognized in many terpretations would support that the western flank of the Domeyko
other subduction-related active orogens worldwide such as the Kirthar Cordillera also represent an important morpho-tectonic limit. Along
and Sulaiman Mountain belts, Pakistan (Banks and Warburton, 1986), with this region, similar first-order structures are also recognized, thus
the Interandean and Subandean Zones of the Central Andean fold and controlling the initial stages of orogenic growth of a segment of the
thrust belt of Bolivia (Baby et al., 1992; Rojas Vera et al., 2019), the western Central Andes. These structures uplifted the ancient Mesozoic
Argentinian foreland basins (Giambiagi et al., 2003; Horton and Fol­ graben or half-graben basin fill and trigged the deformational mecha­
guera, 2022), in the Jura fold-and-thrust belt in the northeast French nisms that affected the Upper Cretaceous and early Cenozoic sequences.
Alps (Madritsch et al., 2008) in the external northern Alpine foreland of In addition, the most important proposed structures for the orogenic
Switzerland (Malz et al., 2016), in the Central-Northern Apennines of growth for the Domeyko Cordillera, and generally, for the Andean
Italy (Scisciani, 2014; Scisciani et al., 2019; Barchi and Tavarnelli, 2022; forearc have been proposed to result from positive tectonic inversion,
Pace et al., 2022a, 2022b), among others. In the outer Andean forearc of due to reverse-reactivations of pre-orogenic normal faults (Amilibia
northern Chile, the interaction between these two contractional struc­ et al., 2008; Fuentes et al., 2018; López et al., 2019, 2022; Martínez
tural styles had rarely been tackled; nonetheless, the mention of thrust et al., 2020a; Martínez et al., 2021b). In this study, we want to propose
systems is scarce (Fuentes et al., 2018; Martinez et al., 2021a; 2022b). that these orogenic processes, in turn, conditioned the generation of
The thrust systems proposed in this work are conditioned by the tectonic thin-skinned thrust systems affecting mainly the Upper Jurassic-Lower
inversion structures that controlled the extrusion of the Mesozoic Cretaceous post-rift deposits and the Upper Cretaceous-early Cenozoic
syn-rift deposits. Inverted normal faults have two mechanisms to synorogenic sequences. A large amount of shortening provoked by
generate thrust systems towards the cover deposits, either post-rift or basement-involved inversion structures was accommodated through the
synorogenic sequences. The first is related to west-directed short-cut thin-skinned thrust systems. These thrust systems are a consequence of
faults, which may form hybrid fold-and-thrust belt systems. The second the propagation of thick-skinned structures towards shallower late
corresponds to high-angle reverse-reactivated normal faults that control Cenozoic stratigraphic levels. The newly-formed low-angle thin-skinned
the development of buttressing structures, which internally generates thrust structures would have attenuated the shortening-effect and many
intraformational detachments affecting the synorogenic sequences. of these structures were hidden under the late Cenozoic gravel se­
quences. Only some minor shortening-related deformation features
5.2. Insights for the orogenic growth of the andean forearc could indicate possible minor reactivations (López et al., 2019). Typi­
cally, thrust systems are associated with basement-involved thrust
The Z1020 seismic line re-interpretation (Fig. 7) and structures faults, comparable to a ‘leading imbricate fan system’, and therefore are
recognized affecting the Upper Cretaceous and Eocene sequences typical in foreland basins. The formed thrusting structures will occur
(Fig. 5e and 11) reveal that the western flank of the Domeyko Cordillera internally in synorogenic deposits, resulting from uplift and erosion by
is controlled by large west-directed reverse-reactivated normal faults, basement-block or older rocks. For example, in neighboring regions as
short-cut faults, and east-directed thin-skinned thrust systems. The latter the thrust systems proposed for the Salar de Atacama Basin and the
are genetically related to the partial reverse-reactivation of pre-existing Cordillera de la Sal. Thrust faults with ramp-flat-ramp geometry uplifted
normal faults and the inherited-structural position of the footwall of the Permian-Triassic rocks, generating an accumulation of the Upper
half-graben basin (Figs. 12 and 13). The pre-shortening restorations of Cretaceous-Cenozoic synorogenic sequences and the formation of
the balanced cross-sections indicate that the region accumulated be­ folding-and-thrusting systems as the Cordillera de la Sal (Mpodozis
tween 4% and 11% of crustal shortening (Fig. 12c and 13c), considering et al., 2005; Arriagada et al., 2006; López et al., 2020). Similarly, in the
only the restoration of inversion structure related to the Mesozoic rift Potrerillos Fold and Thrust Belt, thrusting structures are formed by
basin. These values are similar to those reported by Fuentes et al. (2018) basement-involved thrust faults associated with the shortening and ris­
and Martinez et al., 2021a to the west of the Sierra de Moreno range, ing of the Sierra Castillo Basement Block (Martínez et al., 2020a; López
immediately to the north of this study area. The shortening calculated et al., 2022). In the case of the east-directed thin-skinned thrust system
here represents a relatively modest amount of contraction if compared proposed in this work, there are no Paleozoic basement blocks that have
to the shortening calculated in the Pre-Andean Basins (Amilibia et al., been exposed by shortening- and uplift-related structures and with the
2008; Martínez et al., 2021b). This may be due to the fact that in these same sense of vergence. In fact, basement-involved structures
basins much deformation is achieved by structures developed mentioned here, correspond to west-directed positive tectonic inversion
throughout the entire Cenozoic stratigraphic sequences. On the con­ structures, and are located to the east of the thrust system (e.g., the
trary, when we consider a shortening for all the units involved, specif­ Quebrada Mala segment) and they do not raise Paleozoic basement
ically for the Quebrada Mala segment, the shortening is substantially blocks. Late Cenozoic gravels accumulation hide the deep structures that
greater, reaching a ca. 56% of crustal shortening. In this case, may be occurring below the Central Depression, thus providing re­
thin-skinned thrust systems explain the reason for the higher percent­ striction in the knowledge on the distribution, kinematics and rela­
ages of shortening. These regions commonly lack good constraints on tionship with those systems exposed on the surface. Recently, López
deformation styles, and therefore, structural interpretations underesti­ et al. (2019) proposed structures under the Central Depression associ­
mate the nature of orogenic processes. ated with a positive tectonic inversion process. Nevertheless, Paleozoic
Numerous studies have interpreted that the main morpho-tectonic basement rocks are exposed on the eastern border of the Coastal
break along the forearc of the Central Andes in northern Chile corre­ Cordillera (e.g., Sierra del Tigre Fm., Niemeyer et al., 1977; Cortés et al.,
sponds to the transitional zone between the Domeyko Cordillera and the 2000), which is also controlled by the Atacama Fault System. In this
Pre-Andean Basins (i.g., Salar de Atacama, Punta Negra, Pedernales work, the compelling evidence to support these assumptions is not
Basins). Rigorous evidences obtained from the 2-D reflection seismic available; however, the structural control that the Atacama Fault System
interpretations (Arriagada et al., 2006; Bascuñán et al., 2015; 2019; exerted on the Paleozoic basement blocks recognized in the outer fore­
Martínez et al., 2020a; López et al., 2020, 2022), depositional arc could have had a fundamental role on the development of a thrust

16
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

system such as those proposed in this study. Undoubtedly, we believe thrust systems explain the reason for the higher percentages of
that it is necessary to open the debate about those structures related to shortening in this last segment. In addition, it is possible that an
shortening that may be located towards the outer forearc. The effect of important part of the shortening was taken up and accommodated by
inherited structures on the Andean development could have played a buttressing structures and their associated intraformational
significant role during positive inversion, i.e., in the first stages of detachment-thrust levels.
orogenic growth.
Similarly, strike-slip deformation has also been a structural mecha­ CRediT authorship contribution statement
nism to explain regional-scale high-dipping faults exposed on the
western flank of the Domeyko Cordillera, and immediately to the east of Cristopher López: Writing – review & editing, Writing – original
this study area. We cannot rule out an important role of strike-slip draft, Software, Methodology, Investigation, Formal analysis, Concep­
faulting (Mpodozis and Cornejo, 2012; Niemeyer and Urrutia, 2009; tualization. Renato Cisternas: Software, Methodology, Investigation,
Reutter et al., 1991) during the Eocene-Oligocene reactivations, thus Conceptualization. Sebastián Espinoza: Software, Methodology,
allowing the emplacement of subvolcanic intrusives (Maksaev and Formal analysis. Rodrigo González: Writing – review & editing,
Zentilli, 1999; Mpodozis and Cornejo, 2012; Mpodozis et al., 1993). Investigation, Conceptualization. Fernando Martínez: Writing – review
However, strike-slip deformation could have a fundamental role in & editing, Investigation, Conceptualization. Rodrigo Riquelme:
transitional events that occurred during the positive tectonic inversion. Writing – review & editing, Investigation, Conceptualization. Daniela
Strike-slip deformation would be a common kinematic behavior during Montenegro: Writing – review & editing, Investigation, Conceptuali­
the change between an extensional regime to a contractional regime, zation. Jorge Morales: Writing – review & editing, Investigation.
involving a phase of strike-slip faulting (Peacock et al., 2017).
Finally, we want to highlight that the sub-volcanic intrusive bodies
(e.g., the Sierra Gorda, Spence, and the Lomas Bayas mines, see Fig. 3), Declaration of competing interest
associated with the Upper Cretaceous-Paleocene metallogenetic belt
(Sillitoe, 1992; Sillitoe and Perelló, 2005; Camus, 2003), were emplaced The authors declare that they have no known competing financial
in a structural position related to partly or entirely reverse-reactivated interests or personal relationships that could have appeared to influence
normal faults and/or inversion structures such as short-cut faults or the work reported in this paper.
inversion anticline-and-syncline folds (Fig. 13). The position and dis­ Cortés, 2000, Niemeyer et al., 1997
tribution of the Upper Cretaceous and Paleocene magmatic and volcanic
arc had a strong structural control, and its development would be related Data availability
to positive tectonic inversion processes, and possibly to the continued
development of thin-skinned thrust systems during the early Cenozoic. Data will be made available on request.

6. Concluding remarks Acknowledgment

The descriptions and interpretations presented in this study propose This work was supported by the FONDECYT-ANID, Chile project
a new model of compressive deformation in the western flank of the grant N◦ 1220987, “Geomorphology on the geologic times scales: new
Domeyko Cordillera. These structures are different from the previously perspectives on the geomorphological significance of planation surfaces
formulated structural models that have only interpreted strike-slip as a marker of tectonic and climatic events”. The lead-author thanks the
faulting and tectonic inversion processes along this portion of the support of the VRIDT-UCN2095 project grant 10201295. The authors
outer forearc in northern Chile. Our presented models and structural also thank Enap-Sipetrol for providing the seismic data used in this
mechanisms provide new concepts of tectonic evolution and develop­ study. Finally, the authors thank the constructive suggestions made by
ment of the western Central Andes. The most relevant innovations, Enrico Tavarnelli, Paolo Pace, and Vittorio Scisciani. Special thank is
based on the results of our investigations on the structural geometry and given to Cecilia Avila for valuable tips and reviewing the English.
inferred deformation mechanisms that characterize the western flank of
the Domeyko Cordillera are summarized as follows:
References

1. Two structural styles are recognized on the western flank of the Aguirre-Urreta, M., 1993. Neocomian ammonite biostratigraphy of the andean basin of
Domeyko Cordillera. These structures differ both in kinematics and Argentina and Chile. Rev. Esp. Palaontol. 8 (1), 57–74.
because of involved stratigraphic sequences. The first correspond to Allmendinger, R., Jordan, T., Kay, S.M., Isacks, B., 1997. The evolution of the altiplano-
puna plateau of the central andes. Annu. Rev. Earth Planet Sci. 25, 139–174.
west-directed positive tectonic inversion structures, which expose Amilibia, A., 2002. Inversión Tectónica en la Cordillera de Domeyko, Andes del Norte de
and extrude the Middle-Upper Jurassic syn-rift deposit of the ancient Chile. PhD thesis. Universitat de Barcelona.
Mesozoic rift basin; the second correspond to east-directed thrust Amilibia, A., Sabat, F., McClay, K.R., Muñoz, J.A., Roca, E., Chong, G., 2008. The role of
inherited tectono-sedimentary architecture in the development of the central
systems, affecting mainly the Upper Cretaceous to Eocene volcano- Andean mountain belt: insights from the Cordillera de Domeyko. J. Struct. Geol. 30,
sedimentary synorogenic sequences. 1520–1539.
2. The inversion structures consist of reverse-reactivated normal faults, Arabasz, W., 1971. Geological and Geophysical Studies of the Atacama Fault Zone in
Northern Chile. Unpublished PhD thesis, California Institute of Technology,
inversion anticlines-and-synclines folds, buttressing structures, and Pasadena, California, p. 275.
intraformational detachment-thrust faults. Ardill, J., 1996. Sequence Stratigraphy of the Mesozoic Domeyko Basin, Northern Chile.
3. The thrust systems mainly consist of antiformal stack duplexes. The Ph.D Thesis (Unpublished). University of Liverpool, p. 245.
Ardill, J., Flint, S., Chong, G., Wilke, H., 1998. Sequence stratigraphy of the Mesozoic
formation of these thrusting structures is controlled by a major Domeyko basin, northern Chile. J. Geol. Soc. Lond. 155, 71–88.
detachment fault developed along a regional-scale angular uncon­ Arriagada, C., Cobbold, P., Roperch, P., 2006. Salar de Atacama basin: a record of
formity; here indicated as the Quebrada Mala detachment fault. compressional tectonics in the central Andes since the mid-Cretaceous. Tectonics 25,
1–19.
4. The interaction between inversion structures and thrust systems
Arriagada, C., Roperch, P., Mpodozis, C., Cobbold, P.R., 2008. Paleogene building of the
promoted the development of a west-directed intracutaneous tec­ Bolivian orocline: tectonic restoration of the Central Andes in 2-D map view.
tonic wedge. Tectonics 27 (TC6014), 14.
5. Finally, the minimum shortening obtained from our balanced cross- Baby, P., Hérail, G., Salinas, R., Sempere, T., 1992. Geometry and kinematic evolution of
passive roof duplexes deduced from cross section balancing: example from the
section restoration is ca. 11% for the Río Loa and the Sal Salvador foreland thrust system of the southern Bolivian Subandean Zone. Tectonics 11 (3),
Canyons segment and ca. 56% for the Quebrada Mala segment. The 523–536.

17
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Baeza, L., 1979. Distribución de facies sedimentarias marinas en el Jurásico de Cerritos Duhart, P., Muñoz, J., Quiroz, D., Mestre, A., Varas, G., 2018. Carta Sierra Gorda, Región
Bayos y zonas adyacentes, norte de Chile. In: Actas 3er Congreso Geológico Chileno, de Antofagasta. Servicio Nacional de Geología y Minería, Carta Geológica de Chile.
pp. H45–H61. Arica. 2. In: Serie Geología Básica 198, 1 Mapa Escala 1: 100.000. 1 CD con anexos Santiago.
Bailey, C.M., Giorgis, S., Coiner, L., 2002. Tectonic inversion and basement buttressing: Fuentes, G., Martínez, F., Bascuñan, S., Arriagada, C., Muñoz, R., 2018. Tectonic
an example from the central Appalachian Blue Ridge province. J. Struct. Geol. 24, architecture of the Tarapacá Basin in the northern Central Andes: new constraints
925–936. from field and 2D seismic data. Geosphere 14 (6), 2430–2446.
Banks, C.J., Warburton, J., 1986. ‘Passive-roof’ duplex geometry in the frontal structures García, F., 1967. Geología del Norte Grande de Chile. In: Simposio sobre el Geosinclinal
of the Kirthar and Sulaiman mountain belts, Pakistan. J. Struct. Geol. 8, 229–237. Andino, Publicación, vol. 3. Sociedad Geológica de Chile, Santiago, p. 138.
Issues 3–4. Giambiagi, L.B., Ramos, V.A., Godoy, E., Alvarez, P.P., Orts, S., 2003. Cenozoic
Barchi, M.R., Tavarnelli, E., 2022. Thin vs. thick-skinned tectonics in the Umbria-Marche deformation and tectonic style of the Andes, between 33◦ and 34◦ south latitude.
fold-and-thrust belt: contrast or coexistence?. In: Koeberl, C., Claeys, P., Tectonics 22, 1041.
Montanari, A. (Eds.), From the Guajira Desert to the Apennines, and from Giambiagi, L., Alvarez, P.P., Bechis, F., Tunik, M., 2005. Influencia de las estructuras de
Mediterranean Microplates to the Mexican Killer Asteroid: Honoring the Career of rift triásico-jurásicas sobre el estilo de deformación en las fajas plegadas y corridas
Walter Alvarez, vol. 557. GSA, Special Paper. de aconcagua y malargüe, mendoza. Rev. Asoc. Geol. Argent. 60 (4), 662–671.
Bascuñán, S., Arriagada, C., Le Roux, J., Deckart, K., 2015. Unraveling the Peruvian Giambiagi, L., Bechis, F., García, V., Clark, A.H., 2008. Temporal and spatial
phase of the central andes: stratigraphy, sedimentology and geochronology of the relationships of thick-and thin-skinned deformation: a case study from the malargüe
salar de Atacama basin (22◦ 30–23◦ S), northern Chile. Basin Res. 28 (3), 365–392. fold-and-thrust belt, southern central andes. Tectonophysics 459 (1–4), 123–139.
Bascuñan, S., Maksymowicz, A., Martínez, F., Becerra, J., Arriagada, C., Deckart, K., Grocott, J., Brown, M., Dallmeyer, R.D., Taylor, G.K., Treloar, P.J., 1994. Mechanism of
2019. Geometry and late Mesozoic-Cenozoic evolution of the Salar de Atacama Basin continental growth in extensional arcs: an example from de Andean plate-boundary
(22◦ 30′ –24◦ 30′ S) in the northern Central Andes: new constraints from geophysical, zone. Geology 22, 391–394.
geochronological and field data. Tectonophysics 759, 58–78. Gröschke, M., Hillebrandt, A., von, Prinz, P., Quinzio, L.A., Wilke, H.G., 1988. Marine
Basso, M., 2004. Carta baquedano, región de Antofagasta. Servicio nacional de Geología Mesozoic paleogeography in northern Chile between 21◦ -26◦ S. In: Bahlburg, H.,
y minería (SERNAGEOMIN). Serie Geología Básica 1 (82), 100.000. Breitkreutz, C., Giese, P. (Eds.), The Southern Central Andes (Editors), Lecture Notes
Biese, W., 1961. El Jurásico de Cerritos Bayos, vol. 19. Instituto de Geología, Universidad in Earth Sciences, vol. 17. Springer Verlag, Berlin, pp. 105–117.
de Chile, Santiago, Publication. Hayward, A.B., Graham, R.H., 1996. Some geometrical characteristics of inversion. In:
Bonini, M., 2001. Passive roof thrusting and forelandward fold propagation in scaled Cooper, M.A., Williams, G.D. (Eds.), Inversion Tectonics. Geological Society Special
brittle-ductile physical models of thrust wedges. J. Geophys. Res. 106, 2291–2311. Publication Classics, pp. 17–39.
Bonini, M., Sani, F., Antonielli, B., 2012. Basin inversion and contractional reactivation Héja, G., Ortner, H., Fodor, L., Németh, A., Kövér, S., 2022. Modes of oblique inver- sion:
of inherited normal faults: a review based on previous and new experimental models. a case study from the cretaceous fold and thrust belt of the western transdanubian
Tectonophysics 522–523, 55–88. range (tr), west Hungary. Tectonics 41 (3), e2021TC006728.
Blanco, N., 2008. Estratigrafía y evolución tectono-sedimentaria de la cuenca cenozoica Henriquez, S., DeCelles, P., Carrapa, B., 2018. Cretaceous to middle cenozoic exhumation
de Calama (Chile, 22◦ S). Tesis de Master de Geología Experimental. Universidad de history of the Cordillera de Domeyko and salar de Atacama basin, northern Chile.
Barcelona. Tectonics 38, 395–416.
Boyer, S., Elliott, D., 1982. Thrust Systems, vol. 66. American Association of Petroleum Hervé, M., 1987. Movimiento sinistral en el Cretácico Inferior de la Zona de Falla de
Geologists Bulletin, pp. 1196–1230. Atacama al norte de Paposo (24◦ S), Chile. Rev. Geol. Chile 31, 37–42.
Brocher, T.M., Blakely, R.J., Wells, R.E., 2004. Interpretation of the Seattle Uplift, Horton, B.K., Folguera, A., 2022. Tectonic inheritance and structural styles in the Andean
washington, as a Passive-Roof Duplex. Bulletin of the Seismological Society of fold-thrust belt and foreland basin. In: Mora, A., Zamora, G. (Eds.), Andean
America, pp. 1379–1401. Structural Styles A Seismic Atlas. Elsevier, p. 501.
Butler, R.W.H., 1982. The terminology of structures in thrust belts. J. Struct. Geol. 4, Isacks, B., 1988. Uplift of the central andes plateau and bending of the Bolivian orocline.
239–245. J. Geophys. Res. 93, 3211–3231.
Butler, R.W.H., 1989. The influence of pre-existing basin structure on thrust system Jaillard, E., Soler, P., Carlier, G., Mourier, T., 1990. Geodynamic evolution of the
evolution in the Western Alps. In: Cooper, M.A., Williams, G.D. (Eds.), Inversion northern and central Andes during early to middle Mesozoic times: a Tethyan model.
Tectonics, 44. Geological Society of, London Special Publication, pp. 105–122. Journal of the Geological Society of London 147, 1009–1022.
Calamita, F., Scisciani, V., Tavarnelli, E., 2001. Styles of tectonic inversion within syn- Jensen, O., Vicente, J.C., Davidson, J., Godoy, E., 1976. Etapas de la evolución marina
orogenic basins: examples from the Central Apennines, Italy. Terra. Nova 13 (5), jurásica de la cuenca andina externa (mioliminar) entre los paralelos 26◦ y 29◦ 30’S.
321–326, 2001. In: Congreso Geológico Chileno, vol. 1. Actas, Santiago, pp. A273–A293. No. 1.
Calamita, F., Di Domenica, A., Pace, P., 2018. Macro- and meso-scale structural criteria Liu, S., Yang, Y., Deng, B., Zhong, Y., Wen, L., Sun, W., Li, Z., Jansa, L., Li, J., Song, J.,
for identifying pre-thrusting normal faults within foreland fold-and-thrust belts: Zhang, X., Peng, H., 2021. Tectonic evolution of the Sichuan Basin, southwest China.
insights from the Central-Northern Apennines (Italy). Terra. Nova 30, 50–62. Earth Sci. Rev. 213, 103470.
Camus, F., 2003. Geología de los Sistemas Porfíricos en los Andes de Chile. Servicio López, C., Martínez, F., Maksymowicz, A., Giambiagi, L., Riquelme, R., 2019. What is the
Nacional de Geología y Minería, Chile, p. 267. structure of the forearc region in the Central Andes of northern Chile? An approach
Charrier, R., Farías, M., Maksaev, V., 2009. Tectonic, paleogeographic, and metallogenic from field data and 2-D reflection seismic data. Tectonophysics 769, 228187.
evolution during the Cenozoic in the Andes of Central and Northern Chile and López, C., Martínez, F., Del Ventisette, C., Bonini, M., Montanari, D., Muñoz, B.,
implication for the adjacent regions of Bolivia and Argentina. Rev. Asoc. Geol. Riquelme, R., 2020. East-vergence thrusts and inversion structures: an updated
Argent. 65, 5–35. tectonic model to understand the Domeyko Cordillera and the Salar de Atacama
Cobbold, P.R., Rossello, E.A., Roperch, P., Arriagada, C., Gómez, L.A., Lima, C., 2007. Basin transition in the western Central Andes. J. S. Am. Earth Sci. 103 (103),
Distribution, timing, and causes of Andean deformation across South America. 102741.
Geological Society, London, Special Publications 272 (1), 321–343. López, C., Del Ventisette, C., Bonini, M., Montanari, D., Maestrelli, D., Martínez, F., et al.,
Coloma, F., Valin, X., Oliveros, V., Vásquez, P., Creixell, C., Salazar, E., Ducea, M., 2017. 2022. The relationship between inverted normal faults and pure thrusting during the
Geochemistry of Permian to Triassic igneous rocks from northern Chile (28◦ - tectonic inversion of the Domeyko Cordillera, northern Chile: structural and seismic
30◦ 15′ S): implications on the dynamics of the proto-Andean margin. Andean Geol. interpretation and analog modeling experiments. Tectonics 41, e2022TC007378.
44 (2), 147–178. López-Gamundí, O., Barragan, R., 2012. Structural framework of Lower Cretaceous half
Cornejo, P., Matthews, S., Pérez, C., 2003. The "K-T" compressive deformation event in grabens in the presalt section of the southeastern continental margin of Brazil. In:
northern Chile (24◦ -27◦ S). 10◦ Congreso Geológico Chileno (Concepción). Gao, D. (Ed.), Tectonics and Sedimentation: Implications for Petroleum Systems:
Couzens-Schultz, B.A., Vendeville, B.C., Wiltschko, D.V., 2003. Duplex style and triangle AAPG Memoir 100, pp. 143–158.
zone formation: insights from physical modeling. J. Struct. Geol. 25 (10), Madritsch, H., Schmid, S.M., Fabbri, O., 2008. Interactions between thin- and thick-
1623–1644. skinned tectonics at the northwestern front of the Jura fold-and-thrust belt (eastern
Cortés, J.A., 2000. Hoja palestina, región de Antofagasta. Servicio Nacional de Geología France). Tectonics 27, TC5005.
y Minería, Mapas Geológicos 19 (1) mapa escala 1:100.000. Santiago. Maksaev, V., Zentilli, M., 1999. Fission track thermochronology of the Domeyko
Cristallini, E., Cominguez, A., Ramos, V., 1997. Deep structure of the Metan-Guachipas Cordillera, northern Chile: implications for Andean tectonics and porphyry copper
region: tectonic inversion in Northwestern Argentina. Journal of South American metallogenesis. Explor. Min. Geol. 8, 65–89.
Earth Sciences - J S AMER EARTH SCI 10, 403–421. Malz, A., Madritsch, H., Meier, B., Kley, J., 2016. An unusual triangle zone in the external
Dahlstrom, C., 1969. Balanced cross sections. Can. J. Earth Sci. 6 (4), 743–757. northern Alpine foreland (Switzerland): structural inheritance, kinematics and
de Paula Faria, D.L., dos Reis, A.T., de Souza, O.G., 2017. Three-dimensional implications for the development of the adjacent Jura fold-and-thrust belt.
stratigraphic-sedimentological forward modeling of an Aptian carbonate reservoir Tectonophysics 670, 127–143.
deposited during the sag stage in the Santos basin, Brazil. Mar. Petrol. Geol. 88, Marinovic, N., Lahsen, A., 1984. Hoja Calama. Servicio Nacional de Geología y Minería,
676–695. Santiago, Carta Geológica de Chile, p. 58.
del Rey, A., Deckart, K., Arriagada, C., Martínez, F., 2016. Resolving the paradigm of the Marinovic, N., García, M., 1999. Hoja pampa unión. Región de Antofagasta. Servicio
late Paleozoic–Triassic Chilean magmatism: isotopic approach. Gondwana Res. 37, Nacional de Geología y Minería, Santiago, Mapas Geológicos 9, 1, 100.000.
172–181. Marinovic, N., Cortés, J., García, M., 1996. Estudio geológico regional de la zona
del Rey, A., Deckart, K., Planavsky, N., Arriagada, C., Martínez, F., 2019. Tectonic comprendida entre Sierra del Buitre y Pampa San Román. Serv. Nac. Geol. Min.,
evolution of the southern margin of Pangea and its global implications: evidence Informe Registrado IR- 96–8, 1–140.
from the mid Permian-Triassic magmatism along the Chilean-Argentine border. Martínez, F., López, C., Parra, M., 2020a. Effects of pre-orogenic tectonic structures on
Gondwana Res. 76, 301–303. the cenozoic evolution of andean deformed belts: evidence from the salar de Punta
Dewey, J.F., Bird, J.M., 1970. Mountain belts and the new global tectonics. J. Geophys. Negra basin in the central andes of northern Chile. Basin Res. 32 (6), 1–22.
Res. 75, 2625–2647.

18
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Martinez, F., Fuentes, G., Perroud, S., Bascuñan, S., 2021a. Buried thrust belt front of the Ramos, V.A., 2010. The tectonic regime along Andes: present-day and Mesozoic regimes.
western Central Andes of northern Chile: style, age, and relationship with basement Geol. J. 45, 2–25.
heterogeneities. J. Struct. Geol. 147, 104337. Ramos, V.A., Cegarra, M., Cristallini, E., 1996. Cenozoic tectonics of the High Andes of
Martínez, F., Kania, J., Muñoz, B., Riquelme, R., López, C., 2020b. Geometry and west-central Argentina (30–36◦ S latitude). Tectonophysics 259, 185–200. Issues 1–3.
development of a hybrid thrust belt in an inner forearc setting: insights from the Reutter, K.J., Scheuber, E., Helmcke, D., 1991. Structural evidence of orogen-parallel
Potrerillos Belt in the Central Andes, northern Chile. Chile. J. S. Am. Earth Sci. 98, strike slip displacements in the Precordillera of northern Chile. Geol. Rundsch. 80
102439. (1), 135–153.
Martínez, F., Muñoz, B., López, C., González, R., Parra, M., Riquelme, R., 2021b. Riquelme, R., Tapia, M., Campos, E., Mpodozis, C., Carretier, S., González, R., Muñoz, S.,
Complex basement-involved contractional structures in the Pre-Andean basins of Fernandez-Mort, A., Sanchez, C., Marquardt, C., 2017. Supergene and exotic Cu
northern Chile: a review from seismic data. Tectonics 40 (2), e2020TC006433. mineralization occur during periods of landscape stability in the Centinela Mining
Martínez, F., Parra, M., González, R., López, C., Patiño, A., Muñoz, B., Robledo, F., District, Atacama Desert. Basin Res. 30 (3), 395–425.
Sobel, E., Glodny, S., 2022a. Deciphering the late paleozoic–cenozoic tectonic Rojas Vera, E.A., Giampaoli, P., Gobbo, E., Rocha, E., Olivieri, G., Figueroa, D., 2019.
history of the inner central andes forearc: an update from the salar de Punta Negra Chapter 14 - structure and tectonic evolution of the interandean and subandean
basin of northern Chile. Front. Earth Sci. zones of the central Andean foldfold-thrust belt of Bolivia. In: s), Horton, Brian K.
Martínez, F., López, C., Cisternas, R., 2022b. Loa-geo1: a field regional transect to (Eds.), Andrés Folguera, Andean Tectonics. Elsevier, pp. 399–427.
unravel the structure of the western central andes. J. S. Am. Earth Sci. 119, 104011. Sanchez, C., Stéphanie, B., Riquelme, R., Carretier, S., Bissig, T., Lopez, C., Mpodozis, C.,
May, G., 1997. Oligocene to Recent Evolution of the Calama Basin, Northern Chile. Campos, E., Regard, V., Hérail, G., Marquardt, C., 2017. Exhumation history and
University of Aberdeen, UK, p. 230. Ph.D. Thesis (Unpublished). timing of supergene copper mineralisation in an arid climate: new
May, G., Hartley, A.J., Chong, G., Stuart, F., Turner, P., Kappe, S.J., 2005. Eocene to thermochronological data from the Centinela District, Atacama, Chile. Terra. Nova
Pleistocene lithostratigraphy, chronostratigraphy and tectono-sedimentary evolution 30, 78–85.
of the Calama Basin, northern Chile. Rev. Geol. Chile 32, 33–58. Scheuber, E., Andriessen, P.M., 1990. The kinematics significance of the Atacama Fault
McClay, K.R., Buchanan, P.G., 1992. In: McClay, En K.R. (Ed.), Thrust Faults in Inverted zone, northern, Chile. J. Struct. Geol. 21, 243–257.
Extensional Basins, Thrust Tectonics, pp. 419–434. Scheuber, E., González, G., 1999. Tectonics of the Jurassic – early Cretaceous magmatic
Medina, E., Niemeyer, H., Wilke, H., Cembrano, J., García, M., Riquelme, M., arc of the north Chilean Coastal Cordillera (22◦ –26◦ S): a story of crustal deformation
Espinoza, S., Jensen, A., Chong, G., 2012. Cartas Tocopilla Y María Elena, Región along a convergent plate boundary. Tectonics 18, 895–910.
Antofagasta. Escala 1:100.000. Servicio Nacional de Geología y Minería. Scheuber, E., Bogdanic, T., Jensen, A., Reuter, K., 1994. Tectonic development of the
Montaño, J.M., 1976. Estudio geológico de la zona de Caracoles y áreas vecinas con north Chilean Andes in relation to plate convergence and magmatism since the
especial énfasis en el Sistema Jurásico, Provincia de Antofagasta, II Región, Chile. Jurassic. In: Reuter, K., Scheuber, E., Wigger, P. (Eds.), Tectonics of the Southern
Thesis. Departamento de Geología, Universidad de Chile, Santiago. Central Andes. Springer, Heildelberg, pp. 7–22.
Mortimer, C., 1973. The cenozoic history of the southern Atacama desert, Chile. J. Geol. Scisciani, V., 2009. Styles of positive inversion tectonics in the Central Apennines and in
Soc. Lond. 129, 505–526. the Adriatic foreland: implications for the evolution of the Apennine chain (Italy).
Mount, V.S., Martindale, K.W., Griffith, T.W., Byrd, J.O.D., 2011. Basement-involved J. Struct. Geol. 31 (11), 1276–1294.
contractional wedge structural style: examples from the hanna basin, Wyoming. In: Scisciani, 2014. Positive inversion tectonics in foreland fold-and-thrust belts: a
McClay, K., Shaw, J.H., Suppe, J. (Eds.), Thrust Fault-Related Folding: AAPG Memoir reappraisal of the Umbria-Marche Northern Apennines (Central Italy) by integrating
94, pp. 271–281. geological and geophysical data. Tectonophysics 637, 218–237, 2014.
Mpodozis, C., Ramos, V.A., 1990. The andes of Chile and Argentina. In: Geology of the Scisciani, V., Patruno, S., Tavarnelli, E., Calamita, F., Pace, P., Iacopini, D., 2019. Multi-
Andes and its Relation to Hydrocarbon and Mineral Resources, vol. 11. Circum- phase reactivations and inversions of Paleozoic-Mesozoic extensional basins during
Pacific Council for Energy and Mineral Resources, Houston, Texas, pp. 59–90. Earth the Wilson cycle: case studies from the North Sea (UK) and Northern Apennines
Science Series. (Italy). In: Wilson, W.R., Houseman, G.A., McCaffrey, K.J.W., Doré, A.G., Buiter, S.J.
Mpodozis, C., Ramos, V.A., 2008. Tectónica jurásica en Argentina y Chile: extensión, H. (Eds.), Fifty Years of the Wilson Cycle Concept in Plate Tectonics, vol. 470. Geol.
Subducción Oblicua, Rifting, Deriva y Colisiones? Revista Geológica Argentina 63, Soc., London, Spec. Pub., pp. 205–243
479–495. Sillitoe, R.H., 1992. Gold and copper metallogeny of the central Andes - past, present,
Mpodozis, C., Cornejo, P., 2012. Cenozoic Tectonics and Porphyry Copper Systems of the and future exploration objectives (SEG Distinguished Lecture). Econ. Geol. 87,
Chilean Andes, vol. 16. Society of Economic Geologists, Inc. Special Publication, 2205–2216.
pp. 329–360 (Chapter 14). Sillitoe, R.H., Perelló, J., 2005. Andean Copper Province: Tectonomagmatic Settings,
Mpodozis, C., Marinovic, N., Smoje, I., Cuitiño, L., 1993. Estudio geológico-estructural de Deposit Types, Metallogeny, Exploration, and Discovery. Economic Geology 100th
la Cordillera de Domeyko entre Cerro Limón Verde y Sierra Mariposas. Región de Anniversary, pp. 845–890.
Antofagasta: Servicio Nacional de Geología y Minería [Chile], Santiago, p. 282. Steinmann, G., 1929. Geologie von Peru. Kart Winter.
Informe Registrado IR-93-04. Strugale, M., Schmitt, R.S., Cartwright, J., 2021. Basement geology and its controls on
Mpodozis, C., Arriagada, C., Basso, M., Roperch, P., Cobbold, P., Reich, M., 2005. Late the nucleation and growth of rift faults in the northern Campos Basin, offshore
mesozoic to paleogene stratigraphy of the salar de Atacama basin, antofagasta, Brazil. Basin Res. 33, 1906–1933.
northern Chile: implications for the tectonic evolution of the central andes. Tavarnelli, E., Mazzarini, F., Scialoja, E., Isola, I., 2021. Deformation history of a
Tectonophysics 399 (1–4), 125–154. foredeep basin during the incorporation of its deposits within an advancing orogenic
Muñoz, N., 1989. Geología y estratigrafía de las Hojas Baquedano y Pampa Unión, II wedge: The case of the Oligocene-Early Miocene Macigno Costiero Formation,
Región, Antofagasta, Chile. Memoria. Departamento de Geología, Universidad de southern Tuscany, northern Apennines, Italy. Journal of Structural Geology 147,
Chile, Santiago. 104347.
Nalpas, T., Hérail, G., Mpodozis, C., Riquelme, R., Clavero, J., Dabard, M.P., 2005. Tavarnelli, E., Scisciani, V., Patruno, S., Calamita, F., Pace, P., Iacopini, D., 2019. The
Thermochronological Data and Denudation History along a Transect between role of structural inheritance in the evolution of fold-and-thrust belts: insights from
Chañaral and Pedernales (~26◦ S), North Chilean Andes: Orogenic Implications [ext. the Umbria-Marche Apennines, Italy. In: Koeberl, C., Bice, D.M. (Eds.), 250 Million
abs.]: International Symposium on Andean Geodynamics, 6th. Barcelona, Extended Years of Earth History in Central Italy: Celebrating 25 Years of the Geological
Abstracts, pp. 548–551. Observatory of Coldigioco, vol. 542. GSA Special Papers, pp. 191–211.
Naranjo, J.A., Paskoff, R., 1985. Evolución cenozoica del piedemonte andino en la Taylor, G.K., Grocott, J., Pope, A., Randall, D.E., 1998. Mesozoic fault systems,
Pampa del Tamarugal, norte de Chile (18◦ –21◦ S). In: Proceedings 4th Congreso deformation and fault block rotation in the Andean forearc: a crustal scale strike-slip
Geológico Chileno, vol. 5, pp. 149–164. duplex in the Coastal Cordillera of northern Chile. Tectonophysics 299, 93–109.
Niemeyer, H., Urrutia, C., 2009. Transcurrencia a lo largo de la Falla Sierra de Varas Tomlinson, A.J., Blanco, N., Dilles, J.H., Maksaev, V., Ladino, M., 2018. Carta Calama,
(Sistema de fallas de la Cordillera de Domeyko), norte de Chile. Andean Geol. 36 (1), Región de Antofagasta. Servicio Nacional de Geología y Minería, Carta Geológica de
37–49. Chile. Serie Geológica Básica No.199, 1 mapa escala 1:100.000, 1 CD con anexos,
Niemeyer, H., Venegas, R., González, C.R., Aceñolaza, F.G., 1997. Los terrenos Santiago.
paleozoicos del Salar de Navidad, Región de Antofagasta, Chile. Rev. Geol. Chile 24 Uyeda, S., Kanamori, H., 1979. Back-arc opening and the mode of subduction.
(2), 123–143. J. Geophys. Res. Solid Earth 84 (B3), 1049–1061.
Pace, P., Calamita, F., Tavarnelli, E., 2022a. Along-strike variation of fault-related Vicente, J.C., 2006. Dynamic paleogeography of the Jurassic Andean Basin: pattern of
inversion folds within curved thrust systems: the case of the Central-Northern regression and general considerations on main features. Rev. Asoc. Geol. Argent. 61
Apennines of Italy. Mar. Petrol. Geol. 142, 105731. (3), 408–437.
Pace, P., Calamita, F., Tavarnelli, E., 2022b. Shear zone fabrics and their significance in von Hagke, C., Malz, A., 2017. Triangle zones – geometry, kinematics, mechanics, and
curved, inverted basin-derived thrust systems. J. Struct. Geol. 161, 104663. the need for appreciation of uncertainties. Earth Sci. Rev. 177, 24–42.
Pardo-Casas, F., Molnar, P., 1987. Relative motion of the Nazca (Farallon) and South Williams, G.D., Powell, C.M., Cooper, M.A., 1989. Geometry and kinematics of inversion
American plates since late cretaceous time. Tectonics 6, 233–248. tectonics. Geological Society, London, Special Publications 44 (1), 3–15.
Peacock, D.C.P., Tavarnelli, E., Anderson, M.W., 2017. Interplay between stress Withjack, M., Schlische, R., Olsen, P., 2002. Rift basin structure and its influence on
permutations and overpressure to cause strike-slip faulting during tectonic inversion. sedimentary systems. In: Sedimentation in Continental Rifts, vol. 73, pp. 57–82.
Terra. Nova 29, 61–70. Woodward, N.B., Boyer, S.E., Suppe, J., 1989. Balanced Geological Cross-Sections: an
Prinz, P., Wilke, H., Hilebrandt, A., 1994. Sediment Accumlation and Subsidence History Essential Technique in Geological Research and Exploration. American Geophysical
in the Mesozoic Marginal Basin of Northem Chile. Tectonics of the Southem Central Union, Washington, p. 132.
Andes. Structure and Evolution of an Active Continental Margin. Springer Verlag,
Berlin, pp. 219–232.
Ramírez, C.F., Gardeweg, M., 1982. Hoja toconao región de Antofagasta. Servicio
nacional de Geología y minería. Carta Geol. Chile 58 (1), 250.

19
C. López et al. Journal of South American Earth Sciences 127 (2023) 104376

Wu, D., Zhu, X., Su, Y., Li, Y., Li, Z., Zhou, Y., Zhang, M., 2015. Tectono-sequence Yan, D.-P., Xu, Y.-B., Dong, Z.-B., Qiu, L., Zhang, S., Wells, M., 2016. Fault-related fold
stratigraphic analysis of the lower cretaceous Abu Gabra Formation in the Fula Sub- styles and progressions in fold-thrust belts: insights from sandbox modeling.
basin, Muglad Basin, southern Sudan. Mar. Petrol. Geol. 67, 286–306. J. Geophys. Res. Solid Earth 121, 2087–2111.
Yakovlev, F., Gaidzik, K., Voytenko, V., Frolova, N., 2023. Balanced cross-section
restoration in a complicated folded hinterland structure: shilbilisaj profile, Talas
ridge, Caledonian Tien Shan. Terra. Nova 35, 1–14.

20

You might also like