You are on page 1of 20

Journal of the Geological Society

Magmatic arc fault systems, deformation partitioning and emplacement of granitic


complexes in the Coastal Cordillera, north Chilean Andes (25°30'S to 27°00'S)
John Grocott and Graeme K. Taylor

Journal of the Geological Society 2002; v. 159; p. 425-443


doi:110.1144/0016-764901-124

Email alerting click here to receive free email alerts when new articles cite this article
service
Permission click here to seek permission to re-use all or part of this article
request
Subscribe click here to subscribe to Journal of the Geological Society or the Lyell Collection

Notes

Downloaded by on 2 June 2007

© 2002 Geological Society of London


Journal of the Geological Society, London, Vol. 159, 2002, pp. 425–442. Printed in Great Britain.

Magmatic arc fault systems, deformation partitioning and emplacement of granitic


complexes in the Coastal Cordillera, north Chilean Andes (258309S to 278009S)
1 2
J O H N G RO C OT T & G R A E M E K . TAY L O R
1
School of Earth Sciences and Geography, Centre for Earth and Environmental Science Research, Kingston University,
Kingston-upon-Thames KT1 2EE, UK (e-mail: j.grocott@kingston.ac.uk)
2
Department of Geological Sciences, University of Plymouth, Plymouth PL4 8AA, UK

Abstract: Three fault systems were responsible for Permian to Late Cretaceous deformation of the overriding
plate of the Andean convergent margin in the Coastal Cordillera of northern Chile (258309 to 278009S).
Displacements were linked to crustal growth expressed by the emplacement of a sequence of magmatic arcs.
The Tigrillo Fault System, active from Triassic to Early Cretaceous time, was characterized by arc-normal
extension with increasing left-oblique extension (transtension) from Early Jurassic to Early Cretaceous time.
Stretching of the crust created space for Triassic, Early Jurassic and Early Cretaceous arc basins where
epiclastic, volcaniclastic and volcanic sequences accumulated in continental to shallow marine environments.
Tabular plutonic complexes were emplaced by roof uplift–floor subsidence that allowed a vertical transfer of
material in the crust without significant horizontal extension. The Atacama Fault System was initiated at
c.132 Ma as a (mainly) left strike-slip fault during left-oblique extension of the margin. Elongate, tabular
plutonic complexes were emplaced within the Atacama Fault System between c.132 and c.106 Ma, again by
roof uplift–floor subsidence mechanisms. Ductile–brittle transitions in synplutonic mylonitic rocks of the
Atacama Fault System provided the setting for Kiruna-type Fe-apatite, and Fe oxide (with Cu and/or Au) ores.
The Chivato Fault System was active as an extensional fault system at the eastern side of the Coastal
Cordillera during displacement on the Tigrillo Fault System and later, between c.125 and c.93 Ma, as a
partitioned left-oblique extensional fault system. In post-Early Cretaceous time the Chivato Fault System was
inverted by left-oblique contraction (transpression) when NW-trending transfer faults, some probably
reactivated lateral ramps in the Tigrillo Fault System, accommodated clockwise vertical-axis rotations of 35–
458. Contraction inverted the Atacama Fault System and Tigrillo Fault System and was responsible for west-
vergent, thin-skinned, fold–thrust deformation in stratified rocks throughout the margin.

Keywords: Andes, Coastal Cordillera, island arcs, magmatism, faulting.

The Coastal Cordillera in the south–central Andes of northern aries are characterized by contraction and retreating subduction
Chile provides an exceptional record of overriding plate defor- boundaries by extension in the arc and back arc (Russo & Silver
mation at a convergent margin that has evolved from a retreating 1996). These principles, however, provide only a general guide to
subduction boundary, in Permian to Early Cretaceous time, to an structural style in the overriding plate at convergent margins and
advancing subduction boundary in Late Cretaceous time. Grani- the aim here is to document in detail the response of the
tic magmatism at convergent margins is intrinsic to the growth of overriding plate to deformation at the South American subduc-
continents and is an integral part of Andean geology. However, tion boundary throughout Mesozoic time.
the relationship of magma ascent and emplacement to deforma-
tion and, indeed, the very shape of the intrusions making up
Andean arc batholiths has, until recently, been obscure. Here, we
Geology of the Coastal Cordillera (258309S–278009S)
explore this fundamental relationship and show that this segment Devonian to Carboniferous metasedimentary rocks form the
of the Andes is a prototype for the role of granitic magmatism in basement in this part of the Coastal Cordillera (Figs 2 and 3)
the deformation of the upper plate of continental magmatic arcs. and were originally deposited in a forearc basin (Bahlburg &
We focus on: (1) strains associated with emplacement of plutonic Breitkreuz 1993). Subsequently they were deformed in a late
complexes; (2) the kinematics of magmatic arc fault systems; (3) Carboniferous, SW-vergent thrust belt with gently inclined folds
structures responsible for vertical-axis rotations and partitioning and intense cleavage–bedding anisotropy (Bell 1984). The
of deformation into arc-normal contraction and arc-parallel sequence has characteristics of a mélange SW of Chañaral, where
strike-slip during oblique convergence. it may be part of a Palaeozoic accretionary prism (Bell 1987).
In broad terms it is well known that overriding plate deforma- Permian, Triassic, Jurassic and Early Cretaceous magmatic arc
tion in subduction systems depends on relative rates of subduc- plutonic complexes were emplaced into this metasedimentary
tion and convergence (Dewey 1980; Royden 1993; Waschbusch basement (Berg et al. 1983; Brown 1991; Scheuber & Reutter
& Beaumont 1996). When convergence rate and subduction rate 1992). They have been the focus of several geochronology
differ, the trench migrates with respect to a fixed point in the programmes which have shown that intrusive rocks and synplu-
overriding plate interior (Fig. 1). If the distance between the tonic ductile shear zones, in general, young eastward (Damm &
trench and this point decreases the boundary is an advancing Pichowiak 1981; Berg & Breitkreuz 1983; Naranjo & Puig 1984;
subduction boundary and when it increases the boundary is a Berg & Baumann 1985; Brook et al. 1986; Dallmeyer et al.
retreating subduction boundary. Advancing subduction bound- 1996; Fig. 4). The stratigraphy of corresponding volcanic arc and

425
426 J. G RO C OT T & G . K . TAY L O R

(a) (b)
Upper plate Upper plate
shortening (Vc-Vr>0) extension (Vc-Vr<0)

Fig. 1. (a) Velocity of the upper plate


toward the trench (Vc) exceeds roll-back
continent velocity of the lower plate (Vr) and the
no
advances upper plate is characterized by contractional
trenchward
slab trenchward slab motion of deformation. (b) Lower-plate roll-back (Vr)
roll-back (Vc) roll-back is greater than the velocity of the continent
continental
(Vr) (Vr) interior towards the trench and the upper plate is
(Vc=0) characterized by extension. (After Royden
1993; Russo & Silver 1996).

back-arc rocks of Triassic to Early Cretaceous age has been emplacement of Early Cretaceous plutonic complexes between
revised recently and simplified (Godoy et al. 1997; Godoy & c.132 and c.106 Ma (Grocott et al. 1994; Dallmeyer et al. 1996).
Lara 1998; Lara & Godoy 1998; Marschik & Fontboté 1996; The relationship between ductile and left-slip brittle deformation
Arévalo 1999; Fig. 4). Volcanic rocks east of the Atacama Fault is exemplified on the eastern branch of the Atacama Fault
System, formerly regarded as Cerro Florida Formation (Naranjo System, where left-slip brittle faults rework a 1 km wide,
1978) and correlated with the Bandurrias Formation of Early greenschist to lower amphibolite facies, ductile shear zone at the
Cretaceous age near Copiapó (Segerstrom & Ruiz 1962), have margin of the Sierra Dieciocho plutonic complex (Fig. 3). The
been reassigned to the La Negra Formation. The rocks of the belt of mylonitic rocks has an 40 Ar/39 Ar hornblende isotope
Bandurrias Formation are now regarded as time-equivalent to the correlation age of c.125 Ma (Dallmeyer et al. 1996), identical,
uppermost La Negra Formation and the lower part of the within error, to the U–Pb zircon age for the early intrusive
Chañarcillo Group. They have been reassigned to the Punta del phases of the complex (Berg & Breitkreuz 1983; Fig. 4). High-
Cobre Formation (Fig. 4). temperature ductile shear zones, ductile–brittle shear zones and
Three major arc-parallel fault systems transect this part of the brittle faults all with left-slip kinematics imply a progressive,
Coastal Cordillera: the Tigrillo, Atacama and Chivato fault down-temperature sequence from ductile to brittle deformation
systems (Fig. 2). Faults in each system are linked by NW- and during cooling of the plutonic complex. Late-stage magmatic, Cu
some NE-trending transfer faults. All the fault systems show (Au) and magnetite–apatite mineralization was emplaced in the
evidence of longevity and a history of reactivation. Resolving fault system when the rocks passed through the ductile–brittle
their polyphase history is critical to understanding deformation transition. Manto Verde Mine, a regionally important Cu (Au)
partitioning in the Andean margin and other similar convergent ore deposit (Vila et al. 1997), is located between the left-stepping
boundaries. central and eastern branches of the Atacama Fault System at a
strike-slip relay ramp breached by the Manto Verde fault
(Bonson 1998; Fig. 3).
Tigrillo Fault System The total amount of left strike-slip displacement on brittle
The Tigrillo fault separates Palaeozoic basement from stratified faults of the Atacama Fault System has not been estimated
rocks to the east (Fig. 2) and was recognized in part by Naranjo reliably because cut-off points are difficult to identify (St. Amand
& Puig (1984). It is an east-down, normal-slip fault with Triassic & Allen 1960; Arabasz 1971; Brown et al. 1993). Nevertheless,
to Middle Jurassic sedimentary, volcaniclastic and pyroclastic the discontinuous and overstepping nature of the Atacama Fault
rocks exposed in the hanging wall, and a minimum vertical System fault branches between 258309S and 278009S implies that
separation of 1 km. Displacement on the Tigrillo fault decreases displacement is unlikely to exceed a few kilometres.
to the south and a set of left-stepping, normal-slip and dextral
strike-slip faults (Fig. 2) may link it to east-down, synplutonic,
Chivato Fault System
ductile and ductile–brittle, dip-slip shear zones at the margins of
Las Animas and Las Tazas plutonic complexes (Fig. 3). Ages of The Chivato Fault System (Godoy et al. 1997) was active during
c.160 Ma for the Las Animas plutonic complex and c.132 Ma for emplacement of the Remolino plutonic complex as a ductile
Las Tazas (Dallmeyer et al. 1996; Fig. 4) are consistent with shear zone with dip-slip and left strike-slip displacements (Fig.
increments of normal slip on the Tigrillo fault. The Tigrillo fault 2). The age of this complex is not well constrained, although a
itself, and the overstep system linking it to displacements at the K–Ar biotite age of 93 ! 3 Ma (Lara & Godoy 1998; Fig. 4)
western side of the plutonic complexes, make up the Tigrillo implies that it is younger than the Sierra Dieciocho complex.
Fault System. The Chivato Fault System is similar to the Tigrillo Fault System
and the Atacama Fault System in that the high-temperature
ductile shear zones along it are related to the emplacement of arc
Atacama Fault System
plutons. It further resembles the Atacama Fault System in the
The Atacama Fault System is defined by an arcuate system of emplacement of magnetite–apatite mineralization during the
faults with steeply dipping branches marked by ductile shear ductile–brittle transition in these shear zones (Bonson 1998).
zones reworked by left-slip brittle faults (Naranjo 1987; Brown Unlike the Atacama Fault System, the Chivato Fault System was
et al. 1993; Fig. 2). Ductile deformation was associated with significantly inverted in Late Cretaceous time or later, over-
M AG M AT I C A R C FAU LT S Y S T E M S , C H I L E A N A N D E S 427

printed by brittle reverse-slip displacements and forms part of a Cretaceous time, specifically from an analysis of the emplace-
major belt of left-oblique contraction (transpression) at the ment mechanism of magmatic arc plutonic complexes.
Coastal Cordillera–Precordillera boundary (Godoy et al. 1997;
Arévalo 1999; Fig. 2). This transpressional belt was called the
Central Valley Fault System by Taylor et al. (1998). Some of the Permian and Triassic plutonic complexes
steeply dipping reverse faults in the Chivato Fault System may
Cifuncho
be inverted normal faults of Jurassic to Early Cretaceous age
(Godoy et al. 1997) belonging to the Tigrillo Fault System. Permian and Triassic plutonic complexes comprise monzonite,
Stratified rocks exposed west of the Atacama Fault System granite, granodiorite or tonalite with minor gabbro and diorite
also display west- to NW-vergent folds (Fig. 3) and occasional (Brown 1991). The Cifuncho complex seems to be the oldest of
thrusts consistent with arc-normal to NW–SE shortening. The these, but its age is only poorly constrained by a discordant, U–
style of the structures and the amount of along-strike shortening Pb zircon lower intercept age of 292 ! 14 Ma (Damm &
vary markedly across left-slip, NW-trending transfer faults. Fault Pichowiak 1981). The external contacts are either straight, and
inversion has given rise to buckle folds in the hanging wall of marked by NW-trending faults, or irregular. The latter are
many faults of the Tigrillo Fault System but has not led to characterized by sheet-like apophyses that interfinger with host
reverse-slip separations on any major fault west of the Atacama rocks (Fig. 5a). The sheets are up to 100 m thick with tabular
Fault System (Fig. 3). NE-trending folds in units of the form, steeply dipping edges and gently dipping upper and lower
Chañarcillo Group show that contractional deformation is of contacts parallel to the bedding–cleavage fabric anisotropy in the
post-Albian age, and a maximum age for this deformation of basement (Fig. 3, locality a1 ). No macroscopic magmatic state or
c.81 to c.77 Ma was established by Sylvester & Palacios (1992) crystal–plastic fabrics in the granitic rocks, or emplacement-
and Arèvalo (1999). New 40 Ar/39 Ar geochronology from just east related fabrics in the host rock, were observed. Andalusite and
of the Chivato Fault System near Inca de Oro, and from much muscovite are present in a narrow contact aureole. In contrast,
farther south along the contractional belt near Vallenar (288309S), the contact immediately to the north of this locality is straight
implies that the shortening occurred during emplacement of a and trends NW with a left-separation (apparent displacement) of
suite of Late Cretaceous granitic intrusions in the Precordillera 2.5 km (Fig. 3, locality a2 ). It dips steeply NE, is parallel to
between c.69 and c.65 Ma ago (Gipson 2002). bedding in metasedimentary rocks and there is a narrow thermal
aureole. The dip in the basement flattens out to define an open
synform with an axial surface trace parallel to the contact. This
fold deforms the cleavage as well as bedding, implying that it
NW- and NE-trending faults
post-dates the main cleavage-forming regional deformation.
NW- and NE-trending fault zones are prominent in the Coastal Brittle deformation at the contact is marked by steeply dipping
Cordillera but the NW-trending set is by far the most penetrative minor faults in a 20 m wide fracture zone. Two sets of
(Fig. 2). Most NW faults have left strike-slip displacements that slickenlines are present on fracture surfaces, one down-dip and
post-date emplacement of Lower Cretaceous plutonic complexes one pitching 108NW. Nevertheless, brittle deformation is not
and displacements on the Tigrillo Fault System and Atacama intense and the contact is essentially intrusive, albeit reworked
Fault System. These faults were linked kinematically to left- by minor dip-slip and strike-slip deformation. A lineament,
oblique contraction on the Chivato Fault System (Arévalo 1999). visible on Landsat TM imagery, extends away from the intrusion
Included in this set are two, NW- to NNW-trending fault zones in the same direction as the contact. There are no separations
that substantially offset the Atacama Fault System between 268S across this lineament, either of vertical dykes in the intrusion to
and Taltal (Fig. 2). Brown et al. (1993) estimated a maximum the NW or of the Tigrillo fault and gently dipping strata in its
left strike-slip displacement of 70 km for these faults. This hanging wall to the SE. This rules out significant increments of
estimate was based on restoration of cut-off points of segments both strike-slip and dip-slip post-emplacement displacements on
of the Atacama Fault System with both ductile and brittle the lineament.
characteristics against the NW- to NNW-trending faults. We have
now recognized segments of the Atacama Fault System between
Chañaral–Cachina
the cut-off points originally defined, so that displacement cannot
exceed 20 km (Fig. 2). These major NW- to NNW-trending faults This elongate complex is exposed south of Cifuncho in the
belong to a system that transects the Coastal Cordillera between footwall of the Tigrillo Fault System (Fig. 3). Geochronological
258 and 298S. They are linked to displacements on the Central data are sparse but include a U–Pb zircon age of 230 ! 8 Ma
Valley Fault System at the Coastal Cordillera–Precordillera from the southern part of the complex (Berg & Baumann 1985).
boundary to the east (Taylor et al. 1998). The Chivato Fault The straight, fault-like, NW-trending northern contact is intrusive
System is located within the northern part of the Central Valley but reworked in part by brittle faults (Naranjo & Puig 1984). A
Fault System whereas farther south that fault system is repre- lineament, visible on Landsat TM images, extends to the SE,
sented by the Chañarcillo thrust belt (Arèvalo 1999; Fig. 2). where it is a minor fault that displaces the east-dipping Tigrillo
Although left-slip on most NW-trending faults in the Coastal Fault System, a strip of steeply dipping Triassic volcaniclastic
Cordillera is post-Early Cretaceous in age, and clearly post-dates rocks of the Cifuncho Formation and the lower contact of an
displacement on both the Tigrillo Fault System and the Atacama Upper Jurassic sill in gently dipping Lower Jurassic strata, each
Fault System, many NW faults host late Early Cretaceous with left- or right-separations of less than 200 m (Fig. 3, locality
mineralization, consistent with earlier increments of slip and b1 ). Farther south, a narrow strip of basement separates the
local dilation (Colley et al. 1988). Where such faults curve into eastern contact of the intrusion from a branch of the Tigrillo
the Atacama Fault System, as at Manto Verde (Fig. 3), kinematic Fault System (Fig. 3, locality b2 ). The contact is intrusive, sharp
linkage with the Atacama Fault System during Early Cretaceous and dips c.508E, parallel to bedding and to a 50 m thick granitic
time is implied (Bonson 1998). We now present further evidence sheet in the basement (Fig. 5b). Grain size in the intrusion
that NW-trending faults were active during Permian to Early remains coarse right up to the contact with metasedimentary host
Fig. 2. Overview map of the Coastal Cordillera in northern Chile (258309S–278309S) showing the major plutonic complexes, volcanic, volcaniclastic and
sedimentary stratified units and the major fault systems: the Tigrillo Fault System (TFS), Atacama Fault System (AFS) and the Chivato Fault System
(CFS). Maps compiled from the authors’ own fieldwork, aerial photograph and Landsat image interpretation, and the maps of the Servicio Nacional de
Geologı́a y Minerı́a (Mercado 1978, 1980; Naranjo 1978; Naranjo & Puig 1984; Godoy & Lara 1998; Lara & Godoy 1998).
Jln
base of T projected from 5 km N B AFS C
2
Tac UJm
Tac Tac Jln
Pzm 1
Pzm
Jln Jpa UJm Jln
Tac UJm LKg
Tg Jpa Jln 0
Tg Tg Pzm
Pzm Jpa
Jln -1
Tg
Pzm
Pzm
2 km synplutonic NW-trending
extensional fault fault at W oblique left-slip
with possible margin of fault
AFS internal units, AFS
synplutonic pluton (west) (east)
Las Tazas
component AFS
E (central) F
MVF
2
Las Animas Sierra

M AG M AT I C A R C FAU LT S Y S T E M S , C H I L E A N A N D E S
Las Animas Dieciocho
1 2 3+4 5 1
Flamenco LKg
UJg Jln
Pzm 0
LJg LKg LKg
internal structure unknown UJg
LKg -1
Probable Pzm Probable
Probable Pzm
Probable Pzm
intense fabrics high synplutonic west vergent D1 folds Pzm
crenulated by deformation (schematic - not mapped) high synplutonic deformation
roof-uplift strain overprints D1 overprints D1 in host rocks Las Tazas weak vertical
in host rocks in host rocks intrusive magmatic state
no magmatic state into synplutonic fabric in pluton
fabric in pluton mylonite

Fig. 3. Structural map (see foldout) and cross-sections of the Coastal Cordillera in the Chañaral district of northern Chile (258309S–268459S). The map includes representative orientation data shown on
stereographic projections (Lambert Equal area projections) and palaeomagnetically determined vertical-axis rotations (from Randall et al. 1996). Map and cross-sections were compiled from the authors’
own fieldwork, aerial photograph and Landsat image interpretation, and the maps of the Servicio Nacional de Geologı́a y Minerı́a (Mercado 1978, 1980; Naranjo 1978; Naranjo & Puig 1984; Godoy & Lara
1998; Lara & Godoy 1998).

429
430 J. G RO C OT T & G . K . TAY L O R

50
YPRESIAN

THANET 60 (60.6 ± 1.8)


3
Cabeza deVacas
DANIAN (63 ± 2) 3

post 77Ma
(sinistral)
Hornitos

pression
70

Trans-
Fm.
MAASTRICHTIAN

1
CAMPANIAN "Ocoitas" Intrusion (77 ± 3) ? u/c
80

Cerrillos Fm.
SANTONIAN
CONIACIAN
TURONIAN 90

Sierra de Fraga
Remolino (93.3 ± 3) 2

(sinistral)
CENOMANIAN

tension

Extension in
Trans-
100
ALBIAN 4 onlap
La Borracha (106.3 ± 0.7)
110

Aeropuerto Fm.

(Sinistral)
Bandurrias Fm.
Sierra Dieciocho

Chañarcillo Gp
APTIAN 120 5

tension
(125.7 ± 0.6)

Trans-
4
(126.8 ± 0.5)
Las Tazas (126.8 ± 1.3)
7
BARREMIAN 9
(130.2 ± 0.6)5 (129.2 ± 0.5)
130 (130.3 ± 1.3) 7
HAUTERIVIAN (132.2 ± 0,7) 5
4
Monardes-Moradito
VALANGINIAN (137.6 ± 0.5)
140 (138.9 ± 0.4) 9
(142.1 ± 0.9)

Cobre Fm.
BERRIASIAN

Punta del

Arc-normal
extension
TITHONIAN 150
4
KIMMERIDGIAN Animas (153 ± 0.8) (153.4 ± 0.6)99
OXFORDIAN (155.1 ± 0.7)
7
CALLOVIAN 160 Animas (159.7 ± 1.6)
BATHONIAN

La Negra Fm.
BAJOCIAN 170

AALENIAN Bajocian
Fauna
180 200 m
TOARCIAN above
4 base
Flamenco (east)
190

Arc-normal
(188.7 ± 0.7)

extension?
PLIENSBACHIAN Flamenco (190.1 ± 1.9)
7
Azucar Fm.

hiatus
SINEMURIAN Flamenco (west) 4
Pan de

200 (198.2 ± 0.6)

HETTANGIAN
RHAETIAN
210
Aqua Chica-
Chifuncho

NORIAN 8
Vetado (217 ± 12)
Fm.

220

CARNIAN 8
230 Chañaral (230 ± 8)

LADINIAN
240
6
Chifuncho (292 ± 14)

Stage Time Plutonic Stratigraphical Tectonism


(Ma) Complexes Units

1
K-Ar total rock age (Arévalo 1994)
2,3 2 3
K-Ar biotite ages ( Lara & Godoy 1998; Arévalo 1994)
4,5 40 39 4 5
Ar/ Ar mineral ages ( Dallmeyer et al. 1996 - plutons; Dallmeyer et al. 1996 - synplutonic
mylonites)
6,7,8,9 6 7 8
U-Pb zircon ages ( Damm & Pichowiak 1981; Berg & Breitkreutz, 1983; Berg & Baumann 1985)
9 40 9
Ar/ 39 Ar total rock ages of basaltic andesites in dykes ( Dallmeyer et al. 1996)

Fig. 4. Geochronological and tectonostratigraphic chart for the Coastal Cordillera in northern Chile (258309S–278309S).
M AG M AT I C A R C FAU LT S Y S T E M S , C H I L E A N A N D E S 431

Fig. 5. (a) Contact of Cifuncho pluton and Palaeozoic basement viewed from the east showing sheet-like structure at the pluton margin (locality a1 , Fig.
3). (b) Contact of Cifuncho pluton and Palaeozoic basement looking south near locality b2 (Fig. 3); the contact dips east concordant with bedding in
Palaeozoic metasedimentary rocks. (c) Coarse granite at southern contact of Vetado pluton at locality c, Fig. 3. (d) Roof of Barquito intrusion viewed from
locality d, Fig. 3. (e) Mylonitic migmatite in ductile shear zone at contact of Flamenco pluton with Palaeozoic basement (locality e, Fig. 3). (f) C9 fabrics
in mylonitic shear zone at margin of Flamenco pluton (locality f, Fig. 3).

rocks. These have a narrow metamorphic aureole in which Nearer Chañaral, the same contact is sliced into NE-trending
andalusite is developed. The dip in the host rock decreases away segments by NW-trending faults marked by zones of crush
from the contact to define an open, steeply inclined syncline in breccia and silicification. The trend of contact between fault
bedding and cleavage with an axial surface that trends parallel to pairs is consistently clockwise of the general trend of the contact
the pluton margin. by 20–408, so that vertical-axis rotations, consistent with
432 J. G RO C OT T & G . K . TAY L O R

palaeomagnetic results, could have accompanied left-slip faulting


(Randall et al. 1996). The faults overprint the left-overstep
system in the Tigrillo Fault System (Fig. 2) and, in contrast to
the NW-trending fault at the northern contact of the complex,
show consistent left-separations. Slickenlines confirm these are
oblique left-slip faults (Fig. 3) some of which, traced inboard,
displace plutons as young as at c.93 Ma (Fig. 2).

Vetado
The Vetado plutonic complex lies inboard of the Chañaral–
Cachina complex. It has a U–Pb zircon age of 217 ! 12 Ma
(Berg & Baumann 1985), similar to the stratigraphic age of
Cifuncho and Aqua Chica Formations preserved in structural
basins along the Tigrillo Fault System (Fig. 4). To the west the
contact is subvertical and trends NNE. The host rocks are
intensely ductilely deformed pelitic schists with thin quartzites.
The granite remains coarse grained up to a sharp contact with
metasedimentary rocks (Fig. 5c) that are annealed and have spots
of andalusite and grains of contact-metamorphic muscovite.
Granite veins cut pelitic schists at a high angle but the veins are
not folded, showing that the main schistosity predates emplace-
ment. Close to the contact, andalusite spots are flattened in axial
surfaces of east-vergent minor folds that strike parallel to the
contact. This fabric is interpreted to be due to minor emplace-
ment-related strain. Folding of bedding and the main schistosity
defines an open, NNE-trending synform parallel to the contact
with the intrusion.

Form and emplacement mechanism of Permian and Fig. 6. Criteria allowing identification of synplutonic faults: (a)
Triassic plutonic complexes synplutonic roof uplift produces a synform in the host rocks and/or a
Two of the features of the Permian and Triassic plutonic pluton-side-up dip-slip fault at the pluton margin; (b) a post-
complexes we have described are diagnostic of uplift of the rocks emplacement strike-slip fault offsets the pluton and lithological contacts
forming the pluton roofs to create space for pluton emplacement. away from the pluton by equal amounts; (c) post-emplacement dip-slip
faults displace the pluton and lithological contacts away from the pluton
First, many north–south-trending contacts of these complexes
by amounts depending on the dip of the contacts, and the amount and
terminate against zones of fracturing expressed on Landsat TM
direction of fault slip; (d) synplutonic strike-slip faults have transform-
images as NW-trending lineaments. These fault-like structures
like characteristics with the apparent offset of the pluton opposed to the
show no mappable fault separations away from the pluton (Fig. real shear sense on the fault.
6a), even where they transect geological boundaries with very
different orientations (e.g. Fig. 3, NW and SE of locality a2 ).
This characteristic effectively rules out the possibility of post-
emplacement strike-slip or dip-slip (Fig. 6b and c) or of
synplutonic strike-slip (Fig. 6d), leaving the interpretation that posed along the coast (Fig. 2). The intrusive rocks are tonalites
the fracture zone has accommodated synplutonic vertical displa- to granodiorites with some diorite, and the host rocks comprise
cements, restricted to the contact, during emplacement. A second pelitic to semi-pelitic, strongly deformed to mylonitic schists of
group of structures diagnostic of roof uplift are the open, the Chañaral mélange and Las Tórtolas Formation (Bell 1987).
synformal folds in bedding and cleavage that lie parallel to all Our fieldwork focused on the Flamenco plutonic complex, which
the contacts examined so far, with the exception of the southern has a U–Pb zircon age of 190.1 ! 1.9 Ma (Berg & Breitkreuz
boundary of the Vetado complex. These are interpreted as roof- 1983) and 40 Ar/39 Ar hornblende isotope correlation ages of
uplift folds probably, although not necessarily, associated with 188.7 ! 0.7 Ma and 198.2 ! 0.6 Ma from the western and
roof uplift faults (Fig. 7c and d). Implicit in this emplacement eastern parts of the intrusion, respectively (Dallmeyer et al.
model is that the Permian and Triassic plutonic complexes are 1996).
tabular granites with length and width dimensions that are much At Barquito, schists form a gently dipping roof (Fig. 5d) but at
greater than their thickness (McCaffrey & Petford 1997; Cruden Flamenco, and the north side of Pajonales pluton, the roof is not
1998). exposed and mylonitic foliation dips steeply on all sides (Fig. 3).
The bases of the plutons are not exposed, giving rise to doubt
about their shape. Grocott & Wilson (1997) noted that the
Lower Jurassic plutonic complexes wavelength–thickness relationships of large-scale open folds in
the mylonitic schists were inconsistent with their multilayer
Barquito, Flamenco and Pajonales characteristics and that the wavelength of these folds may reflect
Permian and Triassic intrusions do not crop out south of the thickness of underlying tabular intrusions. Those workers
Chañaral and instead plutonic complexes of an Early Jurassic arc also thought that the map pattern of the pluton was due to
were emplaced into metasedimentary basement rocks now ex- interference between east–west-trending and north- to NE-trend-
M AG M AT I C A R C FAU LT S Y S T E M S , C H I L E A N A N D E S 433

Fig. 7. Models for granitic pluton


emplacement by roof uplift. Folds with
characteristic style and position, relative to
the plutonic complex, form during
reactivation of steeply dipping, pre-existing
fractures in extension ((a) and (d)) or
contraction ((b) and (c)). The amount of
horizontal shortening or extension is
determined by the dip of the fault and the
displacement. (e) Emplacement of granitic
pluton at a dilational jog in a low-angle
detachment fault; a greater amount of heave
(horizontal extension) is implied in
comparison with the roof-uplift models

ing folds, although this idea did not fully account for irregular, S1 and S2 cleavages is to the east, and kinematic indicators in
steeply dipping margins. Previously, Grocott et al. (1994) had zones of high D2 strain (mainly asymmetric boudins) indicate
argued that synplutonic, top-to-the-east displacements in the top-east displacements during D2. Aggregates of stretched
mylonitic envelope were evidence for emplacement at dilational andalusite are partly retrogressed to chlorite but the intense D2
fault jogs during regional extension. New data from the Flamen- fabric is cut by late, undeformed granite sheets, showing
co pluton allow re-evaluation of pluton shape and emplacement unequivocally that D2 was linked to the main episode of
mechanism and we show that a top-east, ductile shear zone emplacement. Close to the contact the host rocks are migma-
formed early during emplacement and was progressively de- titic, and coarse high-temperature D2 fabrics have been
formed and recrystallized as space for the pluton was created by annealed. Evidence of the third phase of deformation is
roof uplift. present close to the western margin of the pluton at Obispito
There are four phases of deformation in the host rocks (Fig. (Fig. 3). Here, high-grade host rocks have an intense, gently
8a and b). The earliest (D1 ) is associated with the SW- to W- north-plunging linear fabric, comprising an L-tectonite fabric
vergent pre-Permian folds and cleavage (S1 ), and SW- to W- and a fold hinge-line lineation (L3 ), caused by crenulation of
transporting shear zones that characterize basement structure the S2 planar fabric during D3 (Fig. 8b). At the eastern
(Bell 1984, 1987). The second and third phases (D2, D3 ) are margin, in contrast, S2 is not crenulated and there is no
syn-emplacement, and the last phase (D4 ) is late syn-emplace- obvious evidence of D3 fabrics. Late syn-emplacement defor-
ment. In the outer part of the aureole, fresh andalusite post- mation (D4 ) is marked by conjugate sets of small C9 shear
dates S1 and represents an early contact hornfelsing of the zones (S4 ), with thin granite veins along the shear plane (Fig.
host rocks. Within 500 m of the contact this andalusite has 5f). S4 shear zones are widespread throughout the zone of
recrystallized during intense reworking of D1 fabrics to define intense D2 strain surrounding the pluton, and the conjugate
a new stretching fabric (L2 ) and a new planar fabric (S2 ; Fig. sets accommodated bulk horizontal flattening and vertical
5e). The vergence direction defined by the intersection of the stretching (Fig. 8b).
434 J. G RO C OT T & G . K . TAY L O R

Fig. 8. (a) Structure in the mylonitic envelope of western part of the Flamenco plutonic complex. (b) Model for the emplacement mechanism. Planar (S2 )
and linear (L2 ) fabric orientations are plotted stereographically in Fig. 3, vi.

vertical extension. This may reflect a ballooning component


Emplacement mechanism for the Flamenco plutonic
during emplacement, caused when the rate of magma supply
complex
exceeded the rate at which space was created by roof uplift,
We envisage the emplacement of the Flamenco pluton as resulting in ductile flattening of the host rocks in the high-
essentially a two-stage process (Fig. 8b). First, a thin tabular temperature envelope (Fig. 8b).
granite was emplaced during regional extension, softening the
host rocks at the contact and effectively introducing a ductile
detachment horizon in the upper crust. Steeply dipping exten-
sional brittle faults at higher crustal levels detached into this Stratified Triassic and Jurassic rocks
zone, in which D2 planar and linear fabrics with top-east
kinematics formed. The second stage involved inflation of the
Cifuncho, Aqua Chica and Pan de Azucar Formations
pluton with space created mainly by roof uplift. Variation in the The Cifuncho and Aqua Chica Formations of Late Triassic age
plunge of L2 (Fig. 3) is consistent with folding of an originally are exposed discontinuously along strike in the hanging wall of
NW-trending, D2 extension lineation and reworking of D1 –D2 the Tigrillo fault (Fig. 3). They comprise coarse clastic continen-
planar fabrics during roof uplift (D3 ). At the western margin, tal strata, with intermediate volcaniclastic rocks towards the base
near Obispito, crenulation of S2 during D3 is consistent with (Mercado 1980). The presence of lithic clasts derived from the
superimposition of a west-down, normal dip-slip shear strain on immediate basement, coupled with very poor sorting, implies a
an east-vergent intersection between S1 and S2 . At the eastern proximal deposition (Suárez & Bell 1992, 1994). The Cifuncho
margin, crenulation of S2 did not occur during roof uplift Formation is overlain by the Pan de Azucar Formation, which
because east-down normal dip-slip was superimposed on an east- comprises mainly shallow marine limestone with thin tuff bands
vergent intersection between S1 and S2 , and thus the planar (Davidson et al. 1976). There is no angular unconformity, but the
fabric was reinforced rather than crenulated during D3 (Fig. 8b). base of the Pan de Azucar Formation is diachronous (Naranjo &
Variation in the plunge of L2 (Fig. 3, vi) cannot be due to Puig 1984). The formations were deposited in pull-apart basins
stretching in roof-uplift shear zones alone because then it would in a NW-trending fault system that extended into central
plunge consistently down-dip on the foliation, and give a small Argentina (Dalziel et al. 1987; Suárez & Bell 1992; Simpson et
circle distribution of poles. Conjugate C9 shear zones (D4 ) are al. 2001) and basaltic, andesitic and silicic volcanic rocks within
consistent with late flattening across the earlier fabrics and the sequences have been linked to subduction.
M AG M AT I C A R C FAU LT S Y S T E M S , C H I L E A N A N D E S 435

La Negra Formation Upper Jurassic and Lower Cretaceous plutonic


complexes
The La Negra Formation crops out in the Coastal Cordillera
between Arica and Chañaral. It is several kilometres thick and Las Animas and Las Tazas
consists of andesite and basaltic andesite lavas, volcaniclastic
The Las Animas plutonic complex is dioritic and has an age of
breccias and high-level sills, with intercalations of marine and
c.160 Ma (Dallmeyer et al. 1996). It is mostly coarse grained but
continental sediments (Mpodozis & Ramos 1990). South of
becomes finer to the north, where it is often difficult to
Taltal, it overlies the Pan de Azucar Formation with erosional
distinguish from volcanic rocks of La Negra Formation. Sheets
unconformity and crops out in the hanging wall of the Tigrillo
of diorite emplaced into feldspar-phyric, amygdaloidal andesite
fault. Marine sedimentary horizons range in age from Pliensba-
show that the complex was intrusive into the formation, the lower
chian at the base near Taltal (Naranjo & Puig 1984) to Bajocian,
part of which also contains Upper Jurassic tonalite to granodior-
200 m above the base, west of Sierra Minillas (Davidson et al.
ite sills of identical radiometric age to the Las Animas complex
1976). Flamenco and Pajonales complexes have 40 Ar/39 Ar iso-
and that probably belong to the same suite (Naranjo & Puig
tope correlation ages that straddle the age of the unconformity
1984).
and imply that they are contemporary with the onset of eruption
To the south, the Las Animas complex consists mainly of
of the La Negra volcanic arc (Fig. 4). A K–Ar age of 159 !
diopside–hornblende–biotite diorite emplaced into greenschist
3 Ma for one of the Upper Jurassic sills cutting the lower part of
facies metasedimentary rocks of the arc basement. The diorite is
the sequence gives a minimum age for the formation (Naranjo &
unfoliated, but in the pelitic host rocks there is a 75 m wide,
Puig 1984).
high-temperature contact aureole with coarse-grained migmatites
and a moderate planar fabric intensity. Foliation is vertical,
trends NNE and is parallel to the contact (Fig. 3). It contains a
steeply plunging stretching fabric and down-to-the-east, S–C
Permian, Triassic and Jurassic deformation: summary fabrics. A second phase of syn-emplacement deformation is
Structures in host rocks of Permian and Triassic plutonic marked by conjugate sets of C9-type minor shear zones with
complexes point to emplacement by roof uplift permitted by granite veins. Bulk strain caused by these zones is horizontal
folding with brittle displacements at steeply dipping pluton flattening with vertical stretch. The ductile shear zone does not
margins. This implies that mainly vertical displacements accom- continue far to the north, where the host rock lithology is more
panied emplacement, and gives few clues about the arc-normal siliceous and there is less synplutonic migmatite. Here, the
component of deformation. More data on the dips of pluton contact is discordant to west-dipping, pre-emplacement quartz
margins and/or the detailed shape of the folds are required before mylonite and steeply dipping C9-type minor shear zones are the
the criteria implicit in Fig. 7 can be applied to determine whether only expression of synplutonic ductile deformation.
horizontal shortening or contraction accompanied pluton empla- The eastern margin of the Las Animas complex was probably
cement. Nevertheless, Late Triassic sequences were deposited in displaced below the erosion level during emplacement of Las
NW-trending rifts and are consistent with overriding plate Tazas complex; an elongate monzonite and granodiorite plutonic
extension and a retreating subduction boundary at this time. We complex emplaced along the western branch of the Atacama
therefore predict that margin-parallel faults and/or folds that Fault System (Fig. 3). The Las Tazas complex has an age of
permitted roof uplift during emplacement of the Permian and c.132 Ma determined by 40 Ar/39 Ar on hornblende (Dallmeyer et
Triassic plutonic complexes will prove to be extensional struc- al. 1996) and U–Pb on zircon (Berg & Breitkreuz 1983). From
tures, many of which were subsequently reworked by later fault the identical ages given by each method, we infer that cooling
displacements, particularly on the Tigrillo Fault System. Marine through the blocking temperatures was rapid, consistent with
transgression represented by the Pan de Azucar Formation may emplacement at high crustal levels (Dallmeyer et al. 1996). The
well be due to thermal subsidence following Triassic rifting complex is bounded by steeply dipping, high-temperature mylo-
(Suárez & Bell 1994). nitic shear zones (Fig. 9a) with the same 40 Ar/39 Ar cooling ages
The La Negra Formation represents an Early Jurassic volcanic as the intrusion (Wilson et al. 2000). The shear zone at the
arc (Scheuber & Reutter 1992; Lucassen & Franz 1994). western margin, in dioritic rocks of the Las Animas complex, is
Although the geochemistry is, in some respects, consistent with a characterized by early syn-emplacement, east-down, dip-slip
back-arc basin setting (Rogers & Hawkesworth 1989), it is the displacements with late syn-emplacement horizontal flattening
sedimentary sequences of the Lautaro Formation farther east that and vertical stretching on conjugate C9 surfaces (Wilson &
represent the corresponding back-arc (Mpodozis & Ramos Grocott 1999). High-temperature fabrics in the shear zone,
1990). Marine beds in the sequence show that subsidence kept anisotropy of magnetic susceptibility (AMS) and solid-state
pace with deposition, consistent with overriding plate extension fabrics in the western part of the complex are all characterized
(Scheuber & Gonzalez 1999). Space for the Lower Jurassic by steep stretching lineations and east-down, non-coaxial flow.
Flamenco plutonic complex was created mainly by roof uplift, This unequivocally demonstrates synkinematic emplacement of
which folded an early synplutonic shear zone. As the complex is the Las Tazas plutonic complex.
the same age as the lower part of the La Negra Formation, it The northern part of the complex comprises five lithological
seems possible that the synplutonic shear zone at Flamenco was units defining an east-younging, sheeted structure (Wilson 1998).
part of a regional pattern of ductile to brittle extensional faults From west to east, magmatic state lineation and the long axes of
that included contemporary basin-margin growth faults, probably the flattening-type AMS ellipsoids rotate from steep to shallow
precursors to the Tigrillo Fault System, although none of these plunge. Steeply dipping synplutonic mylonitic rocks at the east-
have yet been identified. We conclude that in Triassic to Late ern margin are characterized by shallow lineation and left-slip
Jurassic time the overriding plate was characterized by exten- S–C fabrics that overprint, very occasionally preserved, right-
sional deformation at a retreating subduction boundary and that slip kinematic indicators (Wilson 1996). This evidence that
arc plutons were emplaced by a roof uplift mechanism during strike-slip at the eastern margin and dip-slip at the western
extension. margin of the complex were essentially contemporaneous implies
436 J. G RO C OT T & G . K . TAY L O R

Fig. 9. (a) Structure of Las Tazas plutonic complex. (b) Model for the emplacement of Las Tazas pluton. Planar and linear fabrics from the marginal
shear zones are plotted stereographically in Fig. 3. Internal structure of the plutonic complex after Wilson (1998).

that by c.132 Ma a change in arc kinematics from arc-normal complex appears fault-like but is intrusive (Fig. 9a). A fracture
extension to left-oblique extension (transtension) had already zone extends away from the contact to the NW and SE, but there
taken place. are no measurable displacements across it so that it has the
The straight, NW-trending southern contact of the Las Tazas characteristics of a synplutonic fault (Fig. 6a). The western
M AG M AT I C A R C FAU LT S Y S T E M S , C H I L E A N A N D E S 437

margin of the northern pluton is cut and displaced by several (Cruden 1998), because the column of crust below the pluton,
NW-trending faults. These faults have consistent left-separations and the crystallizing magma itself, will experience a weak
along their length, extend beyond the pluton margin and have vertical shear during emplacement (Fig. 9b).
horizontal to gently NW-plunging slickenlines. They clearly post- Although the northern pluton was emplaced during extension
date emplacement and fault-bounded sectors of the contact have to left-oblique extension, AMS and magmatic state fabrics are
rotated clockwise relative to sectors without NW-trending faults steeply dipping flattening fabrics (Wilson 1998). An explanation
(Fig. 9a). This implies that vertical-axis block rotations were for flattening fabrics in the pluton, consistent with the tabular
associated with left-slip on these faults. shape now recognized, is that strain caused by left strike-slip was
progressively superimposed on fabrics formed by dip-slip defor-
mation, from west to east across the pluton. The resultant fabric
Shape and emplacement mechanism
ellipse would be of a flattening type and late progressive left-slip
We have argued that the Permian and Triassic plutons are gently deformation could rotate the linear element towards the horizon-
dipping sheets and that they have steep sides because synplutonic tal. Where they have not been overprinted, originally east-down
faults have been exploited to allow roof uplift during emplace- kinematic indicators may then give a right-slip sense viewed on
ment. The steep contacts and elongate outcrops of the Upper sections cut parallel to the strike-slip movement direction,
Jurassic and Lower Cretaceous complexes do not, therefore, rule consistent with the very localized evidence we have of this on
out gently dipping, tabular form per se. However, in contrast to the eastern margin.
all the older plutonic complexes, deformation in the host rocks
of the Las Animas and the Las Tazas complexes is characterized
Late Jurassic and Early Cretaceous deformation:
by pluton-side-down displacements.
summary
Two types of synplutonic fault are consistent with pluton-side-
down displacements: (1) a steeply dipping, floor depression fault Upper Jurassic and Lower Cretaceous plutonic complexes have
(Fig. 7a and b); (2) a releasing fault-bend in a low-angle the same age as marine sedimentary and volcaniclastic sequences
extensional fault system (Fig. 7e). Grocott et al. (1994) favoured deposited in basins inboard of the magmatic arc and interpreted
the second possibility but did not consider the alternative. Their to be back-arc basins (Fig. 4). The existence of a marine back-
model required 7–10 km of horizontal extension (heave) (for Las arc basin is prima facie evidence of a component of arc-normal
Animas) and c.5 km (for Las Tazas) to make space for the extension at this time. Evidence from dyke dilation directions
plutons, together with linkage between steep faults bounding the points to an increasing component of left-oblique extension
jog and putative low-angle extensional faults to balance the (transtension) in the magmatic arc during this period (Taylor &
section (Fig. 7e). Of course, Late Jurassic to Early Cretaceous Randall 2000). The plutonic complexes contain basaltic andesite
normal faults are present in the Coastal Cordillera, exemplified dykes with dilation directions of 2498 and 2328, for Upper
by the Tigrillo fault, but these are mainly high-angle faults. Jurassic dykes in the Vetado and Las Animas complexes,
There is no evidence for the large stratigraphic omissions on respectively, and 2058 for Lower Cretaceous dykes in the Las
low-angle faults required to accommodate horizontal extension Tazas complex. This change in the dilation direction implies
of these magnitudes in the Coastal Cordillera at the present increasing obliquity of extension from Late Jurassic into Early
erosion level, although this does not rule out the existence of a Cretaceous time and a transition from an extensional to a left-
detachment horizon at depth. Neither is it realistic to argue that oblique extensional (transtensional) regime.
space for the plutons was made by extension (heave) of up to Synplutonic deformation associated with emplacement of the
10 km on the steeply dipping faults observed, because the Lower Cretaceous plutonic complexes implies that this transten-
amount of vertical displacement (throw) required would necessi- sional regime was partitioned into strike-slip and dip-slip compo-
tate very large and abrupt changes in structural level across the nents. Evidence of deformation partitioning is particularly clear
arc, and hence in metamorphic grade, for which there is no for the Las Tazas plutonic complex, where the synplutonic shear
evidence. We therefore interpret the synplutonic shear zones at zone at the western margin is a pluton-side-down, dip-slip shear
the western margin of the complexes as high-angle ductile zone and that on the eastern side is characterized by left strike-
extensional faults that accommodated pluton emplacement by slip deformation. Other Lower Cretaceous plutons, Sierra Diecio-
floor depression and roof uplift, rather than by very large cho (c.125 Ma), La Borracha (c.106 Ma) and Remolino
amounts of horizontal extension (heave) (Fig. 9b). Core com- (c.93 Ma), all have evidence of partitioning of strike-slip and
plex-type extensional detachments reported from the Sierra de dip-slip components of the deformation, and show that a regime
Fraga in the back-arc domain are of later, post-Aptian, age of partitioned transtension characterized the margin throughout
(Mpodozis & Allmendinger 1993; Fig. 4) but could still be Early Cretaceous time. The significance of the Atacama Fault
linked kinematically to high-angle normal faulting during empla- System and the Chivato Fault System in arc kinematics is simply
cement of the late Early Cretaceous plutonic complexes between that they were synplutonic fault systems that accumulated
the Atacama Fault System and the Chivato Fault System (Fig. 2). (mainly) the strike-slip components of this partitioned transten-
sion.
Significance of magmatic state and AMS fabrics
Late-stage deformation
Steeply dipping magmatic state and AMS fabrics in the Las
Tazas pluton have been used to support the view that the The Permian and Triassic plutonic complexes and stratified
northern complex, at least, has a dyke-like form (Wilson 1998). rocks, and faults of the Tigrillo Fault System, are folded by
These fabrics certainly do not exclude a roof-uplift–floor depres- large-wavelength, low-amplitude, east–west-trending folds (Fig.
sion mechanism. Rather, steeply dipping magmatic state and 3). Similarly, the sinuous outcrop of the western contact of the
crystal plastic fabrics are to be expected, at least close to steeply Las Animas plutonic complex is, in part, caused by this type of
dipping marginal shear zones, and are predicted throughout the folding. These large, open folds are a prominent structural
intrusion by the cantilever model for pluton emplacement element between 258 and 278S (Figs. 2 and 3) and possibly
438 J. G RO C OT T & G . K . TAY L O R

farther north (Ferraris & DiBiase 1978). They plunge gently west
and are either bends or buckles formed in response to a late
deformation that post-dates all magmatic arc development and
post-Early Cretaceous inversion of the Tigrillo Fault System and
Chivato Fault System. Evidence for late fault displacements on
the Atacama Fault System is provided by almost ubiquitous
down-dip slickenlines on fault surfaces that may be related to
post-Early Cretaceous inversion, but may also be related to
Miocene to present-day extensional deformation of the upper
crust in the Coastal Cordillera.

Vertical-axis rotation in the Coastal Cordillera


Clockwise vertical-axis rotations reaching 35–458 have been
recorded by investigations of the remanent magnetism in the
Coastal Cordillera between 258S and 278S (Fig. 3). They are
clearly an important element in overriding plate deformation, but
their origin remains controversial. What is clear is that differ-
ential shortening across the Bolivian orocline cannot account for
more than c.108 of the rotations recorded at 268S (Randall et al.
1996) and that some form of block rotation related to fault
displacements is a likely basis for a solution (Forsythe &
Chisholm 1994). Taylor et al. (1998) have argued that the
rotations took place in a post-Early Cretaceous, crustal-scale
strike-slip duplex with linked margin-parallel and NW-trending
faults defining blocks that rotated clockwise during left-oblique
contraction of the Coastal Cordillera (Fig. 10a). This model has
itself proved controversial, particularly concerning the circum-
stances in which clockwise rotation of fault blocks is possible in
a left-slip regime (Beck 1998). Here, we present new evidence
for the link between vertical-axis rotation and strike-slip faulting,
and show that such rotations are an integral part of a change of
upper-plate deformation regime at the South American subduc-
tion boundary from left-oblique extension, in Permian to Early
Cretaceous time, to left-oblique contraction in Late Cretaceous
time.
Graphical models are often used to interpret palaeomagneti-
cally determined vertical-axis rotations (Garfunkel 1989; Nur et
al. 1989; Fig. 11). Characteristic of such models is that blocks Fig. 10. Major NW-trending faults in the Coastal Cordillera (258309S–
are assumed not to deform internally during rotation, and that 298009S) (after Taylor et al. 1998). Orientation of NW-trending fault sets,
faults and blocks rotate by equal amounts. Noting that strike-slip grouped according to their age, Chañaral district (258309S–268459S; area
fault arrays are often bounded by major strike-slip faults, of Fig. 3).
Garfunkel (1989) analysed block rotation in terms of area
(volume) change and length change in the boundary (Fig. 11).
He showed that, provided the blocks are free to change length in faults and to prevent voids developing, extra rotation of the faults
the dimension parallel to the boundary zone, and assuming plane and/or shortening of the rocks in the intervening fault blocks is
strain (no area change), rotation may be either in the same, or in geometrically required at the block corners. Lengthening of fault
the opposite sense, to shearing on the boundary zone, depending blocks parallel to the boundary faults during rotation may be
on the shear sense of the fault array for a given sense of shear on balanced by folding (thickening) perpendicular to the zone. In
the boundary zone. Sense of shear on the fault array will depend the Coastal Cordillera between 258S and 278S, the east–west-
on whether shortening or extension occurs across the boundary trending, late, open folds may be an expression of thickening of
zone. In effect, the fault arrays act like a single slip system in a the fault blocks during rotation.
crystal structure and rotation takes place with a sense dependent
on the longitudinal strain across the boundary zone and not the
The link between fault displacements and vertical-axis
shear sense parallel to the boundary zone (Passchier & Trouw
rotation
1996, p. 89).
Well-known space problems implicit in this approach are In the Coastal Cordillera between 258S and 278S the remanent
particularly obvious at the corners of the blocks in the boundary magnetizations affected by vertical-axis rotations are believed to
zone and are also evident in the lengthening of the deformed be both primary and secondary, and therefore similar amounts of
domain implied by the model (McKenzie & Jackson 1983; Fig. rotation have affected rocks and magnetizations of different ages
11). In nature, the undeformable block assumption must break (Randall et al. 1996; Fig. 3). Rotations are recorded by: (1)
down, at least at the corners of rotating blocks, and is often Lower Jurassic volcanic rocks of the La Negra Formation; (2)
expressed by curvature of the array faults into the boundary Upper Jurassic basaltic–andesite dykes emplaced into the Veta-
faults (Fig. 10a). To accommodate displacement on the array do, Flamenco and Las Animas plutonic complexes; (3) Lower
M AG M AT I C A R C FAU LT S Y S T E M S , C H I L E A N A N D E S 439

shear gradient (K) calculated rotation (R) observed rotation (R)


-0.27 9.13 12
-0.24 8.88 10
-0.21 6.13 9
α -0.35 12.73 11
-0.27 9.77 9.5

δ w
α Las Tazas shear gradient (K) vs rotation (delta)
s

calculated rotation
15
10

(delta)
5
0
-0.4 -0.3 -0.2 -0.1 0
(a) shear gradient (K)

Las Tazas shear gradient (K) vs rotation (delta)

measured rotation
15

(delta)
10
5
w
α
0
α+δ δ
s -0.4 -0.3 -0.2 -0.1 0
(b) shear gradient (K)
K = s/w α = initial angle Fig. 12. Fault slip, block rotation and fault orientation related to NW-
sin δ δ = rotation trending faults cutting the margin of the Las Tazas plutonic complex.
K= . sin(δ-α)
sin α Typical vertical-axis rotations (a) predicted by graphical analysis and (b)
s = displacement measured in situ are each of the order of 10–158, significantly less than
K = cot(α-δ) - cot δ
w = block width palaeomagnetically determined vertical-axis rotation (Fig. 3). Parameters
K and ! are defined in Fig. 11.
cot δ = cot(α-δ) - K K = shear gradient

Fig. 11. Graphical analysis of the relationship between the parameters of nence of late synplutonic dykes in the complex records a vertical-
displacement, block rotation and fault slip during sinistral transpression axis rotation of 35.68 ! 13.28 from a site in the Las Tazas
(modified from Garfunkel 1989). plutonic complex just east of the faulted segment of the pluton
margin (Randall et al. 1996; Fig. 3). This is larger than would be
expected from the 10–158 rotation of the foliation noted above,
Cretaceous basaltic–andesite dykes in the Las Tazas and Remo- albeit at the low end of the range results from the area as a whole.
lino plutonic complexes, which record a similar amount of The general trend of the main stratigraphic and structural
vertical-axis rotation in Lower Cretaceous dykes in plutonic elements of the Coastal Cordillera is parallel to the coast. At a
complexes on both sides of the Atacama Fault System. The smaller scale, however, lithological contacts that trend NE in
youngest dykes studied palaeomagnetically are Lower Cretaceous blocks bounded by left-slip NW-trending faults are ubiquitous
basaltic andesites that cut mylonitic rocks with a 40 Ar/39 Ar between 258S and 278S (Figs. 3 and 10a). The problem of
isotope correlation age on hornblende of c.125 Ma east of the interpretation of these NE trends as a consequence of vertical-
Atacama Fault System, but an upper age bracket for the rotations axis rotation is that it is often impossible to know the pre-
has yet to be defined. faulting trend of the contact, and hence estimate the amount of
At the western margin of the Las Tazas, Sierra Dieciocho and rotation during slip on NW faults. Nevertheless, NW fault-
Remolino plutonic complexes, intensely foliated mylonitic rocks bounded blocks with internal lithological boundaries that strike
cut by NW-trending faults systematically trend 10–158 clockwise clockwise of the overall trend of margin are consistent with the
of the average trend of the shear zones themselves, and of the hypothesis that vertical-axis rotation and left-slip on NW faults
Coastal Cordillera in general (Figs. 2, 3 and 9a). This observation are linked (Forsythe & Chisholm 1994).
implies that clockwise vertical-axis rotations of at least this To investigate the discrepancy between structural and palaeo-
magnitude have accompanied left-slip on faults in this area and is magnetic data at Las Tazas, and to provide a further test of the
supported by graphical analysis showing that displacement mag- hypothesis that rotations are linked to left-slip on NW-trending
nitudes, block width and amounts of rotation (calculated and faults, we plotted orientation of NW-trending faults according to
measured) are consistent with each other (Fig. 12). The rema- age (Fig. 10). The average trend of all NW faults is 3208, but those
440 J. G RO C OT T & G . K . TAY L O R

for which we can demonstrate unequivocally a history of displace- Conclusions


ment that goes back at least as far as Late Jurassic time make a
Deformation of the overriding plate at the Andean subduction
smaller angle to the margin than younger faults. Late Jurassic to
boundary was characterized by extension on the Tigrillo Fault
Early Cretaceous transfer faults in the Tigrillo Fault System and
System in Triassic to Early Cretaceous time with increasing left-
Early Cretaceous synplutonic and synmineralization faults trend
oblique extension (transtension) from Late Jurassic into Early
on average at 3428 (Fig. 10b, ii). Most post-Early Cretaceous
Cretaceous time. Permian to Lower Cretaceous plutonic com-
faults have a trend close to the average of 3208, and the youngest
plexes were emplaced by roof uplift–floor depression mechan-
faults in the margin, including the still active Taltal fault, trend at
isms that allowed a vertical transfer of material in the crust but
3078. The range of these average trends is 358, close to typical
this process, of itself, did not accommodate significant amounts
values of vertical-axis rotation determined palaeomagnetically.
of horizontal extension. The Atacama Fault System was initiated
during emplacement of the Las Tazas plutonic complex at
Interpretation c.132 Ma as a left strike-slip fault system. Emplacement of
plutonic complexes along the Atacama Fault System by roof
Our interpretation is that early increments of post-Early Cretac-
uplift–floor depression involving dip-slip displacements on ex-
eous displacement and up to 228 rotation (342–3208) were
tensional faults is consistent with operation of the Atacama Fault
accommodated on pre-existing faults of the Tigrillo Fault System,
System as part of a partitioned, left-oblique transtensional system
now orientated NNW and first active during Late Jurassic to
between 132 and 106 Ma. Farther east, at the Coastal Cordillera–
Early Cretaceous extensional deformation as right-slip transfer
Precordillera boundary, the Chivato Fault System was active as a
faults (Fig. 2). When these faults rotated into an unfavourable
partitioned left-oblique extensional fault system during emplace-
orientation for further slip, the main set of NW faults formed
ment of Early Cretaceous plutonic complexes between c.125 and
and, with the fault blocks they defined, subsequently rotated c.138
c.93 Ma. The Chivato Fault System may also have had an earlier
(320–3078) to their present orientation with progressive arc-
history of extension that would correlate with displacements on
normal shortening. This implies that NW-trending faults and fault
the Tigrillo Fault System. This style of upper-plate deformation
blocks are redefined incrementally and that the amount of
is consistent with the existence of a retreating subduction at the
vertical-axis rotation recorded palaeomagnetically may be up to
Andean margin from Triassic time to the end of Early Cretaceous
228 more than the rotation experienced by an adjacent fault
time.
belonging to the 3208 set as, for example, at Las Tazas. NW
Subsequent deformation in the Coastal Cordillera was parti-
faults that make a high angle to the margin do so because they
tioned into arc-parallel left strike-slip and arc-normal shortening
are relatively young and have not rotated significantly. The
components. Vertical-axis clockwise rotations of 35–458 were
orientation of the youngest faults may simply be controlled by
produced by a combination of left-slip on NW-trending faults
the failure angle in the overriding plate, although their persis-
and left-slip on north–south-trending margin-parallel faults and
tence through time suggests that deep-seated, NW-trending
accommodated by domino-style rotation of both NW faults and
structures repropagate periodically through the upper crust to re-
fault blocks. Whereas some NW-trending faults were initiated
establish a set of faults at a relatively high angle to the margin.
during this contractional deformation, others were reactivated
lateral ramps in the Tigrillo Fault System. Arc-normal shortening
Relation of clockwise rotations to left-slip margin-parallel was accommodated by: (1) domino-style fault and fault block
shear sense rotation with concomitant arc-parallel extension; (2) transpres-
sion on north–south-trending fault zones, particularly the Chiva-
In the models of Garfunkel (1989), block rotations are linked to
to Fault System at the Coastal Cordillera–Precordillera
displacements on boundary fault systems. In the Coastal Cordil-
boundary; (3) thin-skinned folding and thrusting in Triassic to
lera, the youngest component of strike-slip diplacements on NW
Lower Cretaceous stratified rocks. Partitioned oblique contraction
faults post-date displacement on the Atacama Fault System. Some
of the upper plate is consistent with operation of the Andean
kinematic linkage between NW faults and the Atacama Fault
convergent margin as an advancing subduction boundary in post-
System has been demonstrated at Manto Verde (Bonson 1998)
Early Cretaceous time.
but there is no evidence of major linkage between NW faults and
the Atacama Fault System, consistent with similar magnitudes of We would like to thank colleagues at the Geological Survey of Chile
rotation recorded on either side of the fault system. The Atacama (SERNAGEOMIN) for their contributions, often through vigorous debate,
Fault System cannot, therefore, be regarded as a boundary fault to the ideas expressed in this paper, especially C. Arévalo, P. Cornejo, E.
to the NW fault arrays in the sense of Garfunkel (1989). Godoy, S. Matthews and A. Tomlinson. We also thank C. Bonson, M.
Traced east from the Atacama Fault System, NW faults meet Brown, M. Gardeweg, M. Gipson, D. Leech, C. Mpodozis, D. Randall, P.
the Chivato Fault System–Chañarcillo thrust belt at the Coastal Treloar and J. Wilson for discussion. SERNAGEOMIN provided us with
Cordillera–Pre-Cordillera boundary (Taylor et al. 1998; Arévalo first-class logistical support without which our fieldwork would have been
1999; Fig. 2). This system is characterized by left-oblique much more difficult. In particular, we are indebted to R. Morales for all
contraction partitioned into arc-normal thrusting and arc-parallel his help over many field seasons. S. Bignold (K.U.) compiled the maps
left strike-slip (Arévalo 1999). Significantly, NW faults acted as and drafted the other figures, and A. Belsham (K.U. Media Services)
transfer faults in the thrust belt, just as they do in the stratified processed the photographic images. Both are thanked for their help. The
sequences of the Coastal Cordillera. The combination of NW- research was funded by NERC research grant GR3/99/111 awarded to
G.K.T. and J.G., and a Royal Society Study Visit bursary awarded to J.G.
trending left-slip faults, clockwise vertical-axis rotation and
margin-parallel left-slip faults of the Chivato Fault System is
fully consistent with models of Garfunkel (1989), where sense of References
block rotation is opposite to the sense of shear on the boundary
Arabasz, W.J., Jr. 1971. Geological and geophysical studies of the Atacama fault
faults. This system has accommodated left-oblique contraction of zone in northern Chile. PhD thesis, California Institute of Technology,
the margin during post-Early Cretaceous time, although an upper Pasadena.
age bracket for this deformation has yet to be defined. Arévalo, C. 1994. Carta GeologÍca de Chile, Hoja Los Loros, escala 1:100,000.
M AG M AT I C A R C FAU LT S Y S T E M S , C H I L E A N A N D E S 441

Servicio Nacional de Geologı́a y Minerı́a, Santiago de Chile. plutonismo asociado al sistema de falla Atacama. VII Congreso Geológico
Arévalo, C. 1999. The Coastal Cordillera/Precordillera boundary in the Tierra Chileno, Universidad Católica del Norte, Antofagasta, Actas, 3, 1604–1607.
Amarilla area (278209–278409S; 708059–708209W) northern Chile, and the Grocott, J., Brown, M., Dallmeyer, R.D., Taylor, G.K. & Treloar, P.J.
structural setting of the Candelaria Cu–Au ore deposit. PhD thesis, 1994. Mechanisms of continental growth in extensional arcs: an example
University of Kingston, Kingston-upon-Thames. from the Andean plate-boundary zone. Geology, 22, 391–394.
Bahlburg, H. & Breitkreuz, C. 1993. Differential response of a Devonian– Grocott, J. & Wilson, J. 1997. Ascent and emplacement of granitic plutonic
Carboniferous platform–deeper basin system to sea-level change and complexes in subduction-related extensional environments. In: Holness, M.
tectonics, N. Chilean Andes. Basin Research, 5, 21–40. (ed.) Deformation-enhanced Fluid Transport in the Earth’s Crust and Mantle.
Beck, M.E. Jr 1998. On the mechanism of crustal block rotations in the central Mineralogical Society Series, 7, 173–195.
Andes. Tectonophysics, 299, 75–92. Lara, L. & Godoy, E. 1998. Carta GeologÍca de Chile, Hoja Quebrada Salitrosa,
Bell, C.M. 1984. Deformation produced by subduction of a Palaeozoic turbidite No. 4, escala 1:100,000. Servicio Nacional de GeologÍa y MinerÍa, Santiago
sequence in northern Chile. Journal of the Geological Society, London, 141, de Chile.
339–347. Lucassen, F. & Franz, G. 1994. Arc-related Jurassic igneous and metamorphic
Bell, C.M. 1987. The origin of the Upper Palaeozoic Chañaral melange of rocks in the Coastal Cordillera of northern Chile, Region Antofagasta.
northern Chile. Journal of the Geological Society, London, 144, 599–610. Journal of Metamorphic Geology, 14, 249–265.
Berg, K. & Baumann, A. 1985. Plutonic and metasedimentary rocks from the Marschik, R. & Fontboté, L. 1996. Copper (–iron) mineralization and super-
Coastal Range of northern Chile: Rb–Sr and U–Pb isotopic systematics. imposition of alteration events in the Punta del Cobre belt, northern Chile. In:
Earth and Planetary Science Letters, 75, 101–115. Camus, F., Sillitoe, R.H. & Peterson, R. (eds) Andean Copper Deposits:
Berg, K. & Breitkreuz, C. 1983. Mesozoische Plutone in der nordchilenischen New Discoveries, Mineralization, Styles, and Metallogeny. Society of
Küstenkordillere: petrogenese, geochronologie, Geochemie und Geodynamik Economic Geologists Special Publication, 5, 171–190.
mantelbetonter Magmatite. Geotektonische Forschungen, 66, 1–107. McCaffrey, K.J.W. & Petford, N. 1997. Are granites scale invariant? Journal of
Berg, K., Breitkreuz, C., Damm, K.-W., Pichowiak, S. & Zeil, W. 1983. The the Geological Society, London, 154, 1–4.
North-Chilean Coast Range—an example for the development of an active McKenzie, D. & Jackson, J. 1983. The relationship between strain rates, crustal
continental margin. Geologische Rundshau, 72, 715–731. thickening, palaeomagnetism, finite strain and fault movements within a
Bonson, C. 1998. Fracturing, fluid processes and mineralisation in the Cretaceous deforming zone. Earth and Planetary Science Letters, 65, 182–202.
continental magmatic arc of northern Chile (258159–278159S). PhD thesis, Mercado, M. 1978. Carta Geológica de Chile, Cordillera de la Costa entre
University of Kingston, Kingston-upon-Thames. Chañaral y Caldera, No. 27, escala 1:100,000. Instituto de Investigaciónes
Brook, M., Pankhurst, R.J., Shepherd, T.J., Spiro, B., Snelling, N.J. & Geológica de Chile, Santiago.
Swainbank, I.G. 1986. Andean Geochronology and Metallogenesis. British Mercado, M. 1980. Carta Geológica de Chile, Area Pan de Azúcar, No. 37, escala
Geological Survey Open-File Report WI/IG/86/1. 1:100,000. Instituto de Investigaciónes Geológica de Chile, Santiago.
Brown, M. 1991. Comparative geochemical interpretation of Permian–Triassic Mpodozis, C. & Allmendinger, R.W. 1993. Extensional tectonics: Cretaceous
plutonic complexes of the Coastal Range and Altiplano (258309–268309S), Andes, northern Chile (278S). Geological Society of America Bulletin, 105,
northern Chile. Geological Society of America, Special Paper, 265, 157–177. 1462–1177.
Brown, M., Diaz, F. & Grocott, J. 1993. Displacement history of the Atacama Mpodozis, C. & Ramos, V.A. 1990. The Andes of Chile and Argentina. In:
Fault System 258009–278009S, northern Chile. Geology Society of America Eriksen, G.E., Cañas-Pinochet, M.T. & Reinemund, J.A. (eds) Geology of
Bulletin, 105, 1165–1174. the Andes and its relationship to hydrocarbon and mineral resources.
Colley, H., Treloar, P.J. & Diaz, F. 1988. Gold and silver mineralization in the Circum-Pacific Council for Energy and Mineral Resources, Earth Science
El Salvador region, northern Chile. In: Keays, R., Ramsey, R. & Groves, D. Series, 11, 59–90.
(eds) The Geology of Gold Deposits: the Perspective in 1988. Economic Naranjo, J.A. 1978. Carta Geológica de Chile, Hoja Geologı́a de la zona interior
Geology Monograph, 6, 208–217. de la Cordillera de la Costa entre los 268009 y 268209S, No. 34, escala
Cruden, A.R. 1998. On the emplacement of tabular granites. Journal of the 1:100,000. Instituto de Investigaciónes Geológica de Chile, Santiago.
Geological Society, London, 155, 853–862. Naranjo, J.A. & Puig, A. 1984. Carta Geologı́ca de Chile, Hojas Taltal y
Dallmeyer, R.D., Brown, M., Grocott, J., Taylor, G.K. & Treloar, P.J.T. Chañaral, No. 62-63, escala 1:250,000. Servicio Nacional de GeologÍa y
1996. Mesozoic magmatic and tectonic events within the Andean plate Minerı́a, Santiago.
boundary zone, 268–278309S, north Chile: constraints from 40 Ar/39 Ar mineral Naranjo, J.A. 1987. Interpretación de la actividad Cenozoica superior a lo largo
ages. Journal of Geology, 104, 19–40. de la zona de falla Atacama, norte de Chile. Revista Geológica de Chile, 31,
Dalziel, I.W.D., Storey, B.C., Garrett, S.W., Grunow, A.M., Herrod, L.D.B. 43–55.
& Pankhurst, R.J. 1987. Extensional tectonics and fragmentation of Nur, A., Ron, H. & Scotti, O. 1989. Mechanics of distributed fault and block
Gondwanaland. In: Dewey, J.F., Coward, M.P. & Hancock, P. (eds) rotation. In: Kissel, C. & Laj, C. (eds) Paleomagnetic Rotations and
Continental Extensional Tectonics. Geological Society, London, Special Continental Deformation. NATO ASI Series, 254, 209–228.
Publications, 28, 433–441. Passchier, C.W. & Trouw, R.A.J. 1996. Microtectonics. Springer, Berlin.
Damm, K.W. & Pichowiak, S. 1981. Geodynamik und Magmengenese in der Randall, D.E., Taylor, G.K. & Grocott, J. 1996. Major crustal rotations in the
Küstenkordillere Nordchiles zwischen Taltal und Chañaral. Geotekonische Andean margin: paleomagnetic results from the Coastal Cordillera of north-
Forschungen, 61, 1–166. ern Chile. Journal of Geophysical Research, 101, 15783–15798.
Davidson, J., Godoy, E. & Covacevich, V. 1976. El Bajociano de Sierra Minillas Rogers, G. & Hawkesworth, C.J. 1989. A geochemical traverse across the north
(708309O; 268009S) y Sierra Fraga (698509O; 278009S), Provincia de Atacama, Chilean Andes: evidence for crust generation from the mantle wedge. Earth
Chile: edad y marco geotectónico de la Formación La Negra en esta latitud. I and Planetary Science Letters, 91, 271–285.
Congreso Geológico Chileno, Universidad de Chile, Actas, 1, 255–272. Royden, L.H. 1993. The tectonic expression of slab-pull at continental convergent
Dewey, J.F. 1980. Episodicity, sequence and style at convergent plate boundaries. boundaries. Tectonics, 12, 303–325.
In: Strangeway, D.W. (ed.) The Continental Crust and its Mineral Russo, R.M. & Silver, P.G. 1996. Cordillera formation, mantle dynamics and the
Deposits. Geological Association of Canada, Special Publication, 20, Wilson Cycle. Geology, 24, 511–514.
553–576. Segerstrom, K. & Ruiz, C. 1962. Carta Geológica de Chile, Hoja Copiapó, No.
Ferraris, F.B. & DiBiase, F.F. 1978. Carta Geologı́ca de Chile, Hojas 6, escala 1:100,000. Instituto de Investigaciónes Geológicas de Chile,
Antofagasta, No. 30, escala 1:250,000. Servicio Nacional de Geologı́a y Santiago.
Minerı́a, Santiago. St. Amand, P. & Allen, C.R. 1960. Strike-slip faulting in northern Chile.
Forsythe, R. & Chisholm, L. 1994. Palaeomagnetic and structural constraints on Geological Society of America Bulletin, 71, 1965.
rotations in the north Chilean coast ranges. Journal of South American Earth Scheuber, E. & González, G. 1999. Tectonics of the Jurassic–Early Cretaceous
Sciences, 7, 279–294. magmatic arc of the north Chilean Coastal Cordillera (228–268S): a story of
Garfunkel, Z. 1989. Regional deformation by block translation and rotation. In: crustal deformation along a convergent plate boundary. Tectonics, 18,
Kissel, C. & Laj, C. (eds) Paleomagnetic Rotations and Continental 895–910.
Deformation. NATO ASI Series, 254, 181–208. Scheuber, E. & Reutter, K.J. 1992. Magmatic arc tectonics in the Central Andes
Gipson, M. 2002. Palaeomagnetism and geochronology of the Andean forearc in between 218 and 258S. Tectonophysics, 205, 127–140.
the Vallenar district (298S), Region III, northern Chile. PhD thesis, University Simpson, C., Whitmeyer, S., De Paor, D.G., Gromet, R., Miro, R., Krol, M.A.
of Plymouth. & Short, H. 2001. Sequential ductile through brittle reactivation of major
Godoy, E. & Lara, L. 1998. Carta Geologı́ca de Chile, Hojas Chañaral y Diego fault zones along the accretionary margin of Gondwana in central Argentina.
del Almagro, No. 5-6, escala 1:100,000. Servicio Nacional de GeologÍa y In: Holdsworth, R.E., Strachan, R.A., Magloughlin, J. & Knipe, R.J.
MinerÍa, Santiago de Chile. (eds) The Nature and Tectonic Significance of Fault Zone Weakening.
Godoy, E., Lara, L. & Ugalde, H. 1997. La falla Chivato: borde oriental del Geological Society, London, Special Publications, 186, 233–255.
442 J. G RO C OT T & G . K . TAY L O R

Suárez, M. & Bell, C.M. 1992. Triassic rift-related sedimentary basins in related to the Atacama Fault Zone. In: Camus, F., Sillitoe, R.H. &
northern Chile (248–298S). Journal of South American Earth Sciences, 6, Petersen, R. (eds) Andean Copper Deposits: New Discoveries, Mineraliza-
109–121. tion Styles and Metallogeny. Society of Economic Geologists Special
Suárez, M. & Bell, C.M. 1994. Braided rivers, lakes and sabkahs of the Upper Publication, 5, 157–170.
Triassic Cifuncho Formation, Atacama Region, Chile. Journal of South Waschbusch, P. & Beaumont, C. 1996. Effect of a retreating subduction zone on
American Earth Sciences, 7, 25–33. deformation in simple regions of plate convergence. Journal of Geophysical
Sylvester, H. & Palacios, C. 1992. Transpressional structures in the Research—Solid Earth, 101(B12), 28133–28148.
Andes between the Atacama Fault Zone and the West Fissure System at Wilson, J. 1996. Emplacement of the Las Tazas plutonic complex, Coastal
278S, III Region, Chile. Zentralblatt für Geologie und Paläontologie, 6, Cordillera, northern Chile. PhD thesis, University of Kingston, Kingston-
1645–1658. upon-Thames.
Taylor, G.K.T. & Randall, D.E. 2000. Structural analysis of dyke emplacement Wilson, J. 1998. Magnetic susceptibility patterns in a Cordilleran granitoid: the
directions as an aid to palaeomagnetic studies: an example from northern Las Tazas complex, northern Chile. Journal of Geophysical Research, 103,
Chile. Geophysical Journal International, 141, 253–258. 5257–5267.
Taylor, G.K., Grocott, J., Pope, A. & Randall, D.E. 1998. Mesozoic fault Wilson, J. & Grocott, J. 1999. The emplacement of the Las Tazas complex,
systems, deformation and fault block rotation in the Andean forearc: a crustal northern Chile: the relationship between local and regional strain. Journal of
scale strike-slip duplex in the Coastal Cordillera of northern Chile. Structural Geology, 21, 1513–1523.
Tectonophysics, 299, 93–109. Wilson, J., Dallmeyer, R.D. & Grocott, J. 2000. New 40 Ar/39 Ar dates from the
Vila, T., Lindsay, N. & Zamora, R. 1997. Geology of the Manto Verde copper Las Tazas complex, northern Chile: tectonic significance. Journal of South
deposit, northern Chile: a specularite-rich, hydrothermal–tectonic breccia American Earth Science, 13, 115–122.

Received 12 September 2001; revised typescript accepted 31 January 2002.


Scientific edity by Rob Strachan

You might also like