Cenozoico Transpresional Nazca

Embed Size (px)

Citation preview

  • 7/27/2019 Cenozoico Transpresional Nazca

    1/26

    Late Cenozoic transpressional ductile deformation north of

    the NazcaSouth AmericaAntarctica triple junction

    Jose Cembrano a,*, Alain Lavenu b,c, Peter Reynolds d, Gloria Arancibia c,Gloria Lopez c, Alejandro Sanhueza c

    aDepartamento de Ciencias Geologicas, Universidad Catolica del Norte, Avda Angamos 0610, Antofagasta, Chileb

    Institut Francais de Recherche Scientifique pour le Developpement en Cooperation (IRD),Casilla 53390 Correo Central, Santiago 1, Chile

    cDepartamento de Geologa, Universidad de Chile, Casilla 13518, Correo 21, Santiago, ChiledDepartment of Earth Sciences, Dalhousie University, Halifax, Nova Scotia, Canada B3H 3J5

    Received 23 January 2001; accepted 26 June 2002

    Abstract

    The southern Andes plate boundary zone records a protracted history of bulk transpressional deformation during the

    Cenozoic, which has been causally related to either oblique subduction or ridge collision. However, few structural and

    chronological studies of regional deformation are available to support one hypothesis or the other. We address along- and

    across-strike variations in the nature and timing of plate boundary deformation to better understand the Cenozoic tectonics ofthe southern Andes.

    Two east west structural transects were mapped at Puyuhuapi and Aysen, immediately north of the Nazca South

    AmericaAntarctica triple junction. At Puyuhuapi (44jS), northsouth striking, high-angle contractional and strike-slip ductile

    shear zones developed from plutons coexist with moderately dipping dextral-oblique shear zones in the wallrocks. In Ayse n

    (4546j), top to the southwest, oblique thrusting predominates to the west of the Cenozoic magmatic arc, whereas dextral

    strike-slip shear zones develop within it.

    New 40Ar 39Ar data from mylonites and undeformed rocks from the two transects suggest that dextral strike-slip, oblique-

    slip and contractional deformation occurred at nearly the same time but within different structural domains along and across the

    orogen. Similar ages were obtained on both high strain pelitic schists with dextral strike-slip kinematics (4.4F0.3 Ma, laser on

    muscovitebiotite aggregates, Aysen transect, 45jS) and on mylonitic plutonic rocks with contractional deformation (3.8F0.2

    to 4.2F0.2 Ma, fine-grained, recrystallized biotite, Puyuhuapi transect). Oblique-slip, dextral reverse kinematics of uncertain

    age is documented at the Canal Costa shear zone (45j

    S) and at the Queulat shear zone at 44j

    S. Published dates for theundeformed protholiths suggest both shear zones are likely Late Miocene or Pliocene, coeval with contractional and strike-slip

    shear zones farther north. Coeval strike-slip, oblique-slip and contractional deformation on ductile shear zones of the southern

    Andes suggest different degrees of along- and across-strike deformation partitioning of bulk transpressional deformation.

    The long-term dextral transpressional regime appears to be driven by oblique subduction. The short-term deformation is in

    turn controlled by ridge collision from 6 Ma to present day. This is indicated by most deformation ages and by a southward

    increase in the contractional component of deformation. Oblique-slip to contractional shear zones at both western and eastern

    0040-1951/02/$ - see front matterD 2002 Elsevier Science B.V. All rights reserved.P I I : S 0 0 4 0 - 1 9 5 1 ( 0 2 ) 0 0 3 8 8 - 8

    * Corresponding author. Fax: +56-55-355977.

    E-mail address: [email protected] (J. Cembrano).

    www.elsevier.com/locate/tecto

    Tectonophysics 354 (2002) 289 314

  • 7/27/2019 Cenozoico Transpresional Nazca

    2/26

    margins of the Miocene belt of the Patagonian batholith define a large-scale pop-up structure by which deeper levels of the crust

    have been differentially exhumed since the Pliocene at a rate in excess of 1.7 mm/year.

    D 2002 Elsevier Science B.V. All rights reserved.

    Keywords: Ductile deformation; Late Cenozoic; Nazca South America Antarctica triple junction; Transpression; Liquine Ofqui fault zone

    1. Introduction

    Overall transpressional deformation is expected at

    continental margins where the convergence vector is

    oblique with respect to the plate boundary zone (e.g.

    Sanderson and Marchini, 1984; Dewey et al., 1998;

    Fossen and Tikoff, 1998). Other first-order factors

    controlling the tectonics of convergent margins are the

    age of the subducting plate, the nature and thermal

    structure of the overriding plate, the existence of

    major trench-parallel faults and ridge subduction

    (e.g. Fitch, 1972; Jordan et al., 1983; Jarrard, 1986;

    Beck, 1991; Nelson et al., 1994).

    Kinematic models show that transpressional defor-

    mation arising from oblique convergence is accom-

    plished by distinctive structural styles along and across

    different plate boundaries, mostly depending on the

    angle of obliquity, defined as the angle between the

    convergence vector and the normal to the trench (e.g.Jarrard, 1986; McCaffrey, 1992). For small angles of

    obliquity, transpression is homogeneously distributed

    as in the case of the AustralianPacific plate boundary

    in New Zealand. For large angles of obliquity, complete

    partition of transpression is expected. This is the case of

    the Pacific North America plate boundary of the west-

    ern US, where the San Andreas Fault accommodates

    most of the simple shear component (Teyssier et al.,

    1995). The general case, however, will be that of

    heterogeneous transpression in which discrete domains

    across the plate boundary accommodate wrench-domi-nated or pure-shear dominated transpression (e.g.

    Fitch, 1972; Fossen et al., 1994; Tikoff and Greene,

    1997). Bulk transtension has also been reported at

    convergent plate margins, especially at those that were

    actively retreating (e.g. Grocott et al., 1994).

    In the case of ductile shear zones at obliquely

    convergent plate margins, the strike-slip component

    of transpression is not taken up by slip on a single

    fault but within zones of distributed shear (e.g. Fossen

    et al., 1994; Jones and Tanner, 1995).

    However, the nature and degree of deformation

    partitioning will not only depend on the angle of

    obliquity. For instance, thermally weak intra-arc

    shear zones can accommodate a significant part of

    the bulk transpressional deformation arising from

    oblique convergence affecting the predictions of

    kinematic models (e.g. Saint Blanquat et al., 1998).

    The southern Chilean Andes provides a natural

    laboratory to investigate the nature of long-term

    and short-term transpressional deformation across

    an obliquely convergent continental margin. A major

    problem when trying to establish the nature of the

    relationship between the tectonics of the overriding

    South American plate and the Cenozoic plate kine-

    matic framework has been the almost complete lack

    of detailed structural and thermochronological stud-

    ies in the northern Patagonian region. With the

    exception of regional geology and petrology-oriented

    work (e.g. Bobenrieth et al., 1983; Bartholomew andTarney, 1984; Herve et al., 1995), there has been no

    systematic study of the tectonic evolution of this

    segment of the Andes. The regional geology and

    structure of the foreland, however, has been recently

    studied (Flint et al., 1994; Suarez and De La Cruz,

    2001). One important feature of this segment of the

    Andes is the absence of a Meso-Cenozoic regional

    fold and thrust belt inland of a 1000-km-long Cen-

    ozoic intra-arc fault zone, which has been assumed

    to accommodate much of the convergent deforma-

    tion.Field studies have been difficult to conduct in the

    western slope of the Patagonian Cordillera because

    of dense vegetation and generally poor weather

    conditions. However, a large network of valleys

    and fjords cuts the region, allowing access by boat

    and basic coastal outcrop mapping to be carried out.

    Furthermore, it is possible to study several eastwest

    transects across the orogen by traveling along certain

    fjords and channels. In this paper, we present and

    discuss field, structural and thermochronological data

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314290

  • 7/27/2019 Cenozoico Transpresional Nazca

    3/26

    from two transects across the southern Andes, at

    latitudes 44j and 46j, the Puyuhuapi and Aysen

    transect, respectively (Figs. 1 and 2). Previous work

    by Schermer et al. (1995, 1996) and Cembrano et al.

    (2000) presented a summary of along-strike contrasts

    in the nature and timing of deformation along the

    segment immediately to the north, from 39jS to

    43j

    S.

    Fig. 1. Regional-scale tectonic setting of the Southern Andes. Northeast-trending, trench-parallel lineaments correspond to the Liquin eOfqui

    fault zone (LOFZ). Current position of the NazcaSouth AmericaAntarctica triple junction is shown. S eries of small arrows represent

    convergence vectors between the Nazca and South American plates for the last 48 Ma. (Modified from Pardo-Casas and Molnar, 1987;

    Cembrano et al., 1996; Bourgois et al., 1996; Somoza, 1998 ). Boxes show transect locations.

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 291

  • 7/27/2019 Cenozoico Transpresional Nazca

    4/26

    Fig. 2. Regional-scale geologic map of the southern Chilean Andes between 43jS and 47jS. (modified from SERNAGEOMIN, 1980;

    Bobenrieth et al., 1983; Forsythe and Nelson, 1985; Pankhurst and Herve, 1994; Herve et al., 1995; Pankhurst et al., 1999).

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314292

  • 7/27/2019 Cenozoico Transpresional Nazca

    5/26

    1.1. Tectonic and geologic setting

    An active spreading center, the Chile Ridge, is

    currently subducting under western South Americaand a trench-parallel fault system, the LiquineOfqui

    fault zone (LOFZ), has developed within the Ceno-

    zoic magmatic arc (Fig. 1). The Cenozoic geodynamic

    setting of the southern Chilean Andes is well con-

    strained, showing relatively steady right-lateral obli-

    que subduction of the Farallon (Nazca) plate beneath

    South America since 48 Ma with the exception of

    nearly orthogonal convergence from 26 to 20 Ma

    (Pardo-Casas and Molnar, 1987; Somoza, 1998). At

    present, the angle of obliquity of Nazca South Amer-

    ica plate convergence vector with respect to the

    orthogonal to the trench is f26j for southern Chile

    (Jarrard, 1986). The slab dip is approximately 16j

    (Jarrard, 1986) and the age of the subducting Nazca

    plate decreases from f25 Ma at 38jS to virtually 0

    Ma at 46jS, where the Chile Ridge is currently

    subducting (Herron et al., 1981; Cande and Leslie,

    1986; Bourgois et al., 1996). The limited available

    seismic data suggest that the Chilean forearc between

    39jS and 46jS is currently undergoing trench-orthog-

    onal shortening whereas the volcanic arc is absorbing

    a small trench-parallel component (Chinn and Isacks,

    1983; Cifuentes, 1989; Barrientos and Acevedo,1992; Dewey and Lamb, 1992; Murdie, 1994).

    The North Patagonian Batholith (NPB) occupies a

    central belt, 1000-km long and 200-km wide, within

    the Patagonian Cordillera. The NPB marks the axis of a

    the relatively low-altitude (1 2 km) Main Range

    flanked to the west by the southward continuation

    of the Central Depression or Longitudinal Valley of

    Central Chile (e.g. Herve, 1994; Lavenu and Cem-

    brano, 1999). The Central Depression is a north south

    trending belt characterized by numerous islands and

    fjords composed of basement metamorphic rocks anda younger unit of patchy metasediments and pillow

    metabasalts that lie closer to the NPB. The latter unit,

    named Traiguen Formation, is generally flat-lying west

    of a major northsouth-trending lineament (Canal de

    Moraleda) and highly deformed at and to the east of

    the channel where they are rare (Figs. 1 and 2).

    The basement is mostly composed of metasedi-

    mentary rock, greenstone and chert. Godoy et al.

    (1984) suggested that the basement was Late Paleozoic

    in age based upon field and geochronologic reconnais-

    sance. Herve (1988) interpreted the metamorphic base-

    ment as an accretionary prism, deposited and deformed

    in situ during Late Paleozoic times although much

    younger ages (Jurassic) have been obtained recently(Herve, 1998). Basement rocks are locally deformed

    along northeast-striking ductile shear zones, which are

    meters to hundreds of meters in width.

    Based upon regional mapping and basic geochro-

    nologic dating, the North Patagonian Batholith has

    previously been divided into three roughly defined,

    orogen-parallel belts: A western Jurassic to Creta-

    ceous granodioritic belt, a central Miocene belt of

    granodioritediorite with few granites, and an eastern

    mid-Cretaceous belt constituted by monzogranites and

    granites (Pankhurst et al., 1992, 1999; Herve et al.,

    1993; Pankhurst and Herve, 1994), (Fig. 2).

    Bartholomew and Tarney (1984) proposed that the

    northsouth trending strip of volcanoclastic and sedi-

    mentary rock (Traiguen Formation) located between

    the Paleozoic basement and the main outcrop of the

    Patagonian Batholith represented a Late Cretaceous

    Tertiary intra-arc basin. According to their model,

    crustal thinning arising from east west extension

    was sufficient for true oceanic floor to develop. Their

    model is supported mainly by the geochemical signa-

    ture of the volcanic rocks: tholeiitic rather than calc-

    alkaline in character. Field evidence documenting thegeometric and kinematic relationship between the

    intra-arc basin deposits and the batholith is very

    limited. They proposed that, during a Miocene com-

    pressional event, the basin was inverted and under-

    thrust below the plutonic complex. Easterly dipping

    shear zones outcropping along the margins of the basin

    were presented as evidence for underthrusting. The

    fact that plutonic and metamorphic rocks representing

    lower crustal levels crop out to the east of Canal de

    Moraleda was used as evidence for west-verging

    regional thrust faults (Bartholomew and Tarney, 1984).Herve et al. (1995) proposed an alternative model

    for the origin of the pillow basalts and associated

    metasediments. Based on geologic mapping of the

    Traiguen Formation at Isla Magdalena, they suggested

    that a pull-apart basin nucleated along a leaky trans-

    form fault, represented now by the LOFZ. However,

    as in the former model, little structural and chrono-

    logical evidence was provided to relate coeval dextral-

    strike-slip motion on the fault to basin formation and

    sedimentation. A major difficulty has been to date the

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 293

  • 7/27/2019 Cenozoico Transpresional Nazca

    6/26

    metabasalts and metasediments, which for the most

    part, have yielded poorly constrained Miocene ages

    (Herve et al., 1995).

    1.2. Structural setting

    Cenozoic continental margin deformation along the

    southern Andes is mostly restricted to the magmatic arc,

    where a major fault-system, the LOFZ runs for more

    than a 1000 km in a north south direction (e.g. Herve et

    al., 1979; Forsythe and Nelson, 1985; Dewey and

    Lamb, 1992; Herve, 1994; Cembrano et al., 1996).

    The forearc and foreland regions show very little evi-

    dence of regional-scale deformation. The region under

    study does not have a foreland fold-and-thrust belt as do

    the Andean regions immediately south and north (e.g.

    Ramos, 1989). The development of the Patagonian

    foreland fold-and-thrust belt south of 47jS is attributed

    to the northward migration of the subducting Chile-

    Ridge over the last 14 Ma (Ramos and Kay, 1992).

    According to Herve (1994), the absence of foreland

    deformation north of 47jS hasresulted from the accom-

    modation of a significant part of the convergent defor-

    mation along the Liquine Ofqui fault zone.

    2. Geometry and kinematics of discrete ductileshear zones of the Puyuhuapi and Aysen transects

    Detailed structural work was performed in several

    high strain zones developed from rocks of different

    nature and age along and across the magmatic arc in

    two studied transects. Some of these shear zones are

    located along the eastern lineament of the LOFZ,

    whereas others occur along the western lineament.

    Other, less-well exposed shear zones occur between

    the two main crustal lineaments (Figs. 1 4). Bartho-

    lomew and Tarney (1984) previously identified someof these shear zones at the regional scale, during

    reconnaissance work. However, they provided no data

    on the geometry of stretching lineations or the kine-

    matics of deformation.

    For each shear zone, we analyzed ductile structures

    in mylonitic rocks on the mesoscopic and microscopic

    scales to determine the kinematics and approximate

    conditions of deformation. Observed kinematic indi-

    cators include C/S fabrics (e.g. Berthe et al., 1979),

    asymmetric porphyroclast systems (e.g. Passchier andSimpson, 1986), mica fish (e.g. Lister and Snoke,

    1984), and domino structures (e.g. Simpson and

    Schmid, 1983). Quartz ribbon microstructure (e.g.

    Simpson, 1985) and feldspar microstructure (Simpson,

    1985; Tullis and Yund, 1987; Fitz-Gerald and Stunitz,

    1993) were documented to provide a rough estimate of

    deformation temperatures. Deformation microstruc-

    tures were combined, when possible, with co-existing

    metamorphic mineral assemblages to better constrain

    temperature conditions during deformation.

    Figs. 3 and 4 show the spatial distribution and

    geometry of solid-state fabrics of the Andean region

    between 44j and 46j south along with stereoplots of

    foliations and lineations. The regional structural map

    and the stereoplots show variable dips of foliations

    and shallowly to steeply plunging lineations through-

    out. Most foliations dip to the east along the western

    margin of the Miocene plutonic belt of the North

    Patagonian Batholith whereas they dip to west along

    its eastern margin.

    2.1. Puyuhuapi quarry shear zone

    This shear zone deforms mingled diorite and gran-

    odiorite belonging to a 16.7 Ma pluton (RbSr, whole

    rock; Pankhurst and Herve, 1994) (Figs. 3 and 6). The

    plutonic rocks show a northeast-striking, steeply dip-

    ping magmatic foliation and down-dip mineral line-

    ation defined by euhedral amphibole crystals. Rocks

    are heterogeneously deformed across a 50-m-wide,

    well-exposed, quarry wall located along the eastern

    lineament of the LOFZ. Centimeter- to meter-wide

    shear zones affect the plutonic rocks. The mylonites

    are green-colored and extremely fine-grained; only afew porphyroclasts remain visible to the naked eye.

    The solid-state fabric is subparallel to the magmatic

    fabric and overprints it at the outcrop scale. Steeply

    plunging stretching lineations are defined by quartz

    ribbons and elongate porphyroclast systems of feldspar

    and recrystallized tails. Outcrop-scale kinematic indi-

    Fig. 3. Regional geology and structure of the Puyuhuapi transect showing major ductile shear zones along with respective stereoplots of

    foliations (crosses) and lineations (dots). Location of samples selected for 40Ar 39Ar analysis is shown with stars. The Queulat shear zone

    roughly coincides with the eastern boundary of the Miocene belt of the Patagonian Batholith at these latitudes.

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314294

  • 7/27/2019 Cenozoico Transpresional Nazca

    7/26

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 295

  • 7/27/2019 Cenozoico Transpresional Nazca

    8/26

    Fig. 4. Regional geology and structure of the Aysen transect showing major ductile shear zones along with respective stereoplots of foliations

    (crosses) and lineations (dots). Location of samples selected for 40Ar 39Ar analysis is shown with stars. The Canal Costa shear zone marks the

    western boundary of the Miocene belt of the Patagonian Batholith and the limit between the Central Depression to the west and the Main Range

    to the east.

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314296

  • 7/27/2019 Cenozoico Transpresional Nazca

    9/26

    cators such as C/S fabrics and asymmetric porphyr-

    oclasts suggest top-to-the-west dextral-oblique ductile

    thrusting. Consistent kinematic indicators, such as

    amphibole fish, were found at the microscopic scale

    (Fig. 5A). Diagnostic microstructures such as recrystal-

    lized aggregates of quartz and highly strained, frac-

    tured plagioclase crystals indicate greenschist facies

    conditions for the mylonitic deformation. Actinolite

    occurs around the edges of hornblende porphyroclasts

    (Fig. 5A) and within the foliated matrix where it

    coexists with fine-grained biotite. This mineral associ-ation is consistent with greenschist facies conditions of

    metamorphism, as suggested by the microstructure.

    2.2. Queulat shear zone

    This shear zone, hundreds of meters wide, is com-

    posed of weakly to moderately deformed metasedi-

    ments and metavolcanics that probably belong to the

    Traiguen Formation (Bobenrieth et al., 1983) (Figs. 2

    and 3). The Queulat shear zone roughly coincides with

    the eastern limit of the Miocene plutonic belt of the

    Patagonian batholith in the Puyuhuapi transect. Folia-

    tions, defined by flattened conglomerate pebbles and

    layers of mica and chlorite, strike north-northeast and

    dip moderately to steeply to the west. Stretching line-

    ations show variable pitches but consistently plunge

    to the southwest (Fig. 3). Shear sense is consistently

    dextral-reverse (top to the northeast) along north south

    trending, moderately dipping, ductile shear zones.

    2.3. Canal Jacaf shear zone

    This deformation zone, several tens of meters wide,

    juxtaposes the metamorphic basement to the west and

    the Traiguen Formation to the east (Fig. 3). The shear

    zone is characterized by both a penetrative west-

    dipping, steep to subvertical mylonitic foliation and

    poorly defined down-dip stretching lineation. Kine-

    matic indicators at the outcrop-scale suggest top to the

    east dextral-oblique thrusting of the metamorphic

    basement over the Traiguen Formation.

    Fig. 5. Photomicrographs of high strain rocks from the Puyuhuapi Quarry shear zone (A); Rio Cisnes shear zone (B,C) and Canal Costa shear

    zone (D). Sections cut perpendicular to the foliation and parallel to the stretching lineation. Scale bar is 1 mm. Mineral abbreviations:

    ac=actinolite; hb=hornblende; pl=plagioclase, ms=muscovite, bt=biotite. Fabric abbreviations: C: shear band, S: schistosity. Shear sense is

    indicated on each photograph.

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 297

  • 7/27/2019 Cenozoico Transpresional Nazca

    10/26

    2.4. Rio Cisnes shear zone

    This shear zone is oriented fN 60jE and is repre-

    sented in aerial photographs and satellite images as af30-km-long lineament following the length of Rio

    Cisnes (Figs.2 and 3). In thefield, the shear zone affects

    both the 10 Ma Puyuhuapi granite (Rb Sr, whole rock,

    Herve et al., 1993) and metamorphic basement wall-

    rocks of possible Paleozoic age (Fig. 3). High strain

    shear zones in the pluton and wallrock are centimeter to

    hundreds of meters in width, strike eastnortheast and

    dip moderately to the northwest. The stretching line-

    ation, defined by quartz-ribbons and streaks of musco-

    vite and biotite, plunges shallowly to moderately to the

    southwest. Mesoscopic-scale kinematic indicators,

    such as C/S fabrics and asymmetric porphyroclast

    systems, document dextral-oblique shear with a dip-

    slip, normal component of motion. Shear bands, folia-

    tion fish and mica fish (Fig. 5B,C) also indicate dextral

    shear with a normal, dip-slip component of motion.

    Recrystallized quartz aggregates, and the development

    of C/S fabrics suggest that deformation took place

    under mid-greenschist facies conditions (e.g. Shima-

    moto, 1989; Lister and Snoke, 1984).

    2.5. Islas Cinco Hermanas shear zone

    This high strain shear zone developed from meta-

    morphic rocks of the metamorphic basement at Islas

    Cinco Hermanas (Fig. 4). A north northeast striking

    steep foliation is defined by flattened and stretched

    quartz ribbons and mica foliae. A subhorizontal

    stretching lineation is very well developed. C/S fab-

    rics, z-shaped folds and sigmoidal porphyroclasts are

    seen on horizontal surfaces parallel to the lineation and

    perpendicular to the foliation. Shear sense is consis-

    tently dextral. Metamorphic conditions of mylonitiza-

    tion were probably greenschist facies, as documentedby very fine-grained aggregates of re-crystallized

    quartz and mica.

    2.6. Canal Costa shear zone

    This major shear zone follows the north south

    trend of the Canal deforming granitic rocks of the

    Patagonian Batholith and volcanic and sedimentary

    rocks of the Traiguen Formation (Figs. 2 and 4). At a

    regional scale, the Canal Costa represents a morpho-

    logic boundary between the so-called Central Depres-

    sion to the west and the Main Range to the east,

    dominated by outcrops of the Traiguen Formation and

    the Patagonian Batholith, respectively. Moreover,west of the channel, volcanic and sedimentary rocks

    are almost flat-lying whereas they are highly de-

    formed at the Canal Costa shear zone and do not

    occur extensively to the east of the channel (Figs. 2

    and 4). High strain cataclastic and mylonitic rocks

    occur as a discontinuous strip along the eastern flank

    of the channel. Most of them show north south

    trending foliations dipping steeply to the east. Stretch-

    ing lineations in the mylonites plunge moderately to

    the north, although other directions of plunge are

    observed (Fig. 4). Kinematic indicators document

    top to the southwest dextral-reverse ductile shearing

    (Fig. 5D); a few kinematic indicators showing dip-slip

    normal motions are locally found.

    The regional structural maps and the stereoplots

    show variable dips of foliations and shallowly to steeply

    plunging lineations throughout(Fig. 4). Most foliations

    dip to the east along the western margin of the Miocene

    plutonic belt of the North Patagonian Batholith whereas

    they dip to west along its eastern margin. Kinematic

    indicators in east-dipping mylonitic zones of the Canal

    Costa shear zone document top-to-the-west thrusting of

    the plutonic complex onto the Traiguen Formation. Atthe eastern side (Puyuhuapi and Queulat shear zones),

    kinematic indicators of ductile thrusting and dextral

    transpressional deformation predominate.

    3. Age of deformation

    High strain rocks commonly contain a complex mix

    of relic grains (porphyroclasts) in a matrix of dynam-

    ically recrystallized minerals. Moreover, the relic grains

    are usually partly recrystallized around their margins(core and mantle structure) whereas the so-called

    recrystallized grains often contain domains in which

    dynamic recrystallization has not been fully accom-

    plished as revealed by detailed SEM studies (e.g.

    Trimby et al., 1998). This means that a bulk separate

    of a specific mineral species may contain both relic and

    recrystallized isotopic signatures. Conventional step

    heating is likely to produce a geologically meaningless

    average age, unless age resolution is permitted by the

    fact that the two different generations of this mineral

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314298

  • 7/27/2019 Cenozoico Transpresional Nazca

    11/26

    species have distinctive argon retention properties (e.g.

    West and Lux, 1993). Scheuber et al. (1995) dated a

    continuous series of rocks from the undeformed proto-

    lith to a high-strain mylonite in the middle of a kilo-meter-wide ductile shear zone in northern Chile. In this

    study, they found that40Ar 39Ar ages are fully reset by

    deformation-induced recrystallization in the high-strain

    mylonites while mixed ages in the form of staircase

    spectral patterns were obtained in the low strain zones

    where both relic and recrystallized minerals coexist.For

    those samples in which both pre- and synkinematic

    grains of the same mineral species coexist at the sample

    scale, the 40Ar 39Ar laserprobe technique is one way to

    effectively date deformation (e.g. Reddyet al., 1996). In

    this method, the laser targets mineral grains (or parts

    thereof), which are interpreted to have been dynami-

    cally recrystallized at or below the closure temperature

    of the mineral. Thus, proper identification of targeted

    mineral species is critical, and the laser must be able to

    isolate these from other phases. If the target minerals

    have relatively high potassium concentrations, the laser

    analyses will be relatively insensitive to the inadvertent

    outgassing of contaminants (e.g. low K hornblende).

    For the present study, a total of 12 representative

    samples, all with good structural and kinematic control,

    were selected for dating. For 10 of these, mineral

    separates were prepared for conventional step heating;the remaining two were prepared for laserprobe analy-

    sis. For irradiation in the McMaster University reactor,

    separated mineral concentrates were wrapped in Al foil.

    Laserprobe analyses were carried out on polished slabs

    (f80 Am thick) that were irradiated at the same time as

    the mineral separates. The flux monitor was the horn-

    blende standard, MMhb-1 (assumed age=520F2 Ma).

    An internal tantalum resistance furnace of the double-

    vacuum type was used to carry out the step heating.

    Laserprobe analyses were made with a NdYAG

    system operated in the pulsed mode (12 kHz). Atargeted area on a rock section was fused by moving

    the sample chamber under the laser beam by means of

    an xy translation stage. All isotopic analyses were

    made using a VG 3600 mass spectrometer. Both unde-

    formed and high strain varieties of plutonic rocks were

    analyzed. For the latter, apparent ages of biotite and/or

    muscovite are likely to represent the age of deformation

    because mylonitic deformation was shown to have

    taken place at low to medium greenschist facies con-

    ditions (e.g. Kligfield et al., 1986; Dunlap et al., 1991).

    3.1. Results of the ArAr dating

    Apparent age data from step heating are reported

    with their 1r uncertainties; mean ages are providedwith 2r errors, the latter including the uncertainty in

    the parameter J, but not allowing for error in the

    assumed age of the flux monitor. Lithology, location,

    material dated, intensity of deformation and ArAr

    analytical data for each sample are listed in Table 1.

    For the following discussion, a plateau is defined as a

    set of contiguous steps that together contain at least

    50% of the total 39Ar released, and for which the

    apparent ages are indistinguishable.40Ar 39Ar apparent age spectra (and 37Ar/39Ar

    spectra for hornblendes) are shown in Fig. 6 for

    undeformed (low strain) rocks and in Fig. 7 for high

    strain rocks. Age data as discussed below are sum-

    marized in Table 1.

    Samples 95JC6, 95JC4, 95JC1 and 95JC12 are from

    the Queulat plutonic unit at the Puyuhuapi Quarry

    (Figs. 3 and 6). Sample 95JC6, an undeformed diorite,

    yielded well-defined ages of 14.4F0.6 and 14.4F0.3

    for hornblende and biotite, respectively (Fig. 6a,b).

    These are interpreted as cooling ages of the pluton

    immediately following emplacement perhaps at about

    16 Ma (RbSr whole rock age, Pankhurst and Herve,

    1994). Biotite from sample 95JC4, a low strain varietyof mylonitic granodioritediorite, yielded a spectrum

    that has relatively low ages with high errors in the early

    part of the gas release (Fig. 6c). Ages over the latter part

    of the release are better defined and have an average

    value of 5.3F0.3 Ma. Because this is a low strain rock,

    we interpret this age only as an upper limit to the time of

    solid-state deformation. Hornblende from 95JC1, a low

    strain mingled granodiorite diorite, yielded discordant

    age and 37Ar/39Ar spectra, possibly an indication of the

    presence of more than one generation of amphibole. We

    suggest that the mean age, 20.2F

    0.2 Ma, relates to theearly cooling history of this rock. Biotite from 95JC12,

    a low strain variety of granodiorite, yielded a spectrum

    that is only slightly discordant with a mean age of

    13.3F0.2 Ma. We interpret this as a cooling age,

    perhaps partially reset by later deformation. Biotite

    from 95JC14, a low strain biotite granite from Islas

    Cinco Hermanas in the Aysen fjord (Figs. 4 and 6f),

    yielded a mildly discordant spectrum with a mean age

    of 5.7F0.2 Ma. As for 95JC4 above, we interpret this

    as a partially reset age.

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 299

  • 7/27/2019 Cenozoico Transpresional Nazca

    12/26

    Table 1

    Summary of 40Ar 39Ar analytical data for low strain and high strain rocks of the Puyuhuapi and Aysen transects

    T (jC) mV39Ar 39Ar% Age (Ma)F1r %ATM 37Ar/39Ar 36Ar/40Ar 39Ar/40Ar %IIC

    95-JC-1 Biotite. Low strain mingled granodiorite diorite. Puyuhuapy transect550 41.5 3.0 128.3F3.8 960.0 67.57 0.031041 0.276388 12.75

    600 33.8 2.4 56.9F4.7 308.0 32.82 0.010247 0.151868 16.28

    650 41.4 2.9 25.9F2.7 167.0 15.91 0.005628 0.108746 18.44

    700 79.5 5.7 5.8F1.3 124.0 5.80 0.004197 0.178930 31.51

    750 89.9 6.5 2.1F0.1 115.0 3.88 0.003871 0.308042 59.86

    800 77.6 5.6 1.3F1 110.0 3.37 0.003713 0.337276 81.72

    850 82.9 6.0 0.4F0.7 95.0 2.40 0.003192 0.474628 177.89

    900 84.1 6.0 1.2F0.7 84.5 1.91 0.002839 0.494858 47.30

    950 136.1 9.8 3.1F0.4 57.1 0.94 0.001926 0.560082 9.43

    975 140.8 10.1 3.2F0.3 46.1 0.72 0.001561 0.687272 7.13

    1000 160.2 11.5 3.6F0.2 33.2 0.53 0.001131 0.748429 4.57

    1025 132.8 9.6 3.5F0.3 31.8 0.55 0.001091 0.791464 4.97

    1050 105.2 7.6 3.5F0.3 30.9 0.61 0.001068 0.796960 5.52

    1100 156.8 11.3 3.9F0.2 28.8 0.39 0.000978 0.742279 3.131200 17.5 1.2 1.4F3.2 105.0 2.39 0.003493 0.170124 55.31

    1450 1.6 0.1 104.4F225.2 103.0 33.92 0.003492 0.001414 8.30

    Mean age (950 1100 jC)=3.5F0.2 Ma; J=0.00232F0.0000232 (1%)

    95-JC-1 Hornblende. Low strain mingled granodiorite diorite. Puyuhuapi transect

    650 10.1 3.3 62.5F10.8 88.5 2.83 0.002999 0.007514 1.58

    750 12.3 4.0 15.2F7.6 94.7 2.19 0.003209 0.014355 4.63

    850 13.4 4.4 2.9F4.8 97.0 2.88 0.003293 0.041715 30.37

    950 42.6 14.1 19.5F2.1 80.2 9.99 0.002721 0.041969 16.61

    975 44.0 14.5 21.7F1.5 64.3 12.98 0.002194 0.067942 19.48

    1000 49.0 16.2 17.5F1.3 60.8 13.18 0.002087 0.092551 24.39

    1025 19.1 6.3 14.0F2.7 66.8 11.38 0.002315 0.097334 26.10

    1050 8.2 2.7 14.2F5.3 75.3 10.24 0.002603 0.071654 23.16

    1075 6.3 2.0 10.0F9.0 91.2 13.15 0.003115 0.036222 41.87

    1100 7.3 2.4 15.0F9.2 90.9 15.88 0.003095 0.024947 34.08

    1125 12.4 4.1 18.4F6.2 88.8 15.90 0.003017 0.025098 28.00

    1150 12.0 3.9 27.0F5.8 84.1 15.80 0.002856 0.024399 19.24

    1200 33.9 11.2 31.5F2.7 73.7 15.71 0.002503 0.034417 16.50

    1250 11.8 3.9 21.9F6.4 88.0 15.88 0.002989 0.022554 23.58

    1350 13.7 4.5 12.7F8.3 96.7 15.88 0.003275 0.010757 40.15

    1450 5.0 1.6 53.4F35.2 103 14.03 0.003498 0.002658 7.46

    Mean age (950 1350 jC)=20.2F2 Ma; J=0.00232F0.0000232 (1%)

    95-JC-2 Biotite. Low strain mingled granodiorite diorite. Puyuhuapi transect

    550 19.6 1.5 132.6F8.2 1634.0 71.00 0.043443 0.387645 12.82

    600 10.0 0.7 116.6F11.8 1634.0 60.12 0.036556 0.364203 12.81

    650 12.9 1.0

    56.2F

    8.2 289.0 31.42 0.009248 0.132213 15.80700 32.4 2.5 11.0F3.2 137.0 7.72 0.004587 0.140649 21.76

    750 106.8 8.3 0.3F0.8 103.0 1.66 0.003469 0.593526 192.52

    800 273.9 21.3 1.7F0.2 39.8 0.43 0.001367 1.394780 7.76

    850 268.2 20.9 1.7F0.1 24.1 0.32 0.000865 1.719379 5.82

    900 157.0 12.2 1.3F0.2 39.7 0.47 0.001381 1.747207 10.91

    950 60.6 4.7 0.5F0.6 124.0 1.27 0.003841 1.877213 88.76

    975 35.7 2.7 1.5F0.6 374.0 1.97 0.006809 3.500425 42.98

    1000 48.9 3.8 0.1F0.4 100.0 1.22 0.003034 2.374611 12370.68

    1025 56.8 4.4 0.7F0.3 62.3 0.89 0.002059 1.846029 39.47

    1050 52.6 4.1 0.5F0.3 69.5 0.83 0.002234 1.837870 44.83

    1100 109.0 8.5 1.8F0.2 39.8 0.20 0.00131 1.315038 3.51

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314300

  • 7/27/2019 Cenozoico Transpresional Nazca

    13/26

    T (jC) mV39Ar 39Ar% Age (Ma)F1r %ATM 37Ar/39Ar 36Ar/40Ar 39Ar/40Ar %IIC

    95-JC-2 Biotite. Low strain mingled granodiorite diorite. Puyuhuapi transect

    1200 37.1 2.8 1.7F1.6 87.8 0.28 0.002891 0.289009 5.22

    1450 0.9 0.0 66.2F463.1 101.0 36.56 0.003419 0.000922 15.31

    Mean age (800900 jC)=1.6F0.2 Ma; J=0.00232F0.0000232 (1%)

    95-JC-4 Biotite. Low strain mingled granodioritediorite. Puyuhuapi transect. Puy. Quarry shear zone

    550 31.7 2.2 98F6.8 417.0 61.95 0.013616 0.132815 16.38

    600 22.3 1.5 52.2F7.4 240.0 32.65 0.007926 0.109694 17.81

    650 28.9 2.0 17.0F5.6 126.0 13.11 0.004261 0.065689 23.51

    700 56.6 4.0 1.0F0.3 101.0 3.82 0.003426 0.080565 132.82

    750 84.0 5.9 2.4F1.7 88.4 2.10 0.002972 0.196596 27.22

    800 69.3 4.9 2.0F1.9 89.5 2.43 0.003004 0.214004 38.11

    850 84.5 6.0 3.6F1.2 69.5 1.72 0.002334 0.345849 15.06

    900 77.2 5.5 3.9F1.1 59.0 1.43 0.001988 0.427411 11.66

    950 102.7 7.3 4.9F0.8 48.1 0.91 0.001621 0.432076 5.90

    975 125.7 8.9 5.1F0.5 29.9 0.71 0.001028 0.558256 4.411000 204.8 14.6 5.3F0.3 18.1 0.34 0.000624 0.62939 2.05

    1025 225.6 16.1 5.6F0.2 11.0 0.24 0.000386 0.644185 1.36

    1050 150.2 0.7 5.4F0.2 15.1 0.31 0.000525 0.637975 1.84

    1100 95.8 6.8 5.1F0.5 28.5 0.46 0.000965 0.561897 2.85

    1200 35.8 2.5 3.5F2.4 77.8 1.27 0.002574 0.256589 11.42

    1450 4.7 0.3 30.0F46.9 104.0 14.31 0.003515 0.006162 14.21

    Mean age (950 1100 jC)=5.3F0.3 Ma; J=0.00232F0.0000232 (1%)

    95-JC-4 hornblende. Low strain mingled granodioritediorite. Puyuhuapi transect. Quarry shear zone

    650 18.1 5.8 49.9F8.0 124.0 37.38 0.004218 0.021157 21.42

    750 15.2 4.8 29.4F9.8 109.0 25.49 0.003707 0.014044 25.81

    850 14.1 4.5 16.5F7.6 107.0 20.37 0.003633 0.019659 37.64

    950 14.8 4.7 14.0F6.0 110.0 18.72 0.003713 0.030779 41.23

    975 9.7 3.1

    13.5F

    7.7 110.0 25.13 0.003716 0.032562 57.31

    1000 13.4 4.3 1.7F5.7 101.0 25.99 0.003413 0.033072 504.01

    1025 30.5 9.7 13.5F2.7 85.6 24.43 0.002894 0.043870 57.80

    1050 59.7 19.0 25.7F1.9 76.3 19.54 0.002582 0.038096 24.98

    1100 18.6 5.9 25.2F3.3 64.8 18.46 0.002200 0.057060 23.96

    1125 4.6 1.4 18.3F10.7 81.0 24.55 0.002718 0.041820 43.30

    1150 6.2 1.9 26.5F8.3 74.2 25.56 0.002503 0.039426 31.63

    1175 12.9 4.1 34.7F4.8 68.1 22.51 0.002304 0.037658 21.64

    1200 11.9 3.8 38.4F5.2 65.9 22.33 0.002229 0.036261 19.48

    1250 38.5 12.3 37.5F2.1 63.5 21.07 0.002150 0.040068 18.81

    1350 34.1 10.8 37.2F2.4 63.7 21.83 0.002158 0.040106 19.62

    1450 10.1 3.2 58.0F7.3 69.3 23.52 0.002335 0.021646 14.05

    Mean age (11751350 jC)=37.2F3 Ma; J=0.00232F0.0000232 (1%)

    95-JC-6 Biotite. Undeformed granodiorite. Puyuhuapi transect. Seno Queulat

    550 1.3 0.0 18.2F41.6 86.5 36.29 0.002733 0.027781 64.46

    600 7.1 0.3 15.9F12.4 78.1 6.47 0.002558 0.055596 13.08

    650 36.4 1.9 15.5F3.3 63.6 1.34 0.002120 0.096896 2.79

    700 108.0 5.8 14.8F0.9 41.5 0.47 0.001390 0.163844 1.03

    750 204.6 11.1 14.6F0.3 16.7 0.23 0.000558 0.236413 0.50

    800 310.8 16.9 14.3F0.2 7.5 0.14 0.000252 0.267607 0.31

    850 196.9 10.7 14.4F0.2 7.8 0.26 0.000265 0.264022 0.58

    900 102.9 5.6 14.2F0.4 15.3 0.71 0.000518 0.244703 1.60

    950 107.3 5.8 14.3F0.6 20.7 0.70 0.000695 0.228504 1.59

    (continued on next page)

    Table 1 (continued)

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 301

  • 7/27/2019 Cenozoico Transpresional Nazca

    14/26

    T (jC) mV39Ar 39Ar% Age (Ma)F1r %ATM 37Ar/39Ar 36Ar/40Ar 39Ar/40Ar %IIC

    95-JC-6 Biotite. Undeformed granodiorite. Puyuhuapi transect. Seno Queulat

    1000 104.9 5.7 14.2F

    0.5 17.2 0.77 0.000582 0.240189 1.751050 209.7 11.4 14.2F0.4 11.2 0.43 0.000380 0.258885 0.97

    1100 287.1 15.6 14.3F0.2 8.4 0.37 0.000286 0.265203 0.84

    1200 150.3 8.2 14.3F0.3 14.1 0.78 0.000482 0.247193 1.76

    1450 3.0 0.1 250.4F47.7 53.7 454.86 0.001918 0.007100 82.78

    Mean age (650 1200 jC)=14.4F0.3 Ma; J=0.00232F0.0000232 (1%)

    95-JC-6 Hornblende. Undeformed granodiorite. Puyuhuapi transect. Seno Queulat

    650 18.0 2.1 16.0F3.9 79.1 0.68 0.002686 0.054145 1.38

    750 61.6 7.3 11.3F1.0 54.8 0.23 0.001862 0.166933 0.65

    850 61.5 7.3 10.9F0.9 47.1 0.46 0.001606 0.201649 1.36

    900 29.5 3.5 12.7F1.6 50.3 1.60 0.001730 0.162285 4.04

    950 33.0 3.9 15.5F2.0 61.7 5.19 0.002115 0.102282 10.81

    975 51.1 6.0 17.5F1.4 56.3 8.24 0.001936 0.102916 15.24

    1000 225.6 26.8 15.2F0.4 47.0 9.23 0.001609 0.144158 19.551025 152.1 18.0 14.3F0.5 28.2 9.07 0.001030 0.204592 20.37

    1050 59.9 7.1 14.3F0.8 23.3 6.49 0.000926 0.215059 14.55

    1075 40.1 4.7 12.9F1.0 32.2 3.65 0.001172 0.213963 9.04

    1100 42.2 5.0 12.9F1.3 47.5 5.63 0.001668 0.166737 13.99

    1125 23.9 2.8 12.9F2.6 66.9 9.81 0.002314 0.104947 24.31

    1150 7.9 0.9 7.8F8.0 91.9 12.08 0.003140 0.042667 49.06

    1200 6.5 0.7 0.4F11.9 99.0 12.65 0.003390 0.023506 840.20

    1250 4.3 0.5 5.7F19.4 102.0 12.56 0.003467 0.015287 69.57

    1350 13.4 1.5 1.2F8.7 99.0 11.29 0.003373 0.015062 290.56

    1450 10.0 1.1 8.5F15.8 98.6 2.44 0.003342 0.006368 9.12

    Mean age (10001125 jC)=14.4F0.6 Ma; J=0.00232F0.0000232 (1%)

    95-JC12 Biotite. Undeformed granodiorite. Puyuhuapi transect. Seno Queulat

    550 6.8 0.4

    115.6F

    12.5 544.0 69.12 0.016499 0.145725 14.89

    600 15.7 0.9 18.6F5.2 129.0 19.28 0.004327 0.065755 31.57

    650 57.6 3.5 9.8F1.4 76.4 3.23 0.002568 0.099398 10.50

    700 196.4 12.0 13.4F0.3 36.6 0.54 0.001231 0.196102 1.29

    750 207.8 12.7 13.2F0.2 13.2 0.39 0.000443 0.272260 0.95

    800 283.2 17.3 13.5F0.1 7.9 0.25 0.000268 0.282850 0.60

    850 140.5 8.6 13.1F0.2 10.4 0.40 0.000351 0.281216 0.98

    900 90.2 5.5 12.8F0.4 15.4 0.62 0.000521 0.270966 1.56

    950 79.4 4.8 12.6F0.4 19.9 0.56 0.000665 0.262184 1.44

    975 51.4 3.1 12.6F0.5 15.8 0.81 0.000540 0.271966 2.07

    1000 86.0 5.2 13.1F0.3 14.0 0.41 0.000468 0.269544 1.01

    1025 79.9 4.8 13.3F0.3 9.3 0.24 0.000308 0.279619 0.58

    1050 94.9 5.8 13.7F0.2 7.1 0.03 0.000230 0.279282 0.07

    1100 197.8 12.1 13.9F

    0.2 8.9 0.00 0.000290 0.272419 0.021200 43.3 2.6 13.2F0.9 34.9 0.14 0.001140 0.202711 0.35

    1450 1.2 0.0 96.8F314.6 102.0 10.89 0.003449 0.001018 2.92

    Mean age (850 1200 jC)=13.3F0.2 Ma; J=0.00232F0.0000232 (1%)

    96GA01 Biotite. Mylonite. Puyuhuapy transect. Puyuhuapi Quarry shear zone

    550 3.9 0.3 32.7F20.6 92.0 0.07 0.003107 0.010492 0.08

    600 12.2 1.0 17.9F9.1 94.4 0.01 0.003193 0.013503 0.03

    650 30.1 2.5 11.6F4.0 93.6 0.02 0.003168 0.023668 0.07

    700 49.2 4.2 6.8F2.2 92.7 0.11 0.003134 0.046644 0.57

    750 49.3 4.2 6.6F1.8 92.0 0.02 0.003113 0.051777 0.13

    800 48.1 4.1 9.7F2.6 92.9 0.00 0.003144 0.031672 0.02

    Table 1 (continued)

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314302

  • 7/27/2019 Cenozoico Transpresional Nazca

    15/26

    T (jC) mV39Ar 39Ar% Age (Ma)F1r %ATM 37Ar/39Ar 36Ar/40Ar 39Ar/40Ar %IIC

    96GA01 Biotite. Mylonite. Puyuhuapy transect. Puyuhuapi Quarry shear zone

    850 81.2 6.9 7.3F

    1.6 91.6 0.03 0.003099 0.049540 0.15900 88.0 7.5 6.3F0.8 79.5 0.03 0.002686 0.140882 0.16

    950 170.6 14.6 5.0F0.4 74.4 0.01 0.002514 0.222908 0.12

    1000 263.8 22.6 4.1F0.2 51.6 0.02 0.001742 0.512718 0.21

    1050 213.1 18.3 4.2F0.1 29.5 0.07 0.000997 0.730535 0.59

    1100 143.1 12.3 4.4F0.2 24.9 0.01 0.000840 0.730962 0.12

    1150 8.7 0.7 12.9F1.3 24.5 0.28 0.000818 0.248039 0.73

    1200 1.3 0.1 60.6F24.6 64.4 2.32 0.002145 0.024916 1.39

    1450 0.7 0.0 206.6F163 87.9 2.27 0.002965 0.002414 0.49

    Mean age (10001100 jC)=4.2F0.2 Ma; J=0.002438F0.000018 (0.7%)

    96GA03 Biotite. Mylonite. Puyuhuapy transect. Puyuhuapi Quarry shear zone

    600 12.1 1.3 9.6F6.3 95.5 0.21 0.003231 0.020024 0.74

    700 56.8 6.5 6.7F2.2 94.6 0.31 0.003203 0.034368 1.55

    750 38.2 4.4 5.1F1.9 94.2 0.26 0.003186 0.049435 1.74800 42.7 4.9 6.9F2.0 92.9 0.26 0.003144 0.044138 1.24

    850 65.8 7.5 9.2F1.6 90.3 0.16 0.003054 0.045940 0.59

    900 60.8 7.0 8.9F0.8 71.4 0.12 0.002415 0.138857 0.45

    950 99.3 11.4 4.4F0.5 74.6 0.10 0.002524 0.246991 0.76

    1000 187.4 21.5 3.9F0.2 49.4 0.07 0.001672 0.559732 0.64

    1050 176.8 20.3 3.8F0.2 40.6 0.36 0.001384 0.67938 3.18

    1100 103.9 11.9 3.5F0.3 58.1 0.39 0.001970 0.521886 3.75

    1150 18.0 2.0 5.0F1.8 79.5 1.19 0.002690 0.175269 7.78

    1200 3.0 0.3 5.5F14.6 96.1 5.33 0.003246 0.030209 32.26

    1450 2.9 0.3 13.4F34.5 97.9 6.78 0.003311 0.006694 17.01

    Mean age (10001100 jC)=3.8F0.3 Ma. J=0.002434F0.000018 (0.7%)

    95GA04 Biotite. Low strain mingled granodioritediorite. Puyuhuapi transect. Puy. Quarry shear zone

    550 2.4 0.2 14.4F13.8 89.3 0.49 0.002990 0.031890 1.15

    600 13.4 1.1 4.5F2.9 93.0 0.20 0.003137 0.066126 1.45

    650 30.1 2.5 6.1F1.5 88.2 0.12 0.002980 0.083825 0.65

    700 52.7 4.5 5.6F0.8 82.5 0.11 0.002786 0.134880 0.69

    750 59.2 5.0 5.9F0.7 79.7 0.06 0.002694 0.148544 0.36

    800 54.6 4.6 6.3F1.3 89.8 0.11 0.003038 0.069375 0.58

    850 81.8 7.0 6.2F0.7 83.2 0.06 0.002813 0.117719 0.35

    900 80.7 6.9 4.9F0.4 67.7 0.07 0.002286 0.282460 0.52

    950 136.4 11.6 4.6F0.2 53.6 0.05 0.001811 0.433324 0.38

    1000 275.4 23.5 4.0F0.1 33.2 0.04 0.001124 0.711898 0.35

    1050 235.7 20.1 4.1F0.1 28.8 0.18 0.000978 0.743037 1.45

    1100 131.0 11.2 4.3F0.1 27.0 0.08 0.000914 0.732523 0.61

    1200 10.3 0.8 3.6F2.7 87.1 1.20 0.002920 0.153202 11.08

    1450 4.2 0.3 23.0F

    14.5 92.3 1.07 0.003117 0.014366 1.59Mean age (950 1050 jC)=4.2F0.1 Ma; J=0.002432F0.000018 (0.7%)

    96-GA-26 Muscovite. High strain bt-ms schist. Puyuhuapi transect. Puerto Cisnes shear zone

    550 2.4 0.1 7.8F6.2 76.9 0.00 0.002385 0.111454 0.00

    600 5.6 0.3 8.6F1.7 44.8 0.11 0.001571 0.272863 0.44

    650 14.0 0.7 7.3F1.2 44.7 0.00 0.001536 0.318981 0.02

    675 17.2 0.9 6.9F1.1 44.8 0.03 0.001538 0.338126 0.18

    700 21.7 1.2 5.1F1.5 66.2 0.03 0.002259 0.278627 0.20

    725 31.8 1.7 5.9F1.0 52.4 0.02 0.001786 0.335482 0.11

    750 43.0 2.3 5.6F0.7 48.6 0.01 0.001658 0.384821 0.08

    Table 1 (continued)

    (continued on next page)

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 303

  • 7/27/2019 Cenozoico Transpresional Nazca

    16/26

    T (jC) mV39Ar 39Ar% Age (Ma)F1r %ATM 37Ar/39Ar 36Ar/40Ar 39Ar/40Ar %IIC

    96-GA-26 Muscovite. High strain bt-ms schist. Puyuhuapi transect. Puerto Cisnes shear zone

    775 67.2 3.7 5.8F

    0.7 61.9 0.00 0.002101 0.273282 0.01800 242.4 13.4 6.2F0.2 51.5 0.00 0.001745 0.324493 0.01

    825 323.1 17.8 6.3F0.1 25.6 0.00 0.000869 0.489688 0.00

    850 223.4 12.3 6.3F0.1 18.5 0.00 0.000627 0.535572 0.00

    900 275.6 15.2 6.3F0.1 25.0 0.00 0.000848 0.496512 0.00

    950 158.7 8.7 6.2F0.3 37.3 0.00 0.001266 0.421508 0.00

    1000 92.6 5.1 6.1F0.5 46.4 0.00 0.001575 0.362133 0.04

    1100 209.0 11.5 6.1F0.3 43.5 0.00 0.001474 0.381153 0.02

    1200 74.3 4.1 5.9F0.7 63.0 0.01 0.002139 0.258096 0.06

    1450 5.9 0.3 64.8F32.5 104.0 0.23 0.003525 0.002732 0.10

    Mean age (7501000 jC)=6.2F0.2 Ma; J=0.00232F0.0000232 (1%)

    95-JC-14 Biotite. Undeformed granite. Aysen transect. Isla Cinco Hermanas

    550 11.2 0.6 67.8F6.8 356.0 39.55 0.011387 0.150647 16.09

    600 43.8 2.5

    4.6F

    1.8 111.0 6.14 0.003754 0.109710 42.33650 149.8 8.7 4.3F0.5 78.2 1.21 0.002629 0.206765 8.77

    700 297.2 17.3 5.8F0.2 42.6 0.44 0.001430 0.408531 2.41

    750 366.1 21.3 6.2F0.1 24.6 0.29 0.000825 0.503958 1.50

    800 200.6 11.7 5.9F0.2 42.6 0.44 0.001424 0.400065 2.36

    850 87.4 5.1 4.7F0.4 40.3 0.95 0.001342 0.509703 6.28

    900 67.1 3.9 4.5F0.5 47.2 1.13 0.001561 0.477224 7.97

    950 84.7 4.9 5.2F0.5 51.0 0.72 0.001688 0.387755 4.43

    975 70.5 4.1 4.9F0.7 66.7 0.68 0.002213 0.278840 4.35

    1000 76.0 4.4 5.3F0.4 38.0 0.66 0.001251 0.478727 3.99

    1025 70.9 4.1 6.0F0.4 35.4 0.22 0.001146 0.443894 1.20

    1050 56.8 3.3 6.4F0.5 38.6 0.02 0.001240 0.394752 0.07

    1100 92.0 5.3 6.3F0.5 43.6 0.06 0.001430 0.367440 0.32

    1200 35.8 2.0 4.6F1.3 73.2 0.72 0.002397 0.237803 4.92

    1450 1.4 0.0

    106.8F

    202.5 103.0 27.84 0.003495 0.001509 6.62

    Mean age (7001200 jC)=5.7F0.2 Ma; J=0.00232F0.0000232 (1%)

    95GA17 Muscovitebiotite bands. Laser spots on a polished slab. High strain schist. Aysen transect. Isla Cinco Hermanas shear zone

    Spot mV 39Ar Age (Ma)F2r %ATM 37Ar/39Ar 36Ar/40Ar 39Ar/40Ar %IIC

    1 24.7 6.8F0.4 34.5 0.04 0.001169 0.034083 0.01

    2 44.3 7.7F0.3 39.4 0.05 0.001333 0.027900 0.02

    3 34.5 6.2F0.5 61.6 0.10 0.002087 0.021826 0.05

    4 45.2 4.6F0.3 59.4 0.10 0.002010 0.030686 0.06

    5 40.2 8.0F0.3 38.2 0.09 0.001295 0.027313 0.03

    6 38.4 5.6F0.4 63.4 0.16 0.002146 0.023076 0.09

    7 48.7 4.6F0.2 47.1 0.16 0.001594 0.039999 0.10

    8 67.6 5.9F0.2 40.6 0.07 0.001376 0.035211 0.03

    9 26.3 7.5F0.4 45.0 0.15 0.001525 0.025958 0.06

    10 30.9 5.3F0.4 52.0 0.14 0.001761 0.031881 0.08

    11 26.2 9.2F0.4 25.9 0.19 0.000876 0.028424 0.07

    12 25.0 8.5F0.5 42.2 0.20 0.001429 0.024006 0.07

    Mean age (spots 1 12 )=6.4F0.6 Ma; J=0.000197F0.00002 (10.1%)

    95-GA-19 Muscovite. High strain schist. Aysen transect. Isla Cinco Hermanas shear zone

    T (jC) mV39Ar 39Ar% Age (Ma)F1r %ATM 37Ar/39Ar 36Ar/40Ar 39Ar/40Ar %IIC

    550 6.6 0.4 6.1F8.1 92.2 0.04 0.003119 0.053169 0.21

    600 54.0 3.4 3.1F0.8 68.3 0.05 0.002312 0.415424 0.56

    Table 1 (continued)

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314304

  • 7/27/2019 Cenozoico Transpresional Nazca

    17/26

    Biotites from sample 96GA01 and 96GA03, both

    high strain mylonites, gave spectra with scattered,

    poorly defined ages over the first half of the gas

    release, but with plateaus over the final high temper-ature heating steps. The plateau ages are 4.2F0.2 and

    3.8F0.2 Ma, respectively (Fig. 7a,b). We interpret

    these as the time of high strain solid-state contractio-

    nal ductile deformation that affected the Queulat

    plutonic unit. Sample 96GA26 is from a high strain

    pelitic schist in the Rio Cisnes shear zone (Table 1,

    Fig. 3). A muscovite separate from this rock gave a

    spectrum with a very well-defined plateau (11 steps,

    over > 95% of the gas release) at an age of 6.2F0.2

    Ma (Fig. 7c). This age could be interpreted as the time

    when these rocks cooled to muscovite closure temper-

    atures following intrusion of the Puerto Cisnes pluton

    (at ca. 10 Ma). More likely, it dates solid-state defor-

    mation and recrystallization of muscovite because therocks were deformed under greenschist facies condi-

    tions, and the muscovite grains define extensional

    crenulation cleavage indicating dextral transtensional

    deformation.

    Samples 95GA17 and 95GA19 are from high

    strain quartzmica schists collected from Islas Cinco

    Hermanas in the Aysen fjord (Fig. 4). A fine-grained

    muscovite separate from 95GA19 yielded a spectrum

    that has ages generally increasing from a low off3

    4 Ma to a high off8 Ma (Fig. 7d). This may result

    T (jC) mV39Ar 39Ar% Age (Ma)F1r %ATM 37Ar/39Ar 36Ar/40Ar 39Ar/40Ar %IIC

    650 86.4 5.6 3.4F0.3 37.9 0.04 0.001286 0.749365 0.41675 80.1 5.1 3.8F0.3 29.7 0.04 0.001010 0.755144 0.35

    700 72.4 4.6 3.4F0.3 37.1 0.04 0.001260 0.767762 0.40

    725 71.8 4.6 3.8F0.2 13.0 0.03 0.000447 0.940648 0.32

    750 75.8 4.9 4.0F0.2 15.6 0.03 0.000533 0.877467 0.26

    775 110.9 7.1 4.9F0.2 22.5 0.02 0.000765 0.659609 0.18

    800 89.3 5.7 5.7F0.2 13.1 0.02 0.000448 0.635116 0.14

    825 106.8 6.9 5.0F0.2 17.7 0.02 0.000603 0.682395 0.13

    850 114.5 7.4 4.9F0.2 20.4 0.02 0.000695 0.671934 0.13

    900 217.5 14.1 5.5F0.2 22.6 0.01 0.000767 0.578451 0.08

    950 204.9 13.2 5.5F0.2 27.1 0.01 0.000918 0.552298 0.10

    1000 138.6 8.9 7.2F0.3 24.6 0.02 0.000834 0.431066 0.11

    1100 90.9 5.8 8.6F0.5 43.9 0.08 0.001487 0.270910 0.29

    1200 21.6 1.4 9.8F3.4 87.6 0.37 0.002965 0.052524 1.23

    Total gas age=5.2F0.2 Ma; J=0.00232F0.0000116 (0.5%)

    95GA19 Muscovite biotite bands. Laser spots on a polished slab. High strain schist. Aysen transect. Isla Cinco Hermanas shear zone

    Spot mV 39Ar Age (Ma)F2r %ATM 37Ar/39Ar 36Ar/40Ar 39Ar/40Ar %IIC

    1 2.0 7.4F6.4 75.1 0.38 0.002541 0.011825 0.16

    2 1.9 4.9F4.7 74.8 0.34 0.002530 0.017946 0.20

    3 3.7 4.7F1.8 63.0 0.23 0.002130 0.027820 0.14

    4 21.3 4.6F0.3 43.7 0.06 0.001480 0.042544 0.04

    5 20.1 4.2F0.6 74.6 0.13 0.002525 0.021266 0.09

    6 19.2 4.0F0.4 63.1 0.01 0.002136 0.032294 0.01

    7 11.1 4.9F0.9 67.2 0.01 0.002273 0.023639 0.00

    8 10.5 4.4F0.9 68.8 0.01 0.002327 0.024947 0.00

    9 5.4 5.9F

    7.1 91.4 0.17 0.003093 0.005105 0.0810 2.9 4.2F5.2 85.4 0.42 0.002891 0.011999 0.29

    11 1.3 13.9F64.2 93.9 2.34 0.003179 0.001532 0.60

    12 9.5 5.3F1.7 79.6 0.21 0.002696 0.013489 0.11

    Mean age (spots 112)=4.7F1.0 Ma; J=0.000197F0.00002 (10.1%)

    Analysis are from bulk mineral separates unless otherwise indicated (laserspots on oriented slabs). Mean ages are reported with a 2r error37Ar 39Ar, 36Ar 40Ar and 40Ar 39Ar ratios are corrected for mass spectrometer discrimination, interfering isotopes and system blanks.

    %IIC=interfering isotopes correction.

    Table 1 (continued)

    95-GA-19 Muscovite. High strain schist. Aysen transect. Isla Cinco Hermanas shear zone

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 305

  • 7/27/2019 Cenozoico Transpresional Nazca

    18/26

    Fig. 6. 40Ar 39Ar apparent age spectra for undeformed (low strain) samples of the two transects. Samples 95JC6, 95JC4, 95JC1, 95JC12 are

    from the Puyuhuapi transect, sample 95JC14 is from the Aysen transect. Mean and plateau (for 95JC6 Bt) ages as discussed in text are indicated.

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314306

  • 7/27/2019 Cenozoico Transpresional Nazca

    19/26

    Fig. 7. 40Ar 39Ar apparent age spectra for samples of high strain rocks of the two transects. Samples 96GA01, 96GA03 and 96GA26 are from

    the Puyuhuapi transect, samples 95GA17 and 95GA19 are from the Ayse n transect. Plateau ages as discussed in text are indicated.

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 307

  • 7/27/2019 Cenozoico Transpresional Nazca

    20/26

    from the partial resetting of old muscovite grains by a

    high strain deformation event atf34 Ma. Other-

    wise, the spectrum might represent the mixture of two

    populations of grains: (i) a relict population repre-sented by older ages, and (ii) a recrystallized popula-

    tion represented by the youngest ages. Results from

    the laserprobe dating of these two samples, described

    below, favors the second of these two hypotheses.

    The laser dating, carried out on selected thick

    sections, targeted in each case muscovite biotite

    fine-grained aggregates in sigmoidal lenses parallel

    to the foliation. Results for each sample are shown in

    Fig. 7e,f on plots of apparent age versus 39Ar abun-

    dance. For 95GA19, many of the spot/area analyses

    yielded relatively small amounts of gas (i.e.

  • 7/27/2019 Cenozoico Transpresional Nazca

    21/26

    the Patagonian Batholith and wallrocks. The bulk of

    the transpressional deformation took place at around 4

    Ma as documented by 40Ar 39Ar dating on recrystal-

    lized biotite and muscovite from several greenschistfacies shear zones. A slightly older dextral transten-

    sional event may have taken place in the region at

    about 6 Ma.

    The present-day spatial distribution of geologic

    units, the overall topography of the orogen, and the

    mostly contractional/dextral strike-slip nature of their

    boundaries strongly suggests the development of a

    transpressional pop-up structure defining the Main

    Range of the Patagonian Cordillera (Fig. 8). The

    Cenozoic magmatic arc rocks are thrusted westwards

    over the basement/Traiguen Formation and appear to

    be thrusted eastward over the Cretaceous belt of the

    Patagonian batholith. This regional-scale pop-up like

    structure is very similar to that obtained through three-

    dimensional numerical dynamic modeling of trans-

    pressional deformation at obliquely convergent plate

    margins (Braun and Beaumont, 1995). Recent exper-

    imental analog modeling of transpressional deforma-

    tion by Schreurs and Colletta (1998) also shows

    striking similarities with the field data. Furthermore,

    the series of en echelon structures joining the two

    main boundaries of the Cenozoic plutonic belt in

    southern Chile resembles the duplex-like structureobserved in the same analog models. The observation

    that the Cenozoic belt of the NPB, which constitutes

    the axis of the Main Range, has been differentially

    exhumed with respect to the Central Depression to the

    west and the foreland region to the east has a

    fundamental bearing on orogenic processes operating

    at convergent margins. A long-lived transpressional

    magmatic arc, such as the one described here, con-

    centrates the bulk of deformation by crustal thicken-

    ing and strike-slip movements giving rise to a

    topographic relief by double-verging thrust zones asdepicted by Braun and Beaumonts numerical model.

    Ridge collision and oblique subduction have been

    proposed as alternative driving mechanisms of trans-

    pressional deformation at the leading edge of the

    South American plate. Collision of successive seg-

    ments of the Chile Ridge took place between 6 and 3

    Ma, close to the present position of the triple junction

    (Fig. 9). Dextral right-oblique subduction, in turn, has

    prevailed during most of the Cenozoic (Fig. 1).

    Theoretically, ridge subduction favors margin-orthog-

    onal contraction close to the indenter and oblique-slip

    to strike-slip deformation a few hundred kilometers

    away from it (Tapponier and Molnar, 1976; Nelson et

    al., 1994) (Fig. 10a). On the other hand, obliquesubduction is thought to produce overall transpres-

    sional deformation along ancient and present-day

    plate boundaries (e.g. Jarrard, 1986; Beck, 1991;

    McCaffrey, 1992; Teyssier et al., 1995) (Fig. 10b).

    We envisage oblique subduction as the long-term

    driving mechanism of dextral transpression in the

    southern Andes. Previous studies proposed that dex-

    tral transtension occurred locally along the magmatic

    arc during early Tertiary times when subduction was

    Fig. 9. Migration of the Chile Ridge and fracture zones with respect

    to the plate boundary during the last 6 Ma (Bourgois et al., 1996).

    Several northnorthwest trending short segments of the ridge were

    consecutively subducted from around 6 Ma to present-day close to

    the Taitao Peninsula.

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 309

  • 7/27/2019 Cenozoico Transpresional Nazca

    22/26

    highly dextral-oblique and intraplate basin formationaccompanied by tholeiitic basalts may have occurred

    in a leaky transform environment (e.g. Herve et al.,

    1995). The basin was later inverted in Miocene

    Pliocene (?) times when less oblique convergence

    took place. Less oblique convergence could have led

    to contraction and overthrusting of the Patagonian

    Batholith over the basinal deposits.

    During the Pliocene, subduction of the Chile Ridge

    must have played a more significant role in the

    tectonics of the southern Andes than previously rec-

    ognized. The series of ridge segments that havecollided with the continent at about the same latitude

    from 6 Ma have probably enhanced the contractional

    component of dextral-oblique transpressional defor-

    mation, close to the ridge indenter. Dextral-strike slip

    deformation documented a few hundred kilometers to

    the north at 42jS (Cembrano et al., 1996) and dextral

    transtensional deformation recorded at 6 Ma along

    east northeast trending shear zones at 44jS, are

    kinematically compatible with bulk dextral transpres-

    sional deformation induced by the indenter effect of

    Fig. 10. Models for the tectonic consequences of ridge collision (a), and oblique subduction (b), in the tectonics of plate boundaries. (Tapponier

    and Molnar, 1976; Nelson et al., 1994; Beck, 1991).

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314310

  • 7/27/2019 Cenozoico Transpresional Nazca

    23/26

    the Chile Ridge. Furthermore, coeval zones of obli-

    que-reverse slip and strike-slip ductile deformation

    zones of similar orientation between 42jS and 47jS

    suggest that overall dextral transpression at the

    Nazca South America plate boundary zone has been

    partitioned into zones of contraction-dominated and

    strike-slip dominated kinematics within the magmatic

    arc. In contrast, the forearc and foreland regions haveremained nearly undeformed (Figs. 8 and 11).

    5. Conclusion

    The transect across part of the plate boundary zone

    at Puyuhuapi (44jS) revealed several centimeter-to-

    hundreds-of-meters wide north-east trending ductile

    shear zones that affect Paleozoic metamorphic rocks,

    mid-Tertiary stratified rocks and Miocene plutonic

    rocks. Some of these shear zones, e.g. the Puyuhuapi

    and Queulat shear zones, roughly define the eastern

    boundary of the Miocene belt of the North Patagonian

    Batholith. Steeply dipping meter-wide mylonite zones

    display dextral oblique reverse sense of shear in

    the plutons whereas wider, lower strain shear zones,

    show top-to-the-east ductile shear kinematics. Two40

    Ar

    39

    Ar step heating analyses on fine-grained bio-tite separates from mylonitic zones within Miocene

    plutons yielded 3.8F0.2 and 4.2F0.2 Ma ages. These

    were interpreted as the time of a regional Pliocene

    dextral transpressional deformation event.

    The Aysen transect (45jS) revealed northeast-trend-

    ing ductile shear zones developed within Paleozoic

    metamorphic rocks, Miocene plutons and mid-Terti-

    ary metasedimentary metavolcanic rocks. Dextral,

    high-strain shear zones were observed in the Paleo-

    zoic basement at Islas Cinco Hermanas. These rocks

    Fig. 11. Cartoon of the Nazca South America plate boundary zone showing how overall heterogeneous transpressional deformation arising

    from oblique convergence and ridge subduction has been accommodated during the last 6 Ma. Discrete zones of dextral-reverse shear, dextral

    strike-slip and margin-orthogonal contraction are found along and across the plate boundary, concentrated in the magmatic arc. The component

    of plate boundary contraction is enhanced close to the Chile Ridge collision zone. Dextral strike-slip deformation is favored as the distance from

    the collision zone increases. Convergence vector (shown with arrow) has been slightly oblique with respect to the orthogonal to the trench over

    the last 6 Ma.

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 311

  • 7/27/2019 Cenozoico Transpresional Nazca

    24/26

    were analyzed by two methods: (i) conventional step

    heating (on a muscovite separate) and (ii) laser heating

    (in situ on bands of fine-grained mica). The data

    for these rocks suggest that recrystallization accom-panying deformation occurred principally at about 4

    4.5 Ma. In general, this later event is more readily

    resolved by the laserprobe method than it is by conven-

    tional step heating.

    Similar40Ar 39Ar dates (4 4.5 and 6 Ma) obtained

    from ductile shear zones of different kinematics (rang-

    ing from pure strike-slip to pure contractional) strongly

    suggest that bulk transpressional deformation has been

    partitioned in a complex way along and across the

    southern Andes magmatic arc during Pliocene.

    The increased contractional component of overall

    transpressional deformation in the southern transects,

    close to the Pliocene and present-day positions of the

    triple junction, suggests a strong causal link between

    consecutive episodes of Pliocene ridge collision and

    continental margin tectonics. Oblique subduction, a

    process that has taken place since 48 Ma, has been a

    driving mechanism for long-term dextral transpression

    along the continental margin and cannot be causally

    separated from ridge collision. Both ridge collision

    and oblique subduction produce bulk transpressional

    deformation along the leading edge of the continent.

    The contractional/dextral-oblique kinematics ofregional scale shear zones at both the western and

    eastern boundaries of the Miocene plutonic belt of the

    Patagonian Batholith suggest Pliocene bulk transpres-

    sional continental deformation has been accommo-

    dated through coeval thrusting, oblique-slip and

    strike-slip deformation. These orogen-parallel crustal

    shear zones have produced differential exhumation of

    Miocene plutons and wallrocks with respect to rela-

    tively undeformed forearc and foreland regions.

    Because the presently exposed intra-arc shear zones

    formed atf

    350j

    C, at least 7 km of rock must havebeen eroded away during the last 4 Ma (for a thermal

    gradient of 50 jC/km). This gives an average exhu-

    mation rate in excess of 1.7 mm/year for the Miocene

    plutons and deformed wallrocks.

    Acknowledgements

    The present work is part of the first authors PhD

    thesis at Dalhousie University, Nova Scotia. A Killam

    fellowship funded the stay of JC in Canada. The

    Chilean National Science Foundation (FONDECYT)

    funded fieldwork and laboratory analyses through

    grants 1950497 to JC and 2960019 to AS. FranciscoHerve (Universidad de Chile), Liz Schermer (Western

    Washington University) and Bill McClelland (Uni-

    versity of California) participated in the early stages of

    the fieldwork. Marcos Zentilli, Becky Jamieson and

    Nick Culshaw (Dalhousie University) read and made

    important suggestions and comments that signifi-

    cantly improved the content and scope of this paper.

    Dave Prior (University of Liverpool), external exam-

    iner of Cembranos (1998) thesis, is thanked for his

    enthusiasm and encouraging discussion. C. Androni-

    cos and D. Cunningham made important suggestions

    to an early version of this paper. Editor K. Hodges,

    Eric Nelson and an anonymous reviewer are thanked

    for their thorough revisions and suggestions.

    References

    Barrientos, S.E., Acevedo, P., 1992. Seismological aspects of the

    19881989 Lonquimay (Chile) volcanic eruption. Journal of

    Volcanology and Geothermal Research 53, 73 87.

    Bartholomew, D.S., Tarney, J., 1984. Crustal extension in the south-

    ern Andes (45j 46jS). In: Kokelaar, B.P., Howells, M.F.,

    Roach, R.A. (Eds.), Volcanic Processes in Marginal Basins. Spe-

    cial Publication, Geological Society of London, pp. 195205.

    Beck, M.E., 1991. Coastwise transport reconsidered: lateral dis-

    placements in oblique subduction zones, and tectonic conse-

    quences. Physics of the Earth Planetary Interiors 68, 18.

    Berthe, D., Choukroune, P., Jegouzo, P., 1979. Orthogneiss, mylon-

    ite and non-coaxial deformation of granites: the example of the

    South Armorican shear zone. Journal of Structural Geology 1,

    3142.

    Bobenrieth, L., Daz, F., Davidson, J., Portigliati, C., 1983. Mapa

    metalogenico XI region, Sector Norte Continental, comprendido

    entre 45j lat. S y el lmite con la X region. Informe inedito

    3931, SERNAGEOMIN.

    Bourgois, J., Martin, H., Lagabrielle, Y., Le Moigne, J., Frutos Jara,J., 1996. Subduction erosion related to spreading-ridge subduc-

    tion: Taitao peninsula (Chile margin triple junction area). Geol-

    ogy 24, 723726.

    Braun, J., Beaumont, C., 1995. Three-dimensional numerical ex-

    periments of strain partitioning at oblique plate boundaries: im-

    plications for contrasting tectonic styles in the southern Coast

    Ranges, CA, and central South Island, New Zealand. Journal of

    Geophysical Research 100, 1805918074.

    Cande, S.C., Leslie, R.B., 1986. Late Cenozoic tectonics of the

    southern Chile Trench. Journal of Geophysical Research 91,

    471496.

    Cembrano, J., 1998. Kinematics and timing of intra-arc deforma-

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314312

  • 7/27/2019 Cenozoico Transpresional Nazca

    25/26

    tion, southern Chilean Andes. PhD thesis, Dalhousie University,

    Canada.

    Cembrano, J., Herve, F., Lavenu, A., 1996. The LiquineOfqui

    fault zone: a long-lived intra-arc fault system in southern Chile.

    Tectonophysics 259, 5566 (special issue on Andean Geo-dynamics).

    Cembrano, J., Schermer, E., Sanhueza, A., Lavenu, A., 2000. Along

    strike-variations in the nature and timing of deformation along

    an intra-arc shear zone, the Liquine Ofqui fault zone, southern

    Chilean Andes. Tectonophysics 319, 129149.

    Chinn, D.S., Isacks, B.L., 1983. Accurate source depths and focal

    mechanisms of shallow earthquakes in western South America

    and in the New Hebrides island arc. Tectonics 2, 529563.

    Cifuentes, I.L., 1989. The 1960 Chilean earthquakes. Journal of

    Geophysical Research 94 (B1), 665680.

    Dewey, J.F., Lamb, S.H., 1992. Active tectonics of the Andes.

    Tectonophysics 205, 7995.

    Dewey, J.F., Holdsworth, R.E., Strachan, R.A., 1998. Transpression

    and transtension zones. In: Holdsworth, R.E., Strachan, R.A.,

    Dewey, J.F. (Eds.), Continental Transpressional and Transten-

    sional Tectonics. Geological Society, London, Special Publica-

    tion 135, pp. 114.

    Dunlap, J.W., Teyssier, C., McDougall, I., Baldwin, S., 1991. Ages

    of deformation from K/Ar and 40Ar/39Ar dating of white micas.

    Geology 19, 12131216.

    Fitch, T.J., 1972. Plate convergence, transcurrent faults, and internal

    deformation adjacent to southeast Asia and the western Pacific.

    Journal of Geophysical Research 77, 44324460.

    Fitz-Gerald, J.D., Stunitz, H., 1993. Deformation of granitoids at

    low metamorphic grade: I. Reactions and grain size reduction.

    Tectonophysics 221, 269297.

    Flint, S.S., Prior, D.J., Agar, S.M., Turner, P., 1994. Stratigraphicand structural evolution of the Tertiary Cosmelli Basin and its

    relationship to the Chile triple junction. Journal of the Geolog-

    ical Society of London 151, 251268.

    Forsythe, R.D., Nelson, E., 1985. Geological manifestation of ridge

    collision: evidence for the Golfo de Penas, Taitao basin, south-

    ern Chile. Tectonics 4, 477495.

    Fossen, H., Tikoff, B., 1998. Extended models of transpression and

    transtension, and application to tectonic settings. In: Holds-

    worth, R.E., Strachan, R.A., Dewey, J.F. (Eds.), Continental

    Transpressional and Transtensional Tectonics. Geological Soci-

    ety, London, Special Publication 135, pp. 15 33.

    Fossen, H., Tikoff, B., Teyssier, C., 1994. Strain modeling of trans-

    pressional and transtensional deformation. Norsk Geologis k

    Tidsskrift 74, 134145.Godoy, E., Davidson, J., Herve, F., Mpodozis, C., Kawashita, K.,

    1984. Deformacion sobreimpuesta y metamorfismo progresivo

    en un prismade acrecion paleozoico:Archipielago de losChonos,

    Aysen, Chile. Actas IX Congr. Geologico Argentino, Bariloche,

    IV. Asociacion Geologica Argentina, Bariloche Argentina,

    pp. 211 232.

    Grocott, J., Treloar, J., Brown, M., Dallmeyer, R.D., Taylor, G.K.,

    1994. Mechanisms of continental growth in extensional arcs: an

    example from the Andean plate-boundary zone. Geology 22,

    391394.

    Herron, E.M., Cande, S.C., Hall, B.R., 1981. An active spreading

    center collides with a subduction zone; a geophysical survey of

    the Chile margin triple junction. Geological Society of America,

    Memoir 154, 683701.

    Herve, F., 1988. Late Paleozoic subduction and accretion in south-

    ern Chile. Episodes 11, 183188.Herve, F., 1994. The southern Andes between 39j and 44jS lati-

    tude: the geological signature of a transpressive tectonic regime

    related to a magmatic arc. In: Reutter, K.-J., Scheuber, E., Wig-

    ger, P.J. (Eds.), Tectonics of the Southern Central Andes.

    Springer, Berlin, pp. 243248.

    Herve, F., 1998. Late Triassic rocks in the subduction complex of

    Aysen, southern Chile. Event Stratigraphy of Gondwana. Jour-

    nal of African Earth Sciences 10, 224. Abstracts Special Issue

    Gondwana.

    Herve, F., Araya, E., Fuenzalida, J.L., Solano, A., 1979. Edades

    radiometricas y tectonica neogena en el sector costero de Chiloe

    continental, X Region. II Congreso Geologico Chileno, Actas,

    vol. 1, pp. F1F8.

    Herve, F., Pankhurst, R.J., Drake, R., Beck, M., Mpodozis, C.,

    1993. Granite generation and rapid unroofing related to strike-

    slip faulting, Aysen, Chile. Earth and Planetary Science Letters

    120, 375386.

    Herve, F., Pankhurst, R.J., Drake, R., Beck, M., 1995. Pillow meta-

    basal ts in a mid-Tertiary exten siona l basin adjacent to the

    LiquineOfqui fault zone: the Isla Magdalena area, Aysen,

    Chile. Journal of South American Earth Sciences 8, 3346.

    Jarrard, R.D., 1986. Relations among subduction parameters. Re-

    views of Geophysics 24, 217284.

    Jones, R.R., Tanner, P.W.G., 1995. Strain partitioning in transpres-

    sion zones. Journal of Structural Geology 17, 793802.

    Jordan, T.E., Isacks, B.L., Allmendinger, R.W., Brewer, J.A., Ra-

    mos, V.A., Ando, C.J., 1983. Andean tectonics related to geom-etry of subducted Nazca plate. Geological Society of America

    Bulletin 94, 341361.

    Kligfield, D.L., Hunziker, J., Dallmeyer, R.D., Schmid, S., 1986.

    Dating of deformation phases using KAr and ArAr techni-

    ques: results from the northern Apennines. Journal of Structural

    Geology 8, 781 798.

    Lavenu, A., Cembrano, J., 1999. Compressional and transpression-

    al stress pattern for the Pliocene and Quarternary (Andes of

    central and southern Chile). Journal of Structural Geology 21,

    16691691.

    Lister, G.S., Snoke, A.W., 1984. S C mylonites. Journal of Struc-

    tural Geology 6, 617638.

    McCaffrey, R., 1992. Oblique plate convergence, slip vectors, and

    forearc deformation. Journal of Geophysical Research 97,89058915.

    Murdie, R.E., 1994. Seismicity and neotectonics associated with the

    subduction of an active ocean ridge transform system Southern

    Chile. PhD Thesis, University of Liverpool.

    Nelson, E., Forsythe, R., Arit, I., 1994. Ridge collision tectonics in

    terrane development. Journal of South American Earth Sciences

    7 (3/4), 271278.

    Pankhurst, R.J., Herve, F., 1994. Granitoid age distribution and

    emplacement control in the North Patagonian batholith in Aysen

    (44j 47jS). 7j Congreso Geologico Chileno II. Universidad

    de Concepcion, Chile, pp. 14091413.

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314 313

  • 7/27/2019 Cenozoico Transpresional Nazca

    26/26

    Pankhurst, R., Herve, F., Rojas, L., Cembrano, J., 1992. Magmatism

    and tectonics in continental Chiloe, Chile (42j and 42j30VS).

    Tectonophysics 205, 283 294.

    Pankhurst, R.J., Weaver, C.D., Herve, F., Larrondo, P., 1999. Mes-

    ozoic-Cenozoic evolution of the north Patagonian Batholith inAysen, southern Chile. Journal of the Geological Society of

    London 156, 673 694.

    Pardo-Casas, F., Molnar, P., 1987. Relative motion of the Nazca

    (Farallon) and South American plates since Late Cretaceous

    times. Tectonics 6, 233248.

    Passchier, C.W., Simpson, C., 1986. Porphyroclast systems as kine-

    matic indicators. Journal of Structural Geology 8, 831843.

    Ramos, V.A., 1989. Foothills structure in northern Magallanes Ba-

    sin, Argentina. American Association of Petroleum Geologists

    73, 887903.

    Ramos, V.A., Kay, S.M., 1992. Southern Patagonian plateau basalts

    and deformation: backarc testimony of ridge collisions. Tecto-

    nophysics 205, 261282.

    Reddy, S.M., Kelley, S.P., Wheeler, J., 1996. A 40Ar/39Ar laser

    probe study of micas from the Sesia Zone, Italian Alps; impli-

    cations for metamorphic and deformation histories. Journal of

    Metamorphic Geology 14, 493508.

    Saint Blanquat, M., Tikoff, B., Teyssier, C., Vigneresse, J.L., 1998.

    Transpressional kinematics and magmatic arcs. In: Holdsworth,

    R.E., Strachan, R.A., Dewey, J.F. (Eds.), Continental Transpres-

    sional and Transtensional Tectonics. Geological Society, Lon-

    don. Special Publication 135, pp. 327340.

    Sanderson, D., Marchini, R.D., 1984. Transpression. Journal of

    Structural Geology 6, 449458.

    Schermer, E.R., Cembrano, J., Sanhueza, A., 1995. Kinematics and

    timing of intra-arc shear, southern Chile. Geological Society of

    America Abstracts with Programs, A409.Schermer, E.R., Cembrano, J., Sanhueza, A., McClelland, W.C.,

    1996. Geometry, kinematics and timing of intra-arc shear, south-

    ern Chile. International Geological Congress Proceedings, Bei-

    jing, China, vol. 1, p. 214.

    Scheuber, E., Hammerschmidt, K., Friedrichsen, H., 1995. 40Ar 39Ar and Rb Sr analyses from ductile shear zones from the Ata-

    cama fault zone, northern Chile: the age of deformation. Tectono-

    physics 250, 61 87.

    Schreurs, G., Colletta, B., 1998. Analogue modeling of faulting in

    zones of continental transpression and transtension. In: Holds-

    worth, R.E., Strachan, R.A., Dewey, J.F. (Eds.), Continental

    Transpressional and Transtensional Tectonics. Geological Soci-

    ety, London. Special Publications 135, pp. 59 79.SERNAGEOMIN, 1980. 1:1.000.000 scale geologic map of Chile

    Servicio Nacional de Geologia y Mineria, Santiago, Chile.

    Shimamoto, T., 1989. The origin of SC mylonites and a new fault

    zone model. Journal of Structural Geology 11, 5164.

    Simpson, C., 1985. Deformation of granitic rocks across the brittle

    ductile transition. Journal of Structural Geology 5, 503511.

    Simpson, C., Schmid, S.H., 1983. An evaluation of criteria to de-

    duce the sense of movement in sheared rocks. Bulletin of the

    Geological Society of America 94, 12811288.

    Somoza, R., 1998. Updated Nazca (Farallon) South America rela-

    tive motions during the last 40 Ma. Implication for Mountain

    building in the central Andean region. Journal of South Amer-

    ican Earth Sciences 11, 211215.

    Suarez, M.R., De La Cruz, R., 2001. Tectonics in the eastern central

    Patagonian Cordillera (45 30S-47 30S). Journal of the Geolog-

    ical Society of London 157, 9952001.

    Tapponier, P., Molnar, P., 1976. Slip-line theory and large-scale

    continental tectonics. Nature 264, 319324.

    Teyssier, C., Tikoff, B., Markley, M., 1995. Oblique plate motions

    and continental tectonics. Geology 23, 447450.

    Tikoff, B., Greene, D., 1997. Stretching lineation in transpressional

    shear zones: an example from the Sierra Nevada Batholith, CA.

    Journal of Structural Geology 19, 2939.

    Trimby, P.W., Prior, D.J., Wheeler, J., 1998. Grain boundary hier-

    archy development in a quartz mylonite. Journal of Structural

    Geology 20, 917935.

    Tullis, J., Yund, R.A., 1987. Transition from cataclastic flow todislocation creep of feldspar: mechanisms and microstructures.

    Geology 15, 606609.

    West, D.P., Lux, D.R., 1993. Dating mylonitic deformation by the

    ArAr method: an example from the Norumbega Fault Zone,

    Maine. Earth and Planetary Science Letters 120, 221237.

    J. Cembrano et al. / Tectonophysics 354 (2002) 289314314