You are on page 1of 26

Tectonophysics 354 (2002) 289 314 www.elsevier.

com/locate/tecto

Late Cenozoic transpressional ductile deformation north of the NazcaSouth AmericaAntarctica triple junction
Cembrano a,*, Alain Lavenu b,c, Peter Reynolds d, Gloria Arancibia c, Jose pez c, Alejandro Sanhueza c Gloria Lo
a

gicas, Universidad Cato lica del Norte, Avda Angamos 0610, Antofagasta, Chile Departamento de Ciencias Geolo b veloppement en Coope ration (IRD), Institut Franc ais de Recherche Scientifique pour le De Casilla 53390 Correo Central, Santiago 1, Chile c a, Universidad de Chile, Casilla 13518, Correo 21, Santiago, Chile Departamento de Geolog d Department 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 of the southern Andes. n, immediately north of the Nazca South Two east west structural transects were mapped at Puyuhuapi and Ayse America Antarctica triple junction. At Puyuhuapi (44jS), north south striking, high-angle contractional and strike-slip ductile n shear zones developed from plutons coexist with moderately dipping dextral-oblique shear zones in the wallrocks. In Ayse (45 46j), 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, obliqueslip 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 n transect, 45jS) and on mylonitic plutonic rocks with contractional deformation (3.8F0.2 muscovite biotite aggregates, Ayse 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 (45jS) and at the Queulat shear zone at 44jS. Published dates for the undeformed 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

Corresponding author. Fax: +56-55-355977. E-mail address: jcembrano@socompa.ucn.cl (J. Cembrano).

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

290

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

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

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 deformation arising from oblique convergence is accomplished 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 Australian Pacific 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 western 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-dominated 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 kinematic framework has been the almost complete lack of detailed structural and thermochronological studies in the northern Patagonian region. With the exception of regional geology and petrology-oriented work (e.g. Bobenrieth et al., 1983; Bartholomew and et al., 1995), there has been no Tarney, 1984; Herve 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 Cenozoic intra-arc fault zone, which has been assumed to accommodate much of the convergent deformation. 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 east west 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) 289314

291

from two transects across the southern Andes, at n latitudes 44j and 46j, the Puyuhuapi and Ayse 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 43jS.

e Ofqui Fig. 1. Regional-scale tectonic setting of the Southern Andes. Northeast-trending, trench-parallel lineaments correspond to the Liquin fault zone (LOFZ). Current position of the Nazca South America Antarctica triple junction is shown. Series 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.

292

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

Fig. 2. Regional-scale geologic map of the southern Chilean Andes between 43jS and 47jS. (modified from SERNAGEOMIN, 1980; , 1994; Herve et al., 1995; Pankhurst et al., 1999). Bobenrieth et al., 1983; Forsythe and Nelson, 1985; Pankhurst and Herve

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

293

1.1. Tectonic and geologic setting An active spreading center, the Chile Ridge, is currently subducting under western South America e Ofqui and a trench-parallel fault system, the Liquin fault zone (LOFZ), has developed within the Cenozoic magmatic arc (Fig. 1). The Cenozoic geodynamic setting of the southern Chilean Andes is well constrained, showing relatively steady right-lateral oblique 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 America 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-orthogonal 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 , 1994; Lavenu and CemCentral Chile (e.g. Herve brano, 1999). The Central Depression is a north south trending belt characterized by numerous islands and fjords composed of basement metamorphic rocks and a younger unit of patchy metasediments and pillow metabasalts that lie closer to the NPB. The latter unit, n Formation, is generally flat-lying west named Traigue of a major north south-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 metasedimentary rock, greenstone and chert. Godoy et al. (1984) suggested that the basement was Late Paleozoic in age based upon field and geochronologic reconnais-

(1988) interpreted the metamorphic basesance. Herve ment as an accretionary prism, deposited and deformed in situ during Late Paleozoic times although much younger ages (Jurassic) have been obtained recently , 1998). Basement rocks are locally deformed (Herve along northeast-striking ductile shear zones, which are meters to hundreds of meters in width. Based upon regional mapping and basic geochronologic dating, the North Patagonian Batholith has previously been divided into three roughly defined, orogen-parallel belts: A western Jurassic to Cretaceous granodioritic belt, a central Miocene belt of granodiorite diorite with few granites, and an eastern mid-Cretaceous belt constituted by monzogranites and et al., granites (Pankhurst et al., 1992, 1999; Herve , 1994), (Fig. 2). 1993; Pankhurst and Herve Bartholomew and Tarney (1984) proposed that the north south trending strip of volcanoclastic and sedi n Formation) located between mentary rock (Traigue 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 signature of the volcanic rocks: tholeiitic rather than calcalkaline in character. Field evidence documenting the geometric and kinematic relationship between the intra-arc basin deposits and the batholith is very limited. They proposed that, during a Miocene compressional event, the basin was inverted and underthrust 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). et al. (1995) proposed an alternative model Herve for the origin of the pillow basalts and associated metasediments. Based on geologic mapping of the n Formation at Isla Magdalena, they suggested Traigue that a pull-apart basin nucleated along a leaky transform fault, represented now by the LOFZ. However, as in the former model, little structural and chronological evidence was provided to relate coeval dextralstrike-slip motion on the fault to basin formation and sedimentation. A major difficulty has been to date the

294

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

metabasalts and metasediments, which for the most part, have yielded poorly constrained Miocene ages et al., 1995). (Herve 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 et than a 1000 km in a north south direction (e.g. Herve al., 1979; Forsythe and Nelson, 1985; Dewey and , 1994; Cembrano et al., 1996). Lamb, 1992; Herve The forearc and foreland regions show very little evidence 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 ChileRidge over the last 14 Ma (Ramos and Kay, 1992). (1994), the absence of foreland According to Herve deformation north of 47jS has resulted from the accommodation of a significant part of the convergent defor e Ofqui fault zone. mation along the Liquin

scales to determine the kinematics and approximate conditions of deformation. Observed kinematic indi et al., 1979), cators include C/S fabrics (e.g. Berthe asymmetric porphyroclast systems (e.g. Passchier and Simpson, 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 microstructures 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 throughout. 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

2. Geometry and kinematics of discrete ductile n transects shear zones of the Puyuhuapi and Ayse 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). Bartholomew and Tarney (1984) previously identified some of these shear zones at the regional scale, during reconnaissance work. However, they provided no data on the geometry of stretching lineations or the kinematics of deformation. For each shear zone, we analyzed ductile structures in mylonitic rocks on the mesoscopic and microscopic

This shear zone deforms mingled diorite and granodiorite belonging to a 16.7 Ma pluton (Rb Sr, whole , 1994) (Figs. 3 and 6). The rock; Pankhurst and Herve plutonic rocks show a northeast-striking, steeply dipping magmatic foliation and down-dip mineral lineation 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 a few 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) 289314

295

296

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

n transect showing major ductile shear zones along with respective stereoplots of foliations Fig. 4. Regional geology and structure of the Ayse (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) 289314

297

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.

cators such as C/S fabrics and asymmetric porphyroclasts 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 recrystallized aggregates of quartz and highly strained, fractured 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 association 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 composed of weakly to moderately deformed metasediments and metavolcanics that probably belong to the n Formation (Bobenrieth et al., 1983) (Figs. 2 Traigue 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. Foliations, defined by flattened conglomerate pebbles and layers of mica and chlorite, strike north-northeast and dip moderately to steeply to the west. Stretching lineations 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 n Formation to the east (Fig. 3). The shear the Traigue zone is characterized by both a penetrative westdipping, steep to subvertical mylonitic foliation and poorly defined down-dip stretching lineation. Kinematic indicators at the outcrop-scale suggest top to the east dextral-oblique thrusting of the metamorphic n Formation. basement over the Traigue

298

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

2.4. Rio Cisnes shear zone This shear zone is oriented fN 60jE and is represented in aerial photographs and satellite images as a f30-km-long lineament following the length of Rio Cisnes (Figs. 2 and 3). In the field, the shear zone affects both the 10 Ma Puyuhuapi granite (Rb Sr, whole rock, et al., 1993) and metamorphic basement wallHerve 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 east northeast and dip moderately to the northwest. The stretching lineation, defined by quartz-ribbons and streaks of muscovite 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 dipslip, normal component of motion. Shear bands, foliation 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. Shimamoto, 1989; Lister and Snoke, 1984). 2.5. Islas Cinco Hermanas shear zone This high strain shear zone developed from metamorphic 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 fabrics, z-shaped folds and sigmoidal porphyroclasts are seen on horizontal surfaces parallel to the lineation and perpendicular to the foliation. Shear sense is consistently dextral. Metamorphic conditions of mylonitization were probably greenschist facies, as documented by 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 n Formation (Figs. 2 and 4). At a rocks of the Traigue regional scale, the Canal Costa represents a morpho-

logic boundary between the so-called Central Depression to the west and the Main Range to the east, n Formation and dominated by outcrops of the Traigue the Patagonian Batholith, respectively. Moreover, west of the channel, volcanic and sedimentary rocks are almost flat-lying whereas they are highly deformed 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. Stretching 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 n Formation. At the plutonic complex onto the Traigue the 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 dynamically 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 accomplished 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) 289314

299

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 protolith to a high-strain mylonite in the middle of a kilometer-wide ductile shear zone in northern Chile. In this study, they found that 40Ar 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. Reddy et al., 1996). In this method, the laser targets mineral grains (or parts thereof), which are interpreted to have been dynamically 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 analysis. 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 hornblende standard, MMhb-1 (assumed age=520F2 Ma). An internal tantalum resistance furnace of the doublevacuum type was used to carry out the step heating. Laserprobe analyses were made with a Nd YAG system operated in the pulsed mode (1 2 kHz). A targeted area on a rock section was fused by moving the sample chamber under the laser beam by means of an x y translation stage. All isotopic analyses were made using a VG 3600 mass spectrometer. Both undeformed 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 conditions (e.g. Kligfield et al., 1986; Dunlap et al., 1991).

3.1. Results of the Ar Ar dating Apparent age data from step heating are reported with their 1r uncertainties; mean ages are provided with 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 Ar Ar 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. 40 Ar 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 summarized 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 (Rb Sr whole rock age, Pankhurst and Herve 1994). Biotite from sample 95JC4, a low strain variety of mylonitic granodiorite diorite, 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, 20.2F0.2 Ma, relates to the early 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 n fjord (Figs. 4 and 6f), Cinco Hermanas in the Ayse 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.

300 Table 1 Summary of T (jC)

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

40

n transects Ar 39Ar analytical data for low strain and high strain rocks of the Puyuhuapi and Ayse
39

mV

Ar

39

Ar%

Age (Ma)F1r

%ATM

37

Ar/39Ar

36

Ar/40Ar

39

Ar/40Ar

%IIC 12.75 16.28 18.44 31.51 59.86 81.72 177.89 47.30 9.43 7.13 4.57 4.97 5.52 3.13 55.31 8.30

95-JC-1 Biotite. Low strain mingled granodiorite diorite. Puyuhuapy transect 550 41.5 3.0 128.3F3.8 960.0 600 33.8 2.4 56.9F4.7 308.0 650 41.4 2.9 25.9F2.7 167.0 700 79.5 5.7 5.8F1.3 124.0 750 89.9 6.5 2.1F0.1 115.0 800 77.6 5.6 1.3F1 110.0 850 82.9 6.0 0.4F0.7 95.0 900 84.1 6.0 1.2F0.7 84.5 950 136.1 9.8 3.1F0.4 57.1 975 140.8 10.1 3.2F0.3 46.1 1000 160.2 11.5 3.6F0.2 33.2 1025 132.8 9.6 3.5F0.3 31.8 1050 105.2 7.6 3.5F0.3 30.9 1100 156.8 11.3 3.9F0.2 28.8 1200 17.5 1.2 1.4F3.2 105.0 1450 1.6 0.1 104.4F225.2 103.0 Mean age (950 1100 jC)=3.5F0.2 Ma; J=0.00232F0.0000232 (1%)

67.57 32.82 15.91 5.80 3.88 3.37 2.40 1.91 0.94 0.72 0.53 0.55 0.61 0.39 2.39 33.92

0.031041 0.010247 0.005628 0.004197 0.003871 0.003713 0.003192 0.002839 0.001926 0.001561 0.001131 0.001091 0.001068 0.000978 0.003493 0.003492

0.276388 0.151868 0.108746 0.178930 0.308042 0.337276 0.474628 0.494858 0.560082 0.687272 0.748429 0.791464 0.796960 0.742279 0.170124 0.001414

95-JC-1 Hornblende. Low strain mingled granodiorite diorite. Puyuhuapi transect 650 10.1 3.3 62.5F10.8 88.5 2.83 750 12.3 4.0 15.2F7.6 94.7 2.19 850 13.4 4.4 2.9F4.8 97.0 2.88 950 42.6 14.1 19.5F2.1 80.2 9.99 975 44.0 14.5 21.7F1.5 64.3 12.98 1000 49.0 16.2 17.5F1.3 60.8 13.18 1025 19.1 6.3 14.0F2.7 66.8 11.38 1050 8.2 2.7 14.2F5.3 75.3 10.24 1075 6.3 2.0 10.0F9.0 91.2 13.15 1100 7.3 2.4 15.0F9.2 90.9 15.88 1125 12.4 4.1 18.4F6.2 88.8 15.90 1150 12.0 3.9 27.0F5.8 84.1 15.80 1200 33.9 11.2 31.5F2.7 73.7 15.71 1250 11.8 3.9 21.9F6.4 88.0 15.88 1350 13.7 4.5 12.7F8.3 96.7 15.88 1450 5.0 1.6 53.4F35.2 103 14.03 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 1634.0 600 10.0 0.7 116.6F11.8 1634.0 650 12.9 1.0 56.2F8.2 289.0 700 32.4 2.5 11.0F3.2 137.0 750 106.8 8.3 0.3F0.8 103.0 800 273.9 21.3 1.7F0.2 39.8 850 268.2 20.9 1.7F0.1 24.1 900 157.0 12.2 1.3F0.2 39.7 950 60.6 4.7 0.5F0.6 124.0 975 35.7 2.7 1.5F0.6 374.0 1000 48.9 3.8 0.1F0.4 100.0 1025 56.8 4.4 0.7F0.3 62.3 1050 52.6 4.1 0.5F0.3 69.5 1100 109.0 8.5 1.8F0.2 39.8

0.002999 0.003209 0.003293 0.002721 0.002194 0.002087 0.002315 0.002603 0.003115 0.003095 0.003017 0.002856 0.002503 0.002989 0.003275 0.003498

0.007514 0.014355 0.041715 0.041969 0.067942 0.092551 0.097334 0.071654 0.036222 0.024947 0.025098 0.024399 0.034417 0.022554 0.010757 0.002658

1.58 4.63 30.37 16.61 19.48 24.39 26.10 23.16 41.87 34.08 28.00 19.24 16.50 23.58 40.15 7.46

71.00 60.12 31.42 7.72 1.66 0.43 0.32 0.47 1.27 1.97 1.22 0.89 0.83 0.20

0.043443 0.036556 0.009248 0.004587 0.003469 0.001367 0.000865 0.001381 0.003841 0.006809 0.003034 0.002059 0.002234 0.00131

0.387645 0.364203 0.132213 0.140649 0.593526 1.394780 1.719379 1.747207 1.877213 3.500425 2.374611 1.846029 1.837870 1.315038

12.82 12.81 15.80 21.76 192.52 7.76 5.82 10.91 88.76 42.98 12370.68 39.47 44.83 3.51

J. Cembrano et al. / Tectonophysics 354 (2002) 289314 Table 1 (continued ) T (jC) mV


39

301

Ar

39

Ar%

Age (Ma)F1r

%ATM

37

Ar/39Ar

36

Ar/40Ar

39

Ar/40Ar

%IIC 5.22 15.31

95-JC-2 Biotite. Low strain mingled granodiorite diorite. Puyuhuapi transect 1200 37.1 2.8 1.7F1.6 87.8 1450 0.9 0.0 66.2F463.1 101.0 Mean age (800 900 jC)=1.6F0.2 Ma; J=0.00232F0.0000232 (1%)

0.28 36.56

0.002891 0.003419

0.289009 0.000922

95-JC-4 Biotite. Low strain mingled granodiorite diorite. Puyuhuapi transect. Puy. Quarry shear zone 550 31.7 2.2 98F6.8 417.0 61.95 0.013616 600 22.3 1.5 52.2F7.4 240.0 32.65 0.007926 650 28.9 2.0 17.0F5.6 126.0 13.11 0.004261 700 56.6 4.0 1.0F0.3 101.0 3.82 0.003426 750 84.0 5.9 2.4F1.7 88.4 2.10 0.002972 800 69.3 4.9 2.0F1.9 89.5 2.43 0.003004 850 84.5 6.0 3.6F1.2 69.5 1.72 0.002334 900 77.2 5.5 3.9F1.1 59.0 1.43 0.001988 950 102.7 7.3 4.9F0.8 48.1 0.91 0.001621 975 125.7 8.9 5.1F0.5 29.9 0.71 0.001028 1000 204.8 14.6 5.3F0.3 18.1 0.34 0.000624 1025 225.6 16.1 5.6F0.2 11.0 0.24 0.000386 1050 150.2 0.7 5.4F0.2 15.1 0.31 0.000525 1100 95.8 6.8 5.1F0.5 28.5 0.46 0.000965 1200 35.8 2.5 3.5F2.4 77.8 1.27 0.002574 1450 4.7 0.3 30.0F46.9 104.0 14.31 0.003515 Mean age (950 1100 jC)=5.3F0.3 Ma; J=0.00232F0.0000232 (1%) 95-JC-4 hornblende. Low strain mingled granodiorite diorite. Puyuhuapi transect. Quarry shear zone 650 18.1 5.8 49.9F8.0 124.0 37.38 0.004218 750 15.2 4.8 29.4F9.8 109.0 25.49 0.003707 850 14.1 4.5 16.5F7.6 107.0 20.37 0.003633 950 14.8 4.7 14.0F6.0 110.0 18.72 0.003713 975 9.7 3.1 13.5F7.7 110.0 25.13 0.003716 1000 13.4 4.3 1.7F5.7 101.0 25.99 0.003413 1025 30.5 9.7 13.5F2.7 85.6 24.43 0.002894 1050 59.7 19.0 25.7F1.9 76.3 19.54 0.002582 1100 18.6 5.9 25.2F3.3 64.8 18.46 0.002200 1125 4.6 1.4 18.3F10.7 81.0 24.55 0.002718 1150 6.2 1.9 26.5F8.3 74.2 25.56 0.002503 1175 12.9 4.1 34.7F4.8 68.1 22.51 0.002304 1200 11.9 3.8 38.4F5.2 65.9 22.33 0.002229 1250 38.5 12.3 37.5F2.1 63.5 21.07 0.002150 1350 34.1 10.8 37.2F2.4 63.7 21.83 0.002158 1450 10.1 3.2 58.0F7.3 69.3 23.52 0.002335 Mean age (1175 1350 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 86.5 600 7.1 0.3 15.9F12.4 78.1 650 36.4 1.9 15.5F3.3 63.6 700 108.0 5.8 14.8F0.9 41.5 750 204.6 11.1 14.6F0.3 16.7 800 310.8 16.9 14.3F0.2 7.5 850 196.9 10.7 14.4F0.2 7.8 900 102.9 5.6 14.2F0.4 15.3 950 107.3 5.8 14.3F0.6 20.7

0.132815 0.109694 0.065689 0.080565 0.196596 0.214004 0.345849 0.427411 0.432076 0.558256 0.62939 0.644185 0.637975 0.561897 0.256589 0.006162

16.38 17.81 23.51 132.82 27.22 38.11 15.06 11.66 5.90 4.41 2.05 1.36 1.84 2.85 11.42 14.21

0.021157 0.014044 0.019659 0.030779 0.032562 0.033072 0.043870 0.038096 0.057060 0.041820 0.039426 0.037658 0.036261 0.040068 0.040106 0.021646

21.42 25.81 37.64 41.23 57.31 504.01 57.80 24.98 23.96 43.30 31.63 21.64 19.48 18.81 19.62 14.05

36.29 6.47 1.34 0.47 0.23 0.14 0.26 0.71 0.70

0.002733 0.002558 0.002120 0.001390 0.000558 0.000252 0.000265 0.000518 0.000695

0.027781 0.055596 0.096896 0.163844 0.236413 0.267607 0.264022 0.244703 0.228504

64.46 13.08 2.79 1.03 0.50 0.31 0.58 1.60 1.59

(continued on next page)

302 Table 1 (continued ) T (jC) mV


39

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

Ar

39

Ar%

Age (Ma)F1r

%ATM

37

Ar/39Ar

36

Ar/40Ar

39

Ar/40Ar

%IIC 1.75 0.97 0.84 1.76 82.78

95-JC-6 Biotite. Undeformed granodiorite. Puyuhuapi transect. Seno Queulat 1000 104.9 5.7 14.2F0.5 17.2 1050 209.7 11.4 14.2F0.4 11.2 1100 287.1 15.6 14.3F0.2 8.4 1200 150.3 8.2 14.3F0.3 14.1 1450 3.0 0.1 250.4F47.7 53.7 Mean age (650 1200 jC)=14.4F0.3 Ma; J=0.00232F0.0000232 (1%)

0.77 0.43 0.37 0.78 454.86

0.000582 0.000380 0.000286 0.000482 0.001918

0.240189 0.258885 0.265203 0.247193 0.007100

95-JC-6 Hornblende. Undeformed granodiorite. Puyuhuapi transect. Seno Queulat 650 18.0 2.1 16.0F3.9 79.1 750 61.6 7.3 11.3F1.0 54.8 850 61.5 7.3 10.9F0.9 47.1 900 29.5 3.5 12.7F1.6 50.3 950 33.0 3.9 15.5F2.0 61.7 975 51.1 6.0 17.5F1.4 56.3 1000 225.6 26.8 15.2F0.4 47.0 1025 152.1 18.0 14.3F0.5 28.2 1050 59.9 7.1 14.3F0.8 23.3 1075 40.1 4.7 12.9F1.0 32.2 1100 42.2 5.0 12.9F1.3 47.5 1125 23.9 2.8 12.9F2.6 66.9 1150 7.9 0.9 7.8F8.0 91.9 1200 6.5 0.7 0.4F11.9 99.0 1250 4.3 0.5 5.7F19.4 102.0 1350 13.4 1.5 1.2F8.7 99.0 1450 10.0 1.1 8.5F15.8 98.6 Mean age (1000 1125 jC)=14.4F0.6 Ma; J=0.00232F0.0000232 (1%) 95-JC12 Biotite. Undeformed granodiorite. Puyuhuapi transect. Seno Queulat 550 6.8 0.4 115.6F12.5 544.0 600 15.7 0.9 18.6F5.2 129.0 650 57.6 3.5 9.8F1.4 76.4 700 196.4 12.0 13.4F0.3 36.6 750 207.8 12.7 13.2F0.2 13.2 800 283.2 17.3 13.5F0.1 7.9 850 140.5 8.6 13.1F0.2 10.4 900 90.2 5.5 12.8F0.4 15.4 950 79.4 4.8 12.6F0.4 19.9 975 51.4 3.1 12.6F0.5 15.8 1000 86.0 5.2 13.1F0.3 14.0 1025 79.9 4.8 13.3F0.3 9.3 1050 94.9 5.8 13.7F0.2 7.1 1100 197.8 12.1 13.9F0.2 8.9 1200 43.3 2.6 13.2F0.9 34.9 1450 1.2 0.0 96.8F314.6 102.0 Mean age (850 1200 jC)=13.3F0.2 Ma; J=0.00232F0.0000232 (1%) 96GA01 Biotite. Mylonite. 550 3.9 600 12.2 650 30.1 700 49.2 750 49.3 800 48.1 Puyuhuapy transect. Puyuhuapi Quarry shear zone 0.3 32.7F20.6 92.0 1.0 17.9F9.1 94.4 2.5 11.6F4.0 93.6 4.2 6.8F2.2 92.7 4.2 6.6F1.8 92.0 4.1 9.7F2.6 92.9

0.68 0.23 0.46 1.60 5.19 8.24 9.23 9.07 6.49 3.65 5.63 9.81 12.08 12.65 12.56 11.29 2.44

0.002686 0.001862 0.001606 0.001730 0.002115 0.001936 0.001609 0.001030 0.000926 0.001172 0.001668 0.002314 0.003140 0.003390 0.003467 0.003373 0.003342

0.054145 0.166933 0.201649 0.162285 0.102282 0.102916 0.144158 0.204592 0.215059 0.213963 0.166737 0.104947 0.042667 0.023506 0.015287 0.015062 0.006368

1.38 0.65 1.36 4.04 10.81 15.24 19.55 20.37 14.55 9.04 13.99 24.31 49.06 840.20 69.57 290.56 9.12

69.12 19.28 3.23 0.54 0.39 0.25 0.40 0.62 0.56 0.81 0.41 0.24 0.03 0.00 0.14 10.89

0.016499 0.004327 0.002568 0.001231 0.000443 0.000268 0.000351 0.000521 0.000665 0.000540 0.000468 0.000308 0.000230 0.000290 0.001140 0.003449

0.145725 0.065755 0.099398 0.196102 0.272260 0.282850 0.281216 0.270966 0.262184 0.271966 0.269544 0.279619 0.279282 0.272419 0.202711 0.001018

14.89 31.57 10.50 1.29 0.95 0.60 0.98 1.56 1.44 2.07 1.01 0.58 0.07 0.02 0.35 2.92

0.07 0.01 0.02 0.11 0.02 0.00

0.003107 0.003193 0.003168 0.003134 0.003113 0.003144

0.010492 0.013503 0.023668 0.046644 0.051777 0.031672

0.08 0.03 0.07 0.57 0.13 0.02

J. Cembrano et al. / Tectonophysics 354 (2002) 289314 Table 1 (continued ) T (jC) mV


39

303

Ar

39

Ar%

Age (Ma)F1r

%ATM

37

Ar/39Ar 0.03 0.03 0.01 0.02 0.07 0.01 0.28 2.32 2.27

36

Ar/40Ar

39

Ar/40Ar

%IIC 0.15 0.16 0.12 0.21 0.59 0.12 0.73 1.39 0.49

96GA01 Biotite. Mylonite. Puyuhuapy transect. Puyuhuapi Quarry shear zone 850 81.2 6.9 7.3F1.6 91.6 900 88.0 7.5 6.3F0.8 79.5 950 170.6 14.6 5.0F0.4 74.4 1000 263.8 22.6 4.1F0.2 51.6 1050 213.1 18.3 4.2F0.1 29.5 1100 143.1 12.3 4.4F0.2 24.9 1150 8.7 0.7 12.9F1.3 24.5 1200 1.3 0.1 60.6F24.6 64.4 1450 0.7 0.0 206.6F163 87.9 Mean age (1000 1100 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 95.5 700 56.8 6.5 6.7F2.2 94.6 750 38.2 4.4 5.1F1.9 94.2 800 42.7 4.9 6.9F2.0 92.9 850 65.8 7.5 9.2F1.6 90.3 900 60.8 7.0 8.9F0.8 71.4 950 99.3 11.4 4.4F0.5 74.6 1000 187.4 21.5 3.9F0.2 49.4 1050 176.8 20.3 3.8F0.2 40.6 1100 103.9 11.9 3.5F0.3 58.1 1150 18.0 2.0 5.0F1.8 79.5 1200 3.0 0.3 5.5F14.6 96.1 1450 2.9 0.3 13.4F34.5 97.9 Mean age (1000 1100 jC)=3.8F0.3 Ma. J=0.002434F0.000018 (0.7%)

0.003099 0.002686 0.002514 0.001742 0.000997 0.000840 0.000818 0.002145 0.002965

0.049540 0.140882 0.222908 0.512718 0.730535 0.730962 0.248039 0.024916 0.002414

0.21 0.31 0.26 0.26 0.16 0.12 0.10 0.07 0.36 0.39 1.19 5.33 6.78

0.003231 0.003203 0.003186 0.003144 0.003054 0.002415 0.002524 0.001672 0.001384 0.001970 0.002690 0.003246 0.003311

0.020024 0.034368 0.049435 0.044138 0.045940 0.138857 0.246991 0.559732 0.67938 0.521886 0.175269 0.030209 0.006694

0.74 1.55 1.74 1.24 0.59 0.45 0.76 0.64 3.18 3.75 7.78 32.26 17.01

95GA04 Biotite. Low strain mingled granodiorite diorite. Puyuhuapi transect. Puy. Quarry shear zone 550 2.4 0.2 14.4F13.8 89.3 0.49 0.002990 600 13.4 1.1 4.5F2.9 93.0 0.20 0.003137 650 30.1 2.5 6.1F1.5 88.2 0.12 0.002980 700 52.7 4.5 5.6F0.8 82.5 0.11 0.002786 750 59.2 5.0 5.9F0.7 79.7 0.06 0.002694 800 54.6 4.6 6.3F1.3 89.8 0.11 0.003038 850 81.8 7.0 6.2F0.7 83.2 0.06 0.002813 900 80.7 6.9 4.9F0.4 67.7 0.07 0.002286 950 136.4 11.6 4.6F0.2 53.6 0.05 0.001811 1000 275.4 23.5 4.0F0.1 33.2 0.04 0.001124 1050 235.7 20.1 4.1F0.1 28.8 0.18 0.000978 1100 131.0 11.2 4.3F0.1 27.0 0.08 0.000914 1200 10.3 0.8 3.6F2.7 87.1 1.20 0.002920 1450 4.2 0.3 23.0F14.5 92.3 1.07 0.003117 Mean 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 0.00 600 5.6 0.3 8.6F1.7 44.8 0.11 650 14.0 0.7 7.3F1.2 44.7 0.00 675 17.2 0.9 6.9F1.1 44.8 0.03 700 21.7 1.2 5.1F1.5 66.2 0.03 725 31.8 1.7 5.9F1.0 52.4 0.02 750 43.0 2.3 5.6F0.7 48.6 0.01

0.031890 0.066126 0.083825 0.134880 0.148544 0.069375 0.117719 0.282460 0.433324 0.711898 0.743037 0.732523 0.153202 0.014366

1.15 1.45 0.65 0.69 0.36 0.58 0.35 0.52 0.38 0.35 1.45 0.61 11.08 1.59

0.002385 0.001571 0.001536 0.001538 0.002259 0.001786 0.001658

0.111454 0.272863 0.318981 0.338126 0.278627 0.335482 0.384821

0.00 0.44 0.02 0.18 0.20 0.11 0.08

(continued on next page)

304 Table 1 (continued ) T (jC) 96-GA-26 775 800 825 850 900 950 1000 1100 1200 1450 Mean age mV
39

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

Ar

39

Ar%

Age (Ma)F1r

%ATM

37

Ar/39Ar

36

Ar/40Ar

39

Ar/40Ar

%IIC 0.01 0.01 0.00 0.00 0.00 0.00 0.04 0.02 0.06 0.10

Muscovite. High strain bt-ms schist. Puyuhuapi transect. Puerto Cisnes shear zone 67.2 3.7 5.8F0.7 61.9 0.00 242.4 13.4 6.2F0.2 51.5 0.00 323.1 17.8 6.3F0.1 25.6 0.00 223.4 12.3 6.3F0.1 18.5 0.00 275.6 15.2 6.3F0.1 25.0 0.00 158.7 8.7 6.2F0.3 37.3 0.00 92.6 5.1 6.1F0.5 46.4 0.00 209.0 11.5 6.1F0.3 43.5 0.00 74.3 4.1 5.9F0.7 63.0 0.01 5.9 0.3 64.8F32.5 104.0 0.23 (750 1000 jC)=6.2F0.2 Ma; J=0.00232F0.0000232 (1%)

0.002101 0.001745 0.000869 0.000627 0.000848 0.001266 0.001575 0.001474 0.002139 0.003525

0.273282 0.324493 0.489688 0.535572 0.496512 0.421508 0.362133 0.381153 0.258096 0.002732

n transect. Isla Cinco Hermanas 95-JC-14 Biotite. Undeformed granite. Ayse 550 11.2 0.6 67.8F6.8 356.0 600 43.8 2.5 4.6F1.8 111.0 650 149.8 8.7 4.3F0.5 78.2 700 297.2 17.3 5.8F0.2 42.6 750 366.1 21.3 6.2F0.1 24.6 800 200.6 11.7 5.9F0.2 42.6 850 87.4 5.1 4.7F0.4 40.3 900 67.1 3.9 4.5F0.5 47.2 950 84.7 4.9 5.2F0.5 51.0 975 70.5 4.1 4.9F0.7 66.7 1000 76.0 4.4 5.3F0.4 38.0 1025 70.9 4.1 6.0F0.4 35.4 1050 56.8 3.3 6.4F0.5 38.6 1100 92.0 5.3 6.3F0.5 43.6 1200 35.8 2.0 4.6F1.3 73.2 1450 1.4 0.0 106.8F202.5 103.0 Mean age (700 1200 jC)=5.7F0.2 Ma; J=0.00232F0.0000232 (1%)

39.55 6.14 1.21 0.44 0.29 0.44 0.95 1.13 0.72 0.68 0.66 0.22 0.02 0.06 0.72 27.84

0.011387 0.003754 0.002629 0.001430 0.000825 0.001424 0.001342 0.001561 0.001688 0.002213 0.001251 0.001146 0.001240 0.001430 0.002397 0.003495

0.150647 0.109710 0.206765 0.408531 0.503958 0.400065 0.509703 0.477224 0.387755 0.278840 0.478727 0.443894 0.394752 0.367440 0.237803 0.001509

16.09 42.33 8.77 2.41 1.50 2.36 6.28 7.97 4.43 4.35 3.99 1.20 0.07 0.32 4.92 6.62

n transect. Isla Cinco Hermanas shear zone 95GA17 Muscovite biotite bands. Laser spots on a polished slab. High strain schist. Ayse Spot 1 2 3 4 5 6 7 8 9 10 11 12 Mean age mV
39

Ar

Age (Ma)F2r

%ATM

37

Ar/39Ar

36

Ar/40Ar

39

Ar/40Ar

%IIC 0.01 0.02 0.05 0.06 0.03 0.09 0.10 0.03 0.06 0.08 0.07 0.07

24.7 6.8F0.4 34.5 0.04 44.3 7.7F0.3 39.4 0.05 34.5 6.2F0.5 61.6 0.10 45.2 4.6F0.3 59.4 0.10 40.2 8.0F0.3 38.2 0.09 38.4 5.6F0.4 63.4 0.16 48.7 4.6F0.2 47.1 0.16 67.6 5.9F0.2 40.6 0.07 26.3 7.5F0.4 45.0 0.15 30.9 5.3F0.4 52.0 0.14 26.2 9.2F0.4 25.9 0.19 25.0 8.5F0.5 42.2 0.20 (spots 1 12 )=6.4F0.6 Ma; J=0.000197F0.00002 (10.1%)

0.001169 0.001333 0.002087 0.002010 0.001295 0.002146 0.001594 0.001376 0.001525 0.001761 0.000876 0.001429

0.034083 0.027900 0.021826 0.030686 0.027313 0.023076 0.039999 0.035211 0.025958 0.031881 0.028424 0.024006

n transect. Isla Cinco Hermanas shear zone 95-GA-19 Muscovite. High strain schist. Ayse T (jC) 550 600 mV
39

Ar

39

Ar%

Age (Ma)F1r 6.1F8.1 3.1F0.8

%ATM 92.2 68.3

37

Ar/39Ar

36

Ar/40Ar

39

Ar/40Ar

%IIC 0.21 0.56

6.6 54.0

0.4 3.4

0.04 0.05

0.003119 0.002312

0.053169 0.415424

J. Cembrano et al. / Tectonophysics 354 (2002) 289314 Table 1 (continued) n transect. Isla Cinco Hermanas shear zone 95-GA-19 Muscovite. High strain schist. Ayse T (jC) mV
39

305

Ar

39

Ar%

Age (Ma)F1r

%ATM 37.9 29.7 37.1 13.0 15.6 22.5 13.1 17.7 20.4 22.6 27.1 24.6 43.9 87.6

37

Ar/39Ar

36

Ar/40Ar

39

Ar/40Ar

%IIC 0.41 0.35 0.40 0.32 0.26 0.18 0.14 0.13 0.13 0.08 0.10 0.11 0.29 1.23

650 86.4 5.6 3.4F0.3 675 80.1 5.1 3.8F0.3 700 72.4 4.6 3.4F0.3 725 71.8 4.6 3.8F0.2 750 75.8 4.9 4.0F0.2 775 110.9 7.1 4.9F0.2 800 89.3 5.7 5.7F0.2 825 106.8 6.9 5.0F0.2 850 114.5 7.4 4.9F0.2 900 217.5 14.1 5.5F0.2 950 204.9 13.2 5.5F0.2 1000 138.6 8.9 7.2F0.3 1100 90.9 5.8 8.6F0.5 1200 21.6 1.4 9.8F3.4 Total gas age=5.2F0.2 Ma; J=0.00232F0.0000116 (0.5%)

0.04 0.04 0.04 0.03 0.03 0.02 0.02 0.02 0.02 0.01 0.01 0.02 0.08 0.37

0.001286 0.001010 0.001260 0.000447 0.000533 0.000765 0.000448 0.000603 0.000695 0.000767 0.000918 0.000834 0.001487 0.002965

0.749365 0.755144 0.767762 0.940648 0.877467 0.659609 0.635116 0.682395 0.671934 0.578451 0.552298 0.431066 0.270910 0.052524

n transect. Isla Cinco Hermanas shear zone 95GA19 Muscovite biotite bands. Laser spots on a polished slab. High strain schist. Ayse Spot mV
39

Ar

Age (Ma)F2r

%ATM

37

Ar/39Ar

36

Ar/40Ar

39

Ar/40Ar

%IIC 0.16 0.20 0.14 0.04 0.09 0.01 0.00 0.00 0.08 0.29 0.60 0.11

1 2.0 7.4F6.4 75.1 0.38 2 1.9 4.9F4.7 74.8 0.34 3 3.7 4.7F1.8 63.0 0.23 4 21.3 4.6F0.3 43.7 0.06 5 20.1 4.2F0.6 74.6 0.13 6 19.2 4.0F0.4 63.1 0.01 7 11.1 4.9F0.9 67.2 0.01 8 10.5 4.4F0.9 68.8 0.01 9 5.4 5.9F7.1 91.4 0.17 10 2.9 4.2F5.2 85.4 0.42 11 1.3 13.9F64.2 93.9 2.34 12 9.5 5.3F1.7 79.6 0.21 Mean age (spots 1 12)=4.7F1.0 Ma; J=0.000197F0.00002 (10.1%)

0.002541 0.002530 0.002130 0.001480 0.002525 0.002136 0.002273 0.002327 0.003093 0.002891 0.003179 0.002696

0.011825 0.017946 0.027820 0.042544 0.021266 0.032294 0.023639 0.024947 0.005105 0.011999 0.001532 0.013489

Analysis are from bulk mineral separates unless otherwise indicated (laserspots on oriented slabs). Mean ages are reported with a 2r error 37 Ar 39Ar, 36Ar 40Ar and 40Ar 39Ar ratios are corrected for mass spectrometer discrimination, interfering isotopes and system blanks. %IIC=interfering isotopes correction.

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 temperature 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 contractional 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 temperatures following intrusion of the Puerto Cisnes pluton (at ca. 10 Ma). More likely, it dates solid-state deformation and recrystallization of muscovite because the rocks were deformed under greenschist facies conditions, and the muscovite grains define extensional crenulation cleavage indicating dextral transtensional deformation. Samples 95GA17 and 95GA19 are from high strain quartz mica schists collected from Islas Cinco n fjord (Fig. 4). A fine-grained Hermanas in the Ayse muscovite separate from 95GA19 yielded a spectrum that has ages generally increasing from a low of f3 4 Ma to a high of f8 Ma (Fig. 7d). This may result

306

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

Fig. 6. 40Ar 39Ar apparent age spectra for undeformed (low strain) samples of the two transects. Samples 95JC6, 95JC4, 95JC1, 95JC12 are n transect. Mean and plateau (for 95JC6 Bt) ages as discussed in text are indicated. from the Puyuhuapi transect, sample 95JC14 is from the Ayse

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

307

Fig. 7. 40Ar 39Ar apparent age spectra for samples of high strain rocks of the two transects. Samples 96GA01, 96GA03 and 96GA26 are from n transect. Plateau ages as discussed in text are indicated. the Puyuhuapi transect, samples 95GA17 and 95GA19 are from the Ayse

308

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

from the partial resetting of old muscovite grains by a high strain deformation event at f3 4 Ma. Otherwise, the spectrum might represent the mixture of two populations of grains: (i) a relict population represented by older ages, and (ii) a recrystallized population 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 abundance. For 95GA19, many of the spot/area analyses yielded relatively small amounts of gas (i.e. <f10 mV 39Ar), and consequently, imprecise age data. However, the three analyses (4, 5, 6) that produced the largest amounts of gas did give consistent ages (range: 4.0 4.6 Ma; mean: 4.3 F0.3 Ma). For 95GA17, the gas yields were consistently high and all the apparent ages are precise. These range from a low of 4.6F0.3 Ma (analyses 4, 7) to 9.2F0.3 Ma (analysis 11). We interpret the 95GA19 dates as the time of recrystallization to a finer-grained aggregate during high strain deformation. For 95GA17, we suggest that the older ages reflect the isotopic signature of relic muscovite biotite grains, whereas newly crystallized mica grains control the younger ages.

4. Discussion Many authors have speculated that the southern Andes plate boundary zone records a history of transpressional deformation during most of the Cenozoic (e.g. Forsythe and Nelson, 1985; Dewey and Lamb, , 1994; Nelson et al., 1994; Cembrano et 1992; Herve al., 1996). However, there have been no systematic field and thermochronological studies to date that document those speculative ideas. Previous studies have also suggested that Cenozoic deformation of the overriding South American plate has concentrated , 1994). mainly along the magmatic arc (e.g. Herve Although this seems to be reasonably well supported, the actual nature, style and timing of magmatic arc deformation appear to change significantly along and across the strike of the orogen, suggesting different fundamental driving processes (Cembrano et al., 1996; Cembrano et al., 2000). Furthermore, in contrast to the Central Andes, no foreland fold-and-thrust belt has developed in the Patagonian region north of , 1994). 47jS (Ramos and Kay, 1992; Herve Field data presented here document for the first time the geometry, kinematics and timing of deformation in the Cenozoic magmatic arc of the Southern Andes close to the Chile Triple Junction. Dextral strike-slip and contractional to oblique-slip ductile deformation zones affect both the Miocene belt of

n transects. WestFig. 8. Simplified regional cross section showing structural, kinematic and chronological data from the Puyuhuapi and Ayse verging and east-verging dextral-reverse ductile shear zones flank the Miocene belt of the North Patagonian Batholith. The resulting architecture is a regional-scale, transpressional pop-up structure.

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

309

the Patagonian Batholith and wallrocks. The bulk of the transpressional deformation took place at around 4 Ma as documented by 40Ar 39Ar dating on recrystallized biotite and muscovite from several greenschist facies shear zones. A slightly older dextral transtensional 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 n Formation and appear to over the basement/Traigue 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 threedimensional numerical dynamic modeling of transpressional deformation at obliquely convergent plate margins (Braun and Beaumont, 1995). Recent experimental analog modeling of transpressional deformation 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 structure observed 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, concentrates the bulk of deformation by crustal thickening and strike-slip movements giving rise to a topographic relief by double-verging thrust zones as depicted by Braun and Beaumonts numerical model. Ridge collision and oblique subduction have been proposed as alternative driving mechanisms of transpressional deformation at the leading edge of the South American plate. Collision of successive segments 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, oblique subduction is thought to produce overall transpressional 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 dextral 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 north northwest trending short segments of the ridge were consecutively subducted from around 6 Ma to present-day close to the Taitao Peninsula.

310

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

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).

highly dextral-oblique and intraplate basin formation accompanied by tholeiitic basalts may have occurred et al., in a leaky transform environment (e.g. Herve 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 have collided with the continent at about the same latitude from 6 Ma have probably enhanced the contractional component of dextral-oblique transpressional deformation, 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 transpressional deformation induced by the indenter effect of

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

311

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.

the Chile Ridge. Furthermore, coeval zones of oblique-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 have remained nearly undeformed (Figs. 8 and 11).

5. Conclusion The transect across part of the plate boundary zone at Puyuhuapi (44jS) revealed several centimeter-tohundreds-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. Two 40 Ar 39Ar step heating analyses on fine-grained biotite 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. n transect (45jS) revealed northeast-trendThe Ayse ing ductile shear zones developed within Paleozoic metamorphic rocks, Miocene plutons and mid-Tertiary metasedimentary metavolcanic rocks. Dextral, high-strain shear zones were observed in the Paleozoic basement at Islas Cinco Hermanas. These rocks

312

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

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 accompanying 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 conventional step heating. Similar 40Ar 39Ar dates (4 4.5 and 6 Ma) obtained from ductile shear zones of different kinematics (ranging 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 of regional scale shear zones at both the western and eastern boundaries of the Miocene plutonic belt of the Patagonian Batholith suggest Pliocene bulk transpressional continental deformation has been accommodated 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 relatively undeformed forearc and foreland regions. Because the presently exposed intra-arc shear zones formed at f350 jC, at least 7 km of rock must have been eroded away during the last 4 Ma (for a thermal gradient of 50 jC/km). This gives an average exhumation rate in excess of 1.7 mm/year for the Miocene plutons and deformed wallrocks.

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. Francisco (Universidad de Chile), Liz Schermer (Western Herve Washington University) and Bill McClelland (University 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 significantly improved the content and scope of this paper. Dave Prior (University of Liverpool), external examiner of Cembranos (1998) thesis, is thanked for his enthusiasm and encouraging discussion. C. Andronicos 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 1988 1989 Lonquimay (Chile) volcanic eruption. Journal of Volcanology and Geothermal Research 53, 73 87. Bartholomew, D.S., Tarney, J., 1984. Crustal extension in the southern Andes (45j 46jS). In: Kokelaar, B.P., Howells, M.F., Roach, R.A. (Eds.), Volcanic Processes in Marginal Basins. Special Publication, Geological Society of London, pp. 195 205. Beck, M.E., 1991. Coastwise transport reconsidered: lateral displacements in oblique subduction zones, and tectonic consequences. Physics of the Earth Planetary Interiors 68, 1 8. , D., Choukroune, P., Jegouzo, P., 1979. Orthogneiss, mylonBerthe ite and non-coaxial deformation of granites: the example of the South Armorican shear zone. Journal of Structural Geology 1, 31 42. az, F., Davidson, J., Portigliati, C., 1983. Mapa Bobenrieth, L., D nico XI regio n, Sector Norte Continental, comprendido metaloge mite con la X regio n. Informe ine dito entre 45j lat. S y el l 3931, SERNAGEOMIN. Bourgois, J., Martin, H., Lagabrielle, Y., Le Moigne, J., Frutos Jara, J., 1996. Subduction erosion related to spreading-ridge subduction: Taitao peninsula (Chile margin triple junction area). Geology 24, 723 726. Braun, J., Beaumont, C., 1995. Three-dimensional numerical experiments of strain partitioning at oblique plate boundaries: implications for contrasting tectonic styles in the southern Coast Ranges, CA, and central South Island, New Zealand. Journal of Geophysical Research 100, 18059 18074. Cande, S.C., Leslie, R.B., 1986. Late Cenozoic tectonics of the southern Chile Trench. Journal of Geophysical Research 91, 471 496. Cembrano, J., 1998. Kinematics and timing of intra-arc deforma-

Acknowledgements The present work is part of the first authors PhD thesis at Dalhousie University, Nova Scotia. A Killam

J. Cembrano et al. / Tectonophysics 354 (2002) 289314 tion, southern Chilean Andes. PhD thesis, Dalhousie University, Canada. , F., Lavenu, A., 1996. The Liquine Ofqui Cembrano, J., Herve fault zone: a long-lived intra-arc fault system in southern Chile. Tectonophysics 259, 55 66 (special issue on Andean Geodynamics). Cembrano, J., Schermer, E., Sanhueza, A., Lavenu, A., 2000. Along strike-variations in the nature and timing of deformation along e Ofqui fault zone, southern an intra-arc shear zone, the Liquin Chilean Andes. Tectonophysics 319, 129 149. 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, 529 563. Cifuentes, I.L., 1989. The 1960 Chilean earthquakes. Journal of Geophysical Research 94 (B1), 665 680. Dewey, J.F., Lamb, S.H., 1992. Active tectonics of the Andes. Tectonophysics 205, 79 95. 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 Transtensional Tectonics. Geological Society, London, Special Publication 135, pp. 1 14. 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, 1213 1216. Fitch, T.J., 1972. Plate convergence, transcurrent faults, and internal deformation adjacent to southeast Asia and the western Pacific. Journal of Geophysical Research 77, 4432 4460. Fitz-Gerald, J.D., Stunitz, H., 1993. Deformation of granitoids at low metamorphic grade: I. Reactions and grain size reduction. Tectonophysics 221, 269 297. Flint, S.S., Prior, D.J., Agar, S.M., Turner, P., 1994. Stratigraphic and structural evolution of the Tertiary Cosmelli Basin and its relationship to the Chile triple junction. Journal of the Geological Society of London 151, 251 268. Forsythe, R.D., Nelson, E., 1985. Geological manifestation of ridge collision: evidence for the Golfo de Penas, Taitao basin, southern Chile. Tectonics 4, 477 495. Fossen, H., Tikoff, B., 1998. Extended models of transpression and transtension, and application to tectonic settings. In: Holdsworth, R.E., Strachan, R.A., Dewey, J.F. (Eds.), Continental Transpressional and Transtensional Tectonics. Geological Society, London, Special Publication 135, pp. 15 33. Fossen, H., Tikoff, B., Teyssier, C., 1994. Strain modeling of transpressional and transtensional deformation. Norsk Geologisk Tidsskrift 74, 134 145. , F., Mpodozis, C., Kawashita, K., Godoy, E., Davidson, J., Herve n sobreimpuesta y metamorfismo progresivo 1984. Deformacio n paleozoico: Archipie lago de los Chonos, en un prisma de acrecio n, Chile. Actas IX Congr. Geolo gico Argentino, Bariloche, Ayse n Geolo gica Argentina, Bariloche Argentina, IV. Asociacio 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, 391 394. Herron, E.M., Cande, S.C., Hall, B.R., 1981. An active spreading

313

center collides with a subduction zone; a geophysical survey of the Chile margin triple junction. Geological Society of America, Memoir 154, 683 701. , F., 1988. Late Paleozoic subduction and accretion in southHerve ern Chile. Episodes 11, 183 188. , F., 1994. The southern Andes between 39j and 44jS latiHerve tude: the geological signature of a transpressive tectonic regime related to a magmatic arc. In: Reutter, K.-J., Scheuber, E., Wigger, P.J. (Eds.), Tectonics of the Southern Central Andes. Springer, Berlin, pp. 243 248. , F., 1998. Late Triassic rocks in the subduction complex of Herve n, southern Chile. Event Stratigraphy of Gondwana. JourAyse nal of African Earth Sciences 10, 224. Abstracts Special Issue Gondwana. , F., Araya, E., Fuenzalida, J.L., Solano, A., 1979. Edades Herve tricas y tecto nica neo gena en el sector costero de Chiloe radiome n. II Congreso Geolo gico Chileno, Actas, continental, X Regio vol. 1, pp. F1 F8. , F., Pankhurst, R.J., Drake, R., Beck, M., Mpodozis, C., Herve 1993. Granite generation and rapid unroofing related to strike n, Chile. Earth and Planetary Science Letters slip faulting, Ayse 120, 375 386. , F., Pankhurst, R.J., Drake, R., Beck, M., 1995. Pillow metaHerve basalts in a mid-Tertiary extensional basin adjacent to the n, e Ofqui fault zone: the Isla Magdalena area, Ayse Liquin Chile. Journal of South American Earth Sciences 8, 33 46. Jarrard, R.D., 1986. Relations among subduction parameters. Reviews of Geophysics 24, 217 284. Jones, R.R., Tanner, P.W.G., 1995. Strain partitioning in transpression zones. Journal of Structural Geology 17, 793 802. Jordan, T.E., Isacks, B.L., Allmendinger, R.W., Brewer, J.A., Ramos, V.A., Ando, C.J., 1983. Andean tectonics related to geometry of subducted Nazca plate. Geological Society of America Bulletin 94, 341 361. Kligfield, D.L., Hunziker, J., Dallmeyer, R.D., Schmid, S., 1986. Dating of deformation phases using K Ar and Ar Ar techniques: results from the northern Apennines. Journal of Structural Geology 8, 781 798. Lavenu, A., Cembrano, J., 1999. Compressional and transpressional stress pattern for the Pliocene and Quarternary (Andes of central and southern Chile). Journal of Structural Geology 21, 1669 1691. Lister, G.S., Snoke, A.W., 1984. S C mylonites. Journal of Structural Geology 6, 617 638. McCaffrey, R., 1992. Oblique plate convergence, slip vectors, and forearc deformation. Journal of Geophysical Research 97, 8905 8915. 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), 271 278. , F., 1994. Granitoid age distribution and Pankhurst, R.J., Herve n emplacement control in the North Patagonian batholith in Ayse gico Chileno II. Universidad (44j 47jS). 7j Congreso Geolo n, Chile, pp. 1409 1413. de Concepcio

314

J. Cembrano et al. / Tectonophysics 354 (2002) 289314 Schreurs, G., Colletta, B., 1998. Analogue modeling of faulting in zones of continental transpression and transtension. In: Holdsworth, R.E., Strachan, R.A., Dewey, J.F. (Eds.), Continental Transpressional and Transtensional Tectonics. Geological Society, London. Special Publications 135, pp. 59 79. SERNAGEOMIN, 1980. 1:1.000.000 scale geologic map of Chile a y Mineri a, Santiago, Chile. Servicio Nacional de Geologi Shimamoto, T., 1989. The origin of S C mylonites and a new fault zone model. Journal of Structural Geology 11, 51 64. Simpson, C., 1985. Deformation of granitic rocks across the brittle ductile transition. Journal of Structural Geology 5, 503 511. Simpson, C., Schmid, S.H., 1983. An evaluation of criteria to deduce the sense of movement in sheared rocks. Bulletin of the Geological Society of America 94, 1281 1288. Somoza, R., 1998. Updated Nazca (Farallon) South America relative motions during the last 40 Ma. Implication for Mountain building in the central Andean region. Journal of South American Earth Sciences 11, 211 215. Suarez, M.R., De La Cruz, R., 2001. Tectonics in the eastern central Patagonian Cordillera (45 30S-47 30S). Journal of the Geological Society of London 157, 995 2001. Tapponier, P., Molnar, P., 1976. Slip-line theory and large-scale continental tectonics. Nature 264, 319 324. Teyssier, C., Tikoff, B., Markley, M., 1995. Oblique plate motions and continental tectonics. Geology 23, 447 450. Tikoff, B., Greene, D., 1997. Stretching lineation in transpressional shear zones: an example from the Sierra Nevada Batholith, CA. Journal of Structural Geology 19, 29 39. Trimby, P.W., Prior, D.J., Wheeler, J., 1998. Grain boundary hierarchy development in a quartz mylonite. Journal of Structural Geology 20, 917 935. Tullis, J., Yund, R.A., 1987. Transition from cataclastic flow to dislocation creep of feldspar: mechanisms and microstructures. Geology 15, 606 609. West, D.P., Lux, D.R., 1993. Dating mylonitic deformation by the Ar Ar method: an example from the Norumbega Fault Zone, Maine. Earth and Planetary Science Letters 120, 221 237.

, F., Rojas, L., Cembrano, J., 1992. Magmatism Pankhurst, R., Herve , Chile (42j and 42j30VS). and tectonics in continental Chiloe Tectonophysics 205, 283 294. , F., Larrondo, P., 1999. MesPankhurst, R.J., Weaver, C.D., Herve ozoic-Cenozoic evolution of the north Patagonian Batholith in n, southern Chile. Journal of the Geological Society of Ayse 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, 233 248. Passchier, C.W., Simpson, C., 1986. Porphyroclast systems as kinematic indicators. Journal of Structural Geology 8, 831 843. Ramos, V.A., 1989. Foothills structure in northern Magallanes Basin, Argentina. American Association of Petroleum Geologists 73, 887 903. Ramos, V.A., Kay, S.M., 1992. Southern Patagonian plateau basalts and deformation: backarc testimony of ridge collisions. Tectonophysics 205, 261 282. Reddy, S.M., Kelley, S.P., Wheeler, J., 1996. A 40Ar/39Ar laser probe study of micas from the Sesia Zone, Italian Alps; implications for metamorphic and deformation histories. Journal of Metamorphic Geology 14, 493 508. 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 Transpressional and Transtensional Tectonics. Geological Society, London. Special Publication 135, pp. 327 340. Sanderson, D., Marchini, R.D., 1984. Transpression. Journal of Structural Geology 6, 449 458. 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, southern Chile. International Geological Congress Proceedings, Beijing, China, vol. 1, p. 214. Scheuber, E., Hammerschmidt, K., Friedrichsen, H., 1995. 40Ar 39 Ar and Rb Sr analyses from ductile shear zones from the Atacama fault zone, northern Chile: the age of deformation. Tectonophysics 250, 61 87.

You might also like