Sie sind auf Seite 1von 18

Journal of the Geological Society, London, Vol. 163, 2006, pp. 965982. Printed in Great Britain.

Neoproterozoic to Early Palaeozoic events in the Sierra de San Luis: implications


for the Famatinian geodynamics in the Eastern Sierras Pampeanas (Argentina)
NICA G. LO
P E Z D E L U C H I 1,
A N D R E S T E E N K E N 1, S I E G F R I E D S I E G E S M U N D 2, M O
3
2
RO B E RT F R E I & K L AU S W E M M E R
1
Instituto de Geocronologa y Geologa Isotopica, Pabellon INGEIS, Ciudad Universitaria, C1428EHA,
Buenos Aires, Argentina (e-mail: asteenk@gwdg.de)
2
Geoscience Centre of the University of Gottingen, University Gottingen, Goldschmidtstr. 3, 37077 Gottingen, Germany
3
Geological Institute, University of Copenhagen, ster Voldgade 10, DK-1350 Copenhagen, Denmark
Abstract: The application of the sensitive high-resolution ion microprobe (SHRIMP) U/Pb dating technique
to zircon and monazites of different rock types of the Sierra de San Luis provides an important insight into
the provenance and timing of deposition of the sedimentary precursors as well as the metamorphic and
igneous history of the various basement domains. Additional constraints on the Famatinian metamorphic
episode are provided by Pb/Pb stepwise leaching experiments on one staurolite and two garnet separates. The
results indicate that the sedimentary precursors of the Conlara Metamorphic Complex have a maximum age of
c. 590 Ma, whereas the Pringles Metamorphic Complex metasediments appear to be sourced from the
Pampean orogen in the Early Cambrian. Folded xenoliths within the c. 496 Ma El Penon pluton suggest that
the host Conlara Metamorphic Complex underwent a Pampean compression. From a 208 Pb/232 Th monazite age
of 478 Ma for a migmatite from the Nogol Metamorphic Complex, the structural evolution of this basement
complex appears to be entirely post-Pampean. Onset of the Famatinian high-grade metamorphism, between c.
500 Ma and c. 450 Ma, follows a period of crustal extension on the western outboard of Gondwana and might
not be related directly to a Mid-Ordovician accretion of the Cuyania Terrane.

Pampeanas (i.e. the Sierra de Pie de Palo (Casquet et al. 2001)


and the Sierra de Valle Fertil (Fig. 1)) makes it a key locality for
understanding the widely accepted Ordovician tectonic emplacement of the Cuyania Terrane at its present position within the
southwestern margin of Gondwana (Astini et al. 1995; Astini
1998; Astini & Thomas 1999; Quenardelle & Ramos 1999;
Rapalini et al. 1999; Thomas et al. 2002; Thomas & Astini
2003; Astini & Davila 2004).
We provide new time constraints on the early Palaeozoic
evolution of the southwestern margin of Gondwana using sensitive
high-resolution ion microprobe (SHRIMP) techniques and Pb/Pb
stepwise leaching (PbSL) experiments. The SHRIMP technique,
in particular, has proven extremely useful in defining the ages of
multiple events within a single sample. Results presented here
help to define crustal provenances for the metasedimentary rocks
of the various basement domains of the Sierra de San Luis (see
Stewart et al. 2001) as well as an approximation to the emplacement ages of the intrusive rocks. Specific metamorphic minerals
have been subjected to PbSL experiments. Work by Frei &
Kamber (1995) and Frei et al. (1997) has shown that this method
is capable of capturing the timing of metamorphic mineral growth
and therefore it provides another time marker for the fabric
development in poly-deformed metamorphic rocks.
Together with the increasing amount of geochronological data
for the various parts of the Sierras Pampeanas the new results
will help to delineate the crustal domains with respect to their
Pampean or Famatinian history.

The Sierras Pampeanas in central Argentina comprise a number of


morphotectonic terrains whose Cenozoic exhumation is related to
the subhorizontal subduction of the oceanic Nazca plate below the
South American platform (Jordan & Allmendinger 1986). Among
the terrains of the Eastern Sierras Pampeanas (Fig. 1), the most
important are the Sierra de San Luis (Sato et al. 2003b; Steenken
et al. 2004), the Sierras de Cordoba (Rapela et al. 1998b; Martino
2003) and the Sierra de Chepes (Pankhurst et al. 1998). The
metasedimentary and igneous rocks that crop out constitute the
cover of the Pampia Craton (Ramos et al. 1986) or Pampean
Terrane (Rapela et al. 1998b), which collided with the Ro de La
Plata Craton during the Pampean Orogenic Cycle (540520 Ma).
Later accretions to the newly formed western outboard of
Gondwana are the Precordillera (or Cuyania; Ramos et al. 1986)
Terrane during the Famatinian Cycle (500440 Ma) and the
poorly constrained Chilenia Terrane (Ramos & Basei 1997) during
the Achalian Cycle (c. 400 Ma). Different crustal provenances are
proposed for the various basement domains that constitute the
proto-Andean margin of Gondwana, and the different geodynamic
scenarios established west of the Ro de La Plata Craton are still
controversial (see Ramos 1988; Astini 1996; Dalla Salda
et al. 1998; Rapela et al. 1998a,b; Rapela 2000; Acenolaza et al.
2002; Thomas et al. 2002; Thomas & Astini 2003; Steenken
et al. 2004).
The Sierra de San Luis (328109338209S, 658159668209W),
the southern tip of the Eastern Sierras Pampeanas, records an
Ediacaran(?) to Devonian metamorphic and magmatic evolution.
In addition, Early Cambrian sedimentation was suggested for
some of the precursors of the basement domains (Sims et al.
1998; Sollner et al. 2000). Its position between the Late
NeoproterozoicEarly Cambrian Pampean orogen of the Sierras
de Cordoba (Rapela et al. 1998b) and the Western Sierras

Geological setting
The Sierra de San Luis comprises three variable NNESSWstriking basement domains of amphibolite- to granulite-facies
965

966

A. STEENKEN ET AL.

Fig. 1. Simplified sketch map of the protoAndean basement that crops out in
northwestern Argentina (modified from von
Gosen 1998b). The geographical division
into the morphotectonic basement domains
of the Northern and Eastern Sierras
Pampeanas is indicated, as well as the
Western Sierras Pampeanas, developed on
the Cuyania Terrane. ASC, Sierras de
Cordoba; SL, Sierra de San Luis; SCh,
Sierra de Chepes; VF, Sierra de Valle Fertil;
PP, Sierra de Pie de Palo; SF, Sierra
Famatina.

FA M AT I N I A N G E O DY NA M I C E VO L U T I O N O F G O N DWA NA

ortho- and paragneisses, schists, migmatites and amphibolites;


namely, from west to east, the Nogol, Pringles and Conlara
Metamorphic Complexes (Fig. 2 Sims et al. 1997). The subordinate appearance of marbles, calcsilicates and banded iron
formations has been reported for the Nogol Metamorphic
Complex only (Gonzalez et al. 2002b). The domains are
generally separated by an assemblage of metaquartz arenites and
phyllites: the San Luis Formation of Prozzi & Ramos (1988).
Throughout the late Cambrian to Mid Ordovician (i.e. the
Famatinian Cycle) numerous intrusions of (ultra-)mafic to monzogranitic composition were emplaced (Sims et al. 1997, 1998;
Stuart-Smith et al. 1999; von Gosen et al. 2002; Sato et al.
2003a, 2005). The intrusive history terminates in the Devonian
with the emplacement of voluminous discordant granodioritic
and syenogranitic batholiths (Fig. 2) during the Achalian Cycle
(Sims et al. 1997; Lopez de Luchi et al. 2002, 2006; Siegesmund
et al. 2004).
There is continuing discussion about the relative timing of
fabric development within the various basement domains. Relict
NWSE-trending planar fabrics within the Nogol Metamorphic

967

Complex that interfere with the prominent NNESSW foliation


trend of the entire basement complex were used to claim a preFamatinian structural evolution of this western basement domain
(von Gosen & Prozzi 1998; Sato et al. 2001, 2003b; Gonzalez et
al. 2002b, 2004). Although this assumption seemed to be
confirmed by the appearance of banded iron formations and
spinifex textures of olivine in komatiitic metavolcanites (Sato et
al. 2001; Gonzalez et al. 2002b) it was not supported by
geochronological data. Reliable U/Pb zircon and monazite ages
between 458 Ma and 475 Ma (Sato et al. 2003a, 2005) are
exclusively related to the Famatinian magmatometamorphic
evolution.
There is similar controversy concerning the Pringles Metamorphic Complex, which has been subdivided into a amphiboliteto granulite-facies migmatic basement overlain by lower-grade
metamorphic sequences; that is, the mica schist group (von Gosen
& Prozzi 1996) and the San Luis Formation (Prozzi & Ramos
1988). According to von Gosen & Prozzi (1996, 1998) and von
Gosen (1998a,b) the upper sequences experienced an exclusively
Famatinian deformation, whereas the high-grade basement was
deformed for the first time during the Pampean Cycle. New
structural observations (Steenken et al. 2004, 2006) do not
confirm this statement. The inferred sedimentation age of the
protolith indicates a near-exclusive Pampean source (Sims et al.
1998) with most SHRIMP U/Pb zircon core ages at c. 530 Ma.
Early to mid-Cambrian sedimentation was followed by nearisobaric amphibolite- to granulite-facies metamorphism related to
the emplacement (ultra-)mafic rocks in a probable back-arc
environment (Hauzenberger et al. 2001; Steenken et al. 2004).
Alternatively, the sediments that turned into the low-grade
metamorphic units were related to a foreland basin that formed as
a response to the uprising Pampean orogen (Chernicoff & Ramos
2003). Both models are consistent with an andesitic to acidic
contribution found within the clastic material (Lopez de Luchi et
al. 2003). The closure of the basin was the result of the Cuyania
Terrane approaching the western outboard of the Pampean
Terrane. The end of granulite-facies metamorphism at c. 460 Ma
(SHRIMP U/Pb monazite and minor zircon data; Sims et al.
1998) is consistent with U/Pb zircon ages from the basement of
the Sierra de Pie de Palo (Casquet et al. 2001). Dating of igneous
rocks that intrude the complex yielded U/Pb zircon ages of
48  7 Ma in orthogneisses and 478  6 Ma in a felsic segregation in mafic rocks (Sims et al. 1998).
Recent isotopic and geochemical studies have shown a correlation between the Ediacaran Puncoviscana Formation (Jezek et al.
1985) and the metasediments of the Conlara Metamorphic
Complex (Steenken et al. 2004; Zimmermann 2005). A preFamatinian polyphase deformation (Sims et al. 1997; Steenken et
al. 2005) is recorded in this basement complex (Lopez de Luchi
& Cerredo 2001).

Sample description

Fig. 2. Simplified geological map of the Sierra de San Luis (modified


after von Gosen 1998a,b; Steenken et al. 2004). The sample localities for
the discussed geochronological data are indicated.

The granitoids of the Sierra de San Luis exhibit different structural


relations with their hosts as well as distinct internal deformation.
Accordingly, the Ordovician granitoids have been divided in a pre- and
syn-orogenic group relative to the Famatinian cycle (Llambas et al.
1998). Of these, the former are represented by the tonalitic to granodioritic intrusions that are restricted to the San Luis Formation, whereas the
latter form sheet-like granodioritic to granitic plutons that parallel
regional foliations. Recent structural and geochemical considerations
suggest that both suites of Famatinian granitoids intruded the basement
contemporaneously during the late Cambrian to Mid-Ordovician (Steenken et al. 2004, 2006; Lopez de Luchi et al. 2006).

968

A. STEENKEN ET AL.

Granitoids (A 14-02, A 02-02, A 59-01, A 40-01) Among the investigated


samples three exhibit pervasive solid-state fabrics: a leucomonzogranite
(A 14-02) from the Nogol Metamorphic Complex, the northern stock of
the Paso del Rey Granite (A 02-02) from the amphibolite-facies domain
of the Pringles Metamorphic Complex, and the El Penon Granite (A 5901) from the Conlara Metamorphic Complex. The planar fabrics of those
intrusive rocks parallel the regional foliation trend of the host. One
sample taken from the Las Verbenas pluton (A 40-01), which intruded
the western belt of the San Luis Formation (Fig. 2), experienced a
heterogeneous subsolidus overprint of its magmatic structures only. The
relative timing of emplacement of these rocks is indicated from granite
host relationships. In the case of the crustal-derived granites of the
Pringles Metamorphic Complex, such as the Paso del Rey Granite (A 0202), it is evident that those granitoids intruded their metasedimentary
host parallel to its first metamorphic fabric and were subsequently
isoclinally folded during D2 (e.g. Steenken et al. 2006). Similarly, the
tonalites and granodiorites of the San Luis Formation (A 40-01) intruded
the host during its first deformation, although this first phase of
deformation corresponds to the D2 deformation of the high-grade rocks
of the Pringles Metamorphic Complex (Steenken et al. 2006). In contrast,
two phases of deformation within the host of the El Penon Granite were
described preceding the syntectonic emplacement of the pluton (Steenken
et al. 2005). No relative timing of the leucomonzogranites within the
Nogol Metamorphic Complex has been established so far.
The granitoids are composed of variable amounts of quartz, plagioclase
(oligoclaseandesine), K-feldspar, muscovite  biotite  garnet 
epidote/clinozoisite. Accessory minerals are apatite, zircon, titanite and
tourmaline. Microstructures indicate that deformation took place in the
submagmatic state, when rigid plagioclase and K-feldspar laths broke and
the intracrystalline fractures were filled with fine-grained melts of
granitic composition. Continuity of the deformation process below the
solidus but still at high temperatures is indicated by the formation of
anhedral subgrains in plagioclase (Fig. 3a). Cross-hatched twinning of
the K-feldspar denotes to its microcline modification. It frequently shows
flame perthite exsolution lamellae and the formation of myrmekite in
contact with plagioclase. Chessboard-like subgrains in relict quartz grains
(Fig. 3b) attest to deformation temperatures close to the submagmatic
field (Kruhl 1996; Paterson et al. 1998). Quartz recrystallization at lower
temperature is accommodated either by migration processes or static
recrystallization leading to triple junctions. Garnet, found only in A 1402, A 02-02 and A 40-01, appears to be of metamorphic origin from its
textural relation with its matrix. Its formation is associated with fibrolitic
sillimanite (A 02-02). In sample A 14-02 a discontinuous corona of
feldspar aggregates and muscovite around garnet suggests its metamorphic disequilibrium (Fig. 3c).
Nogol Metamorphic Complex The migmatite sample A 39-01 was taken
from the northern part of the Nogol Metamorphic Complex, which is
dominated by metapsammopelites that show a variable degree of
migmatization. The prominent planar fabrics of these rocks are tightly to
isoclinally folded around steep axes. These folds were previously
described as the result of a pre-Famatinian D2 deformation (von Gosen &
Prozzi 1998).
The planar metamorphic fabric is defined by the alignment of finegrained muscovite and biotite crystals. The peak metamorphic conditions
are indicated by the appearance of fibrolitic sillimanite at the expense of
muscovite and biotite. The high-grade metamorphism is accompanied by
the segregation of granitic melts that led to the migmatitic aspect of the
rock. Fabrics are overgrown by arbitrarily oriented undeformed muscovite
plates at up to 1 cm in diameter (Fig. 3d). The formation of a second biotite
on the retrograde path of metamorphism cannot be excluded, as a few of the
biotite crystals are oblique to the foliation and cut the fibrolite needles.
Pringles Metamorphic Complex Two samples of the metasediments have
been subjected to geochronological investigations. Sample A 41/4203 is a
two-mica, garnet- and staurolite-bearing schist from the western transition
zone between the granulite- and amphibolite-facies rocks. Planar fabrics
within this mica schist are defined by muscovite and biotite, although
domains with a less penetrative foliation are present. Garnet grains
locally exhibit a snowball texture pointing to their synkinematic growth.

Sillimanite appears in prismatic crystals of up to 3 cm length. Macroscopically these crystals are parallel the stretching lineation, whereas at
the microscopic scale some of the crystals overprint the foliation,
pointing to a post-deformational growth (Fig. 3e). Sericite rims that
border some of the staurolite crystals indicate retrogression.
The second sample (A 56-01) belongs to a NNE-trending hightemperature mylonite from the western realm of the complex that is
composed of a stable paragenesis of quartz + plagioclase + garnet + Kfeldspar + prismatic sillimanite  cordierite (Fig. 3f). Sillimanite prisms
parallel the stretching lineation. At the microscopic scale two textures are
observed: coarse-grained relicts of the non-mylonitic granulite-facies rock
and the fine-grained mylonitic bands. Within the latter the disequilibrium
of sillimanite is indicated by its replacement by muscovite. Subgrain
boundaries and recrystallization of feldspars indicate mylonitization at
amphibolite-facies conditions.
Conlara Metamorphic Complex Two dominant rock types constitute this
basement domain. The first is a fine-grained banded gneiss, whose
banding is the result of isoclinal folding and metamorphism. Macroscopically, those rocks are composed of quartz, plagioclase, biotite,
muscovite and scarce garnet. Muscovite plates frequently overgrow the
planar fabrics. The other gneissic rock type is generally homogeneous but
locally migmatitic. Transitions between the two rock types seem to be
gradational. The latter type (A 25-01) was subjected to provenance
analyses.
This medium-grained gneiss is composed of quartz, plagioclase
(oligoclase), biotite, muscovite and accessory minerals (tourmaline,
apatite, zircon, opaque minerals). The pervasive banding, defined by finegrained biotite and coarse-grained quartz + plagioclase  biotite, is the
result of the transposition of a previous S1 planar fabric (Lopez de Luchi
& Cerrado 2001). Quartz experienced a static recrystallization that led to
a fine-grained mosaic with triple junctions, although in close proximity to
feldspar crystals parallel subgrains are present. Brown to yellowbrown
biotite appears in two generations. The smaller crystals (up to 0.5 mm in
diameter) help define the S2 foliation, whereas crystals of up to 0.75 mm
are frequently oblique to this foliation. Both generations contain zircon
inclusions. A late-stage alteration of the rock is indicated by the
formation of scarce chlorite and muscovite in the quartz-rich domains.

Geochronology
SHRIMP U/Pb data
To contribute to the continuing discussion on the geodynamic
history of the Sierra de San Luis, six samples from Famatinian(?)
granitoids (three zircon samples) and the metasedimentary
successions (one monazite and two zircon samples) were dated
using the U/Pb SHRIMP method. Reported uncertainties on
SHRIMP analyses are at the 1 level, while uncertainties in any
calculated weighted mean age or concordia age (Ludwig 1998)
are reported at 95% confidence limits if not stated otherwise (see
appendix for details). Ages determined for the intrusions point to
their crystallization during the Famatinian cycle. Inherited
zircons indicate variable sources for the Pringles and Conlara
Metamorphic Complexes, implying for the latter a probable link
and contemporaneity with the Puncoviscana Formation. Constraints on the metamorphic history suggest a common metamorphism of all basement domains during the Famatinian.
Amphibolite-facies conditions within the Pringles basement seem
to continue until the Early Silurian.
Las Verbenas pluton (A 40-01). The tonalite sample yielded
high-quality euhedral, clear zircons with well-developed prismatic habitus (Fig. 4a). Most grains are composed of a weakly
zoned central area overgrown by oscillatory-zoned zircon. The
limited presence of rounded cores suggests some inherited
material. One core analysis provided a 206 U/238 Pb apparent age
of 520  7 Ma (1) (Table 1, analysis 15.1). Nineteen addi-

FA M AT I N I A N G E O DY NA M I C E VO L U T I O N O F G O N DWA NA

969

Fig. 3. (a) High-temperature solid-state fabrics within the Paso del Rey granite (A 02-02). Twinning lamellae of plagioclase are slightly bent. Quartz
aggregates are characterized by grain boundary migration recrystallization fabrics. (Note the appearance of garnet (arrow).) Width of view is c. 15 mm.
(b) Chessboard pattern within quartz from the El Penon pluton (A 59-01) indicates its high-temperature solid-state deformation. Width of view is c.
5.5 mm. (c) Garnet of the leucomonzogranite from the Nogol Metamorphic Complex (A 14-02). The corona of Ms and Pl denotes its disequilibrium.
Width of view is c. 5.5 mm. (d) Recrystallized plagioclasebiotite fabric within the migmatic sample A 39-01 from the Nogol Metamorphic Complex.
The triple junctions between the plagioclase crystals should be noted. The fabrics are overgrown by large discordant muscovite plates (light crystal, upper
half of the image). Width of view is c. 5.5 mm. (e) Staurolite-bearing mica schist (A 41/42-03) from the western realm of the Pringles Metamorphic
Complex. The large staurolite crystals locally overgrow the mica fabric. Width of view is c. 9 mm. (f) Preserved granulite-facies domain showing the
paragenesis of garnet + biotite + plagioclase + sillimanite + quartz within a mylonite from the Pringles Metamorphic Complex (A 56-01). Width of view
is c. 16 mm.

tional analyses were performed on 17 grains (Table 1), and


apart from one analysis that plots above the concordia all data
combine to give a concordia age of 478  4 Ma (95%
confidence, MSWD 1:8), which is considered to reflect the
crystallization age of these zircons (Fig. 5). This Early

Ordovician
nations on
al. 1998)
units; that
cycle (Fig.

crystallization age agree with previous age determiother tonalitic to granodioritic intrusions (Sims et
that were considered as pre-Famatinian intrusive
is, belonging to the early stage of the Famatinian
6).

970

A. STEENKEN ET AL.

Fig. 4. Photomicrographs of the analysed


zircons (ac, e and f; SEM CL images) and
monazites (d; BSE image). (a)
Magmatically zoned zircons from the Las
Verbenas tonalite. (b) Zircons from the
Paso del Rey granite, showing polyphase
growth. Envelopes are generally dark with
minor zoning (see arrow). (c) Rounded
cores (dark zones within the zircons on the
left) within the simply structured zircons
from the El Penon pluton. (d) BSE image
of the monazites, showing monophase
growth (A 39-01; Nogol Metamorphic
Complex). (e) Complex zoned zircons from
sample A 56-01 (Pringles Metamorphic
Complex). (f) View of the variable-shaped
inherited zircons from the Conlara
Metamorphic Complex (A 25-01).

Table 1. Summary of SHRIMP U/Pb zircon data


Grain/spot

206

Pbc

U (ppm)

Th (ppm)

232

Th/238 U

206

Pb*
(ppm)

Pb*/
Pb*

 (%)

207

Pb*/235 U*  (%)

206

Pb*/238 U*  (%)

r (6/8
7/5)

206
206

SHRIMP II, Las Verbenas pluton (A 40-01)


1.1
0.38
228
83
0.38
2.1
0.89
247
108
0.45
2.2
1.54
119
101
0.88
3.1
0.62
254
109
0.44
4.1
0.56
262
93
0.36
6.1
0.9
253
104
0.43
7.1
1.02
232
89
0.39
10.1
0.89
350
154
0.45
11.1
1.33
274
108
0.41
13.1
0.69
359
237
0.68
14.1
0.21
352
122
0.36
16.1
0.83
132
40
0.31
5.1
0.88
237
125
0.54
8.1
0.29
375
124
0.34
12.1
0.61
206
66
0.33
18.1
1.34
194
76
0.41
8.2
1.16
288
92
0.33
9.1
0.93
324
131
0.42
15.1
0.43
341
126
0.38
17.1
0.99
302
135
0.46
SHRIMP II, Paso del Rey, northern stock (A 02-02)
7.1
0.3
564
171
0.31
8.1
0.56
327
230
0.73
12.1
0.48
245
122
0.52
1.2
0.56
534
235
0.45
24.1
0.11
570
319
0.58
5.1
0.71
409
154
0.39
2.1
4.14
1672
426
0.26
4.1
5.82
1612
1355
0.87
6.1
5
1429
201
0.15
10.1
27.67
3264
789
0.25
11.1
7.97
4060
1575
0.4

207

Apparent ages (Ma)


Pb/238 U*

207

Pb/206 Pb*

Disc. (%)

14.8
16.6
8.24
17.3
17.3
16.9
15.7
23.5
18.5
24.5
23.5
9.33
15.6
24.2
14.2
13
18.6
20.7
24.7
18.3

468.3
483.2
492.8
489.1
474.4
477.2
485
481.7
480.8
489.5
481
505
471
466.7
494.5
477.4
461
458.1
519.5
435

 6.1
 6.5
 7.6
 6.4
 6.2
 6.4
 6.9
6
 6.3
 6.4
 6.1
 10
 6.3
 5.7
 6.6
 7.1
 5.9
 5.7
 6.7
 5.9

450
472
504
489
444
479
492
515
518
472
491
497
415
413
438
380
397
396
540
484

 56
 120
 160
 93
 53
 120
 130
 73
 96
 71
 46
 130
 100
 46
 89
 180
 97
 82
 46
 110

4
2
2
0
7
0
1
6
7
4
2
2
13
13
13
26
16
16
4
10

0.0559
0.0565
0.0573
0.0569
0.0558
0.0567
0.057
0.0576
0.0577
0.0565
0.057
0.0571
0.0551
0.055
0.0556
0.0542
0.0546
0.0546
0.0583
0.0568

2.5
5.4
7.5
4.2
2.4
5.3
6
3.3
4.4
3.2
2.1
6
4.5
2
4
8
4.3
3.6
2.1
4.8

0.581
0.606
0.628
0.619
0.588
0.6
0.614
0.616
0.616
0.615
0.609
0.642
0.575
0.57
0.611
0.575
0.558
0.555
0.674
0.547

2.9
5.6
7.7
4.4
2.7
5.5
6.2
3.6
4.6
3.5
2.5
6.3
4.7
2.4
4.2
8.2
4.5
3.9
2.5
5

0.0753
0.0778
0.0794
0.0788
0.0764
0.0768
0.0781
0.0776
0.0774
0.0789
0.0775
0.0816
0.0758
0.07509
0.0797
0.0769
0.07413
0.07364
0.0839
0.06981

1.4
1.4
1.6
1.4
1.4
1.4
1.5
1.3
1.4
1.4
1.3
2.1
1.4
1.3
1.4
1.5
1.3
1.3
1.3
1.4

0.476
0.249
0.209
0.309
0.494
0.255
0.238
0.362
0.297
0.39
0.532
0.328
0.297
0.525
0.329
0.189
0.291
0.335
0.536
0.281

44.2
25.1
18.6
37.5
38.6
26.4
43.3
33.4
32.8
41.4
47.6

562
547.7
542.4
504.2
488.9
464.7
183.8
145
161.8
68.4
80.4

 14
 7.5
 7.3
 6.2
 2.8
 6.2
 2.4
2
 2.3
2
 1.5

507
564
577
508
553
495
476
422
505
930
562

 48
 53
 61
 46
 22
 66
 230
 340
 250
 1300
 380

11
3
6
1
12
6
61
66
68
93
86

0.0574
0.0589
0.0593
0.0574
0.05862
0.0571
0.0566
0.0552
0.0574
0.07
0.059

2.2
2.4
2.8
2.1
1
3
11
15
12
63
18

0.72
0.72
0.717
0.644
0.6368
0.588
0.226
0.173
0.201
0.103
0.102

3.4
2.8
3.1
2.5
1.2
3.3
11
15
12
63
18

0.091
0.0887
0.0878
0.0814
0.07879
0.0747
0.02892
0.02275
0.02541
0.01067
0.01255

2.7
1.4
1.4
1.3
0.59
1.4
1.3
1.4
1.4
3
1.9

0.773
0.503
0.451
0.518
0.499
0.421
0.124
0.09
0.122
0.047
0.106

(continued )

FA M AT I N I A N G E O DY NA M I C E VO L U T I O N O F G O N DWA NA

971

Table 1. (continued )
Grain/spot

206

Pbc

U (ppm)

Th (ppm)

232

Th/238 U

206

Pb*
(ppm)

Pb/238 U*

207

Pb/206 Pb*

74.4
50.42
257.7
341.9
247
88.3
329.4
415.1
426
459.5
251.6
234.6
155.6
388.2
386.6
268.1
412.1
486.8
409.7
452.6
683.5
1006

 1.2
 0.9
 1.7
 1.9
3
 1.9
 2.6
 2.4
 5.8
 5.8
 2.1
 1.6
6
 2.4
5
 1.6
 2.9
 3.9
 5.8
 2.6
 8.6
 22

713
733
405
512
429
528
470
380
403
450
541
465
566
428
449
444
668
413
548
466
667
1093

 440
 590
 180
 110
 560
 990
 130
 110
 150
 25
 390
 170
 140
 79
 100
 78
 150
 76
 63
 31
 43
 75

496.7
492.1
500.8
918.6
517
562
1022.2
1089.7
89.6
192.7
166.7
467.1

 0.7
 1.3
 1.1
 4.6
 32
 2.9
 2.7
 3.4
 3.0
 3.2
 2.0
 2.5

486
494
495
1039

557
1011
1039
1290
310
681
544

8
 13
 10
 17

 26
 14
 10
 1300
 1400
 660
 400

23.5
8.55
21.7
10.6
9.78
11.8
11.1
18.4
24.1
10.1
22.4
14.7
60.4
16.2
9.93
13.3
7.84
14.1
10
10.5
70.6
17.2
10.3
22.2

552.2
530.3
545
534.9
550.1
557.3
562.1
454.2
493.1
318.6
466.8
378.1
510.8
481
401.4
348.2
515
494.2
526.4
572
604
999
2664
417.9

 7.2
 9.2
 7.4
 7.6
 8.8
9
 8.7
 6.1
 6.3
 4.4
 5.9
5
 6.8
 7.1
 6.8
 5.2
 8.9
7
 7.7
 14
 10
 14
 58
 5.8

588
519
506
482
541
492
586
501
488
482
519
486
510
405
386
494
427
425
619
588
653
1115
2699
1249

25.1
44.6
14.7
69.4
106
13.7
75.2
24.6
9.64
7.23
10.5
16.9
29.2
17.8
38.9
17.1
31.2
14.5
19.5
45.8
78.9

590.9
577.6
585.4
591.8
641.5
612
634.1
754.5
753
889
829
1046
1098
1008
949
951
963
973
954
2021
519.9

 6.7
 6.4
 7.7
 6.5
 7.2
 8.3
 7.3
 9.2
 11
 20
 12
 13
 13
 12
 11
 11
 11
 12
 11
 23
 5.9

595
629
557
518
699
757
634
808
1212
968
836
1022
1100
1075
1015
973
963
963
1065
1987
535

26.2
33.4
50
71
48.7
53.6
80.6
50.9
42.7
73.8
35
18.4
32.2
22.7
33.6
22.9
24.2
9.49
25.8
36.1
43.3
14.1
500
39.9
67.2
50.5
81
10.9
61.1
43.8
99.5
76.6
176
744

Pb*/
Pb*

 (%)

207

Pb*/235 U*  (%)

206

Pb*/238 U*  (%)

r (6/8
7/5)

206
206

13.1
9.51
2372
3032
1.32
14.1
12.57
4335
1674
0.4
15.2
3.69
1374
331
0.25
18.1
2.35
1482
336
0.23
20.1
6.95
1350
580
0.44
21.2
14.61
3866
1174
0.31
26.1
2.38
1747
243
0.14
17.2
1.63
876
143
0.17
3.1
2.82
707
140
0.2
5.2
0.22
1161
29
0.03
17.1
7.99
942
427
0.47
22.1
2.41
564
362
0.66
25.1
2.93
1490
1386
0.96
27.1
1.09
421
55
0.14
1.1
1.54
622
242
0.4
21.1
1.13
620
40
0.07
23.1
1.69
420
98
0.24
16.1
0.57
140
37
0.27
9.1
0.7
454
62
0.14
19.1
0.32
576
56
0.1
15.1
0.54
448
144
0.33
14.2
1.1
96
41
0.44
SHRIMP RG, El Penon granite (A 59-01)
1.1
0.08
7255
19
0
3.1
0
585
40
0.07
5.1
0
968
36
0.04
1.2
0.08
383
14
0.04
2.1
51.78
545
16
0.03
4.1
0.15
139
65
0.48
7.1
0.1
413
241
0.6
9.1
0
277
224
0.84
6.1
21.49
6496
642
0.1
8.1
22.2
2285
798
0.36
10.1
14.39
6706
3963
0.61
11.1
8.81
10509
84
0.01
SHRIMP II, Pringles Metamorphic Complex (A 56-01)
5.2

306
29
0.1
9.2
0.18
116
40
0.36
14.1
0.26
286
65
0.24
15.1
0.86
142
27
0.2
17.1
1.19
126
69
0.56
18.1
1.53
150
85
0.59
19.1
1.19
140
27
0.2
1.1
1.08
290
8
0.03
3.1
0.23
352
14
0.04
8.1
0.51
230
8
0.04
9.1
0.23
347
15
0.04
10.1
2.54
275
11
0.04
18.2
2.36
833
23
0.03
2.1
0.9
241
76
0.32
4.1
1.07
178
35
0.2
6.1
1.55
276
95
0.36
13.1
1.71
108
78
0.74
16.1
0.99
204
57
0.29
12.1
0.69
137
37
0.28
19.3
0.74
131
52
0.41
19.2
1.39
825
218
0.27
11.1
0.21
119
50
0.43
7.1
1.09
23
26
1.17
7.2
3.68
371
18
0.05
SHRIMP II, Conlara Metamorphic Complex (A 25-01)
2.1
0.27
304
63
0.22
5.2
0.56
551
33
0.06
6.1
0.85
179
69
0.4
7.1
1.22
828
34
0.04
1.1
2.37
1158
293
0.26
10.1
2.9
157
106
0.7
11.1
0.68
841
112
0.14
1.2
0.54
230
66
0.3
3.1
4.01
89
33
0.39
12.1
5.93
54
13
0.26
17.1
2.38
87
43
0.51
4.1
0.18
111
42
0.39
14.1
1.82
180
92
0.53
8.1
0.71
122
64
0.54
5.1
0.52
285
49
0.18
9.1
0.62
124
98
0.82
13.1
1.26
223
71
0.33
13.2
0.57
103
24
0.24
15.1
0.62
142
62
0.45
16.1

144
144
1.03
8.2
3.63
1055
282
0.28

207

Apparent ages (Ma)


Disc. (%)
90
93
36
33
42
83
30
9
6
2
53
50
73
9
14
40
38
18
25
3
2
8

0.063
0.064
0.0548
0.0575
0.055
0.058
0.0564
0.0542
0.0548
0.05593
0.058
0.0563
0.059
0.0554
0.0559
0.0558
0.0618
0.055
0.0585
0.05634
0.0618
0.0759

21
28
8.1
5.2
25
45
5.9
4.7
6.9
1.1
18
7.6
6.6
3.5
4.5
3.5
6.9
3.4
2.9
1.4
2
3.7

0.101
0.069
0.308
0.432
0.298
0.11
0.408
0.497
0.516
0.5698
0.32
0.288
0.199
0.474
0.476
0.327
0.563
0.595
0.529
0.5651
0.953
1.769

21
28
8.2
5.2
25
45
5.9
4.7
7
1.7
18
7.6
7.7
3.6
4.7
3.6
7
3.5
3.2
1.5
2.4
4.4

0.01161
0.00785
0.04079
0.05448
0.03906
0.01379
0.05243
0.06651
0.06831
0.07388
0.0398
0.03706
0.02443
0.06207
0.06181
0.04247
0.06601
0.07843
0.06561
0.07274
0.1119
0.1689

1.6
1.8
0.67
0.58
1.3
2.2
0.81
0.6
1.4
1.3
0.83
0.7
3.9
0.64
1.3
0.63
0.72
0.84
1.4
0.58
1.3
2.4

0.076
0.066
0.082
0.112
0.05
0.049
0.137
0.127
0.2
0.764
0.047
0.092
0.506
0.179
0.286
0.176
0.103
0.239
0.448
0.382
0.548
0.536

2
0
1
12

1
1
5
93
38
76
14

0.05684
0.05706
0.05708
0.07392

0.05873
0.07288
0.07391
0.084
0.053
0.062
0.058

0.34
0.61
0.47
0.85

1.2
0.67
0.51
66
61
31
18

0.6277
0.624
0.6357
1.561

0.7376
1.727
1.877
0.16
0.22
0.225
0.61

0.37
0.67
0.52
1

1.3
0.73
0.61
66
61
31
18

0.08009
0.07932
0.08078
0.15316
0.0834
0.09109
0.17182
0.18418
0.014
0.03035
0.0262
0.07515

0.15
0.28
0.22
0.54
6.5
0.54
0.29
0.34
3.4
1.7
1.2
0.56

0.394
0.421
0.425
0.534

0.411
0.397
0.556
0.052
0.027
0.039
0.031

 32
 70
 43
 56
 65
 73
 72
 57
 29
 51
 31
 110
 97
 60
 84
 77
 95
 55
 65
 53
 52
 25
 24
 96

6
2
8
11
2
13
4
9
1
34
10
22
0
19
4
29
20
16
15
3
8
10
1
67

0.05957
0.0577
0.0574
0.0568
0.0583
0.057
0.0595
0.0572
0.0569
0.0567
0.05771
0.0569
0.0575
0.0548
0.0544
0.0571
0.0554
0.0553
0.0604
0.0596
0.0614
0.07675
0.1851
0.0821

1.5
3.2
1.9
2.5
3
3.3
3.3
2.6
1.3
2.3
1.4
5
4.4
2.7
3.7
3.5
4.2
2.5
3
2.4
2.4
1.2
1.5
4.9

2
3.6
2.4
2.9
3.3
3.6
3.7
2.9
1.9
2.7
1.9
5.2
4.6
3.1
4.1
3.8
4.5
2.8
3.3
3.4
3
1.9
2.7
5.1

0.0894
0.0857
0.0882
0.0865
0.0891
0.0903
0.0911
0.07299
0.0795
0.05067
0.07509
0.06041
0.0825
0.0775
0.0643
0.0555
0.0832
0.0797
0.0851
0.0928
0.0982
0.1676
0.512
0.06698

1.3
1.7
1.4
1.4
1.5
1.5
1.5
1.4
1.3
1.4
1.3
1.3
1.3
1.4
1.6
1.4
1.5
1.4
1.4
2.4
1.7
1.4
2.3
1.3

0.675
0.467
0.573
0.492
0.448
0.412
0.419
0.466
0.705
0.519
0.679
0.252
0.287
0.47
0.405
0.37
0.337
0.486
0.433
0.71
0.564
0.753
0.84
0.262

 59
 32
 94
 83
 100
 140
 43
 51
 110
 280
 160
 48
 79
 47
 31
 55
 83
 61
 26
 17
 180

1
8
5
14
8
19
0
7
38
8
1
2
0
6
7
2
0
1
10
2
3

0.0598
2.7
0.06071 1.5
0.0587
4.3
0.0577
3.8
0.0627
4.8
0.0645
6.6
0.0609
2
0.0661
2.4
0.0806
5.7
0.0714 14
0.0669
7.6
0.0733
2.4
0.0762
4
0.0752
2.4
0.073
1.5
0.0715
2.7
0.0712
4.1
0.0712
3
0.07488 1.3
0.1221
0.97
0.0581
8

0.096
0.0937
0.0951
0.0962
0.1046
0.0996
0.1034
0.1242
0.1239
0.1478
0.1373
0.1762
0.1856
0.1693
0.1586
0.1589
0.1611
0.1629
0.1595
0.3683
0.084

1.2
1.2
1.4
1.2
1.2
1.4
1.2
1.3
1.6
2.4
1.5
1.3
1.3
1.3
1.2
1.3
1.2
1.4
1.3
1.3
1.2

0.397
0.61
0.303
0.292
0.241
0.21
0.518
0.466
0.264
0.178
0.2
0.489
0.315
0.485
0.63
0.435
0.291
0.412
0.708
0.802
0.145

0.735
0.682
0.698
0.677
0.716
0.71
0.748
0.576
0.624
0.396
0.598
0.474
0.653
0.586
0.482
0.437
0.635
0.608
0.709
0.762
0.832
1.774
13.06
0.759

0.791 3
0.785 1.9
0.77
4.5
0.765 3.9
0.905 4.9
0.885 6.7
0.867 2.3
1.131 2.8
1.377 5.9
1.45 14
1.267 7.7
1.78
2.7
1.95
4.2
1.756 2.7
1.597 2
1.567 3
1.581 4.3
1.599 3.3
1.646 1.8
6.2
1.6
0.673 8.1

Errors are 1; Pbc and Pb* indicate the common and radiogenic portions, respectively. Error in standard calibration was 0.25% (SHRIMP II) and 0.12% (SHRIMP RG). ,
common Pb too high to calculate any meaningful 207 Pb/206 Pb data or age.
*Common Pb corrected using measured 204 Pb.

972

A. STEENKEN ET AL.

Fig. 5. Wetherill concordia diagrams and cumulative probability plots of SHRIMP U/Pb zircon data for the analysed igneous samples from the Sierra de
San Luis. Las Verbenas Tonalite: grey error ellipses are used to calculate the concordia age. Paso del Rey, north: white error ellipses (inner overgrowth
zones) and ruled ellipses (tips with high U and common Pb) are used to calculate the discordia age. Light and dark grey refer to core data. El Penon
Granite: white error ellipses refer to high U (low Th/U) and partly unacceptable high common Pb, whereas grey ellipses indicate near-concordant rim and
core analyses. The cumulative probability plots show the Proterozoic inheritance of the zircon cores.

Paso del Rey granite, northern stock. The zircons from the Paso
del Rey granite (A 02-02) are generally subhedral and heterogeneous, with many grains showing overgrowths of euhedral
zircon (Fig. 4b). In SEM cathodoluminescence (CL) images
these overgrowths are observed on host crystals of various forms
that are presumably inherited grains. The overgrowths have a low
CL response but do have an inner zone of higher response (lower
U) that probably reflects continuous crystallization during changing conditions and magmatic compositions rather than two
different overgrowths (see arrow in Fig. 4b). The rims are
sometimes large enough to be accessed by the SHRIMP but in
almost all cases they have very high U contents, high common
Pb and are generally highly discordant.
A total of 34 analyses were carried out on 27 grains (Table 1).
Most analyses were sited in what is presumed to be the magmatic
overgrowths, to establish an age of crystallization or emplacement
for the granite. Other analyses were sited on cores or discrete,
zoned grains that appeared to be magmatic. The 207 Pb/206 Pb ages
determined for cores range from about 1.09 Ga to about 495 Ma.
The two analysed near-concordant older cores appear in the
cumulative probability plot only (Fig. 5). A group of three core
analyses gives a concordia age of 547  10 Ma (2). This is the

most reliable estimate of the maximum age of the granite,


assuming these cores are inherited. In terms of calculating a
magmatic age for this rock, the problems of discordance only
allow a date to be calculated through regression of all (n 26) the
variably discordant analyses of the overgrowths (interpreted to be
magmatic). Although this approach is limited by the imprecision
of the analyses (especially for the 207 Pb/235 U data) it does yield an
upper intercept with the concordia at 456  30 Ma (MSWD
0:26, probability of fit (P) 1:00; Fig. 5). This regression
includes an analysis of a dark CL embayment (Table 1, analysis
5.2) with very low Th/U of 0.03 that yielded a 206 Pb/238 U age of
460  6 Ma, which is interpreted to represent the best estimate of
the age of this granite. It is apparent that the high-U rims (and
even some of the lower-U inner zones) are severely altered and
produce discordant ages. Therefore, the establishment of a sound
crystallization age for the Paso del Rey granite turned out to be
very difficult and has to involve regional geological constraints.
Previous attempts to elucidate the crystallization age of the granite
are within the error range of the new result. Llambas et al. (1991)
determined a debatable Rb/Sr whole-rock errorchron age of
454  21 Ma, whereas recently a 207 Pb/206 Pb zircon evaporation
age of 491  19 Ma was presented (Steenken et al. 2006).

FA M AT I N I A N G E O DY NA M I C E VO L U T I O N O F G O N DWA NA

973

Fig. 6. Compilation of the available U/Pb


zircon and monazite data and new
geochronological results on the igneous and
metamorphic history of the Sierra de San
Luis. (For references see insert.)
Provenance data are discussed in the text.

El Penon pluton. The zircons from the El Penon sample A 59-01


are almost all simply structured into a core and magmatic
overgrowth (Fig. 4c). The cores are generally rounded and
suggest inheritance from a sedimentary source but could also be
the result of resorption during crystallization. The overgrowths
are usually euhedral, sometimes zoned, and are very dark in CL,
indicating high U contents.
Only 11 analyses were carried out on this sample, and the data
are presented in Table 1 and plotted in Figure 5. The cores show
a range of ages from Mesoproterozoic to Neoproterozoic
Ediacaran. The overgrowths are again highly enriched in U (but
with very low Th/U) and some have unacceptably high common
Pb contents. The common 206 Pb in one spot (analysis 2.1) is over
50% of the total 206 Pb and as a result any calculation of 207 Pb/
206
Pb data or ages is totally compromised. A group of three
analyses provide almost concordant ages (Fig. 5, Table 1;
analyses 1.1, 3.1, 5.1). The weighted mean of the 206 Pb/238 U ages
at 497  8 Ma is indistinguishable from the weighted mean of
the 207 Pb/206 Pb ages at 490  11 (95% confidence, MSWD

0:35, P 0:7). Although high U zircons prone to matrix


problems (Table 1, analysis 1.1) that lead to inaccurate 206 Pb/
238
U ages, the concordance of the two ages and their agreement
with the structural arrangement of the pluton in the host makes a
late Cambrian emplacement reasonable.
Nogol Metamorphic Complex. The monazites from sample A
39-01 are euhedral with a generally prismatic habit, although a
few stubby crystals appear. The back-scattered electron (BSE)
image indicates a week zoning of the crystals. Inclusions are
scarce (Fig. 4d). The SHRIMP analyses produced a homogeneous dataset with respect to U, Pb and Th (Table 2). Common
Pb is generally less than 0.8%. One of 17 analyses has been
rejected, resulting in an average 208 Pb/232 Th age of 478  4 Ma
(95% confidence). This is indistinguishable within error from
two concordant U/Pb ages of 473  4 Ma and 484  4 Ma (95%
confidence) that can be calculated (Fig. 7). These ages are the
result of an Early Ordovician metamorphism. Inheritance of
older ages has not been detected.

A. STEENKEN ET AL.

 25
 22
5
5
5
5
 11
5
5
5
5
7
 15
7
5
 10
6
469
500
467
470
476
477
490
491
476
480
483
483
498
479
497
468
477
0.07352
0.0733
0.07539
0.0753
0.07586
0.07688
0.07577
0.07649
0.07751
0.0793
0.07857
0.07732
0.07865
0.07769
0.07893
0.07757
0.07807

1.0
1.3
0.8
0.83
0.8
0.9
0.88
1.1
0.84
0.88
0.86
0.78
0.81
0.91
0.74
0.73
0.78

Pb*/  (%)
U*
238

1.5
1.4
1.6
2.1
1.4
2.0
2.3
3.0
2.4
3.0
2.4
1.9
1.5
1.9
1.1
1.0
1.0
0.5659
0.5666
0.5889
0.58
0.6081
0.588
0.588
0.579
0.588
0.581
0.587
0.587
0.615
0.617
0.6106
0.6025
0.6143
5.2
4.4
1.0
1.0
1.1
1.0
2.2
1.0
1.0
1.1
1.0
1.4
3.0
1.4
1.0
2.1
1.3
0.0235
0.025
0.0234
0.0235
0.0238
0.0239
0.0245
0.0246
0.0238
0.024
0.0242
0.0242
0.0249
0.024
0.0249
0.0234
0.0239
1.0
1.3
0.8
0.83
0.8
0.9
0.88
1.1
0.84
0.88
0.86
0.78
0.81
0.91
0.74
0.73
0.78
13.6
13.64
13.26
13.28
13.18
13.01
13.2
13.07
12.9
12.61
12.73
12.93
12.71
12.87
12.67
12.89
12.81
0.54
0.39
0.69
0.77
0.73
0.76
0.89
0.74
0.80
0.86
0.82
0.64
0.74
0.95
0.52
0.42
0.54
0.0567
0.05642
0.05785
0.05907
0.05898
0.05829
0.05807
0.06176
0.05884
0.05864
0.05812
0.05757
0.05811
0.05959
0.05665
0.05756
0.05681
1.0
1.3
0.79
0.82
0.8
0.89
0.87
1.1
0.83
0.87
0.85
0.77
0.8
0.91
0.74
0.73
0.78
34306
28824
31968
37949
52685
43904
26828
33120
45784
24866
34369
39158
40802
28846
30579
36700
39426
3923
6261
1882
1800
2353
1922
1414
1832
2031
1311
1395
2552
2214
1179
3385
4778
4357
0.11
0.04
0.15
0.39
0.10
0.35
0.22
0.84
0.48
0.69
0.49
0.31
0.17
0.24
0.07
0.15
0.03
3.1
12.1
8.1
16.1
4.1
13.2
1.1
2.1
6.1
7.1
14.1
9.1
13.1
5.1
11.1
10.1
15.1

Errors are 1; Pbc and Pb* indicate the common and radiogenic portions, respectively
*Common Pb corrected using measured 204 Pb.

13.59
13.64
13.25
13.23
13.17
12.96
13.17
12.96
12.84
12.53
12.67
12.89
12.69
12.84
12.66
12.87
12.81
 24
 11
 31
 43
 25
 39
 46
 60
 50
 63
 49
 38
 28
 37
 17
 16
 13
445
456
478
449
535
432
464
413
413
336
382
416
482
516
455
466
494
 3.7
 3.7
 4.1
 4.2
 4.1
 4.3
 4.6
 4.3
 4.4
 4.5
 4.5
 4.0
 4.4
 4.8
 4.0
 3.7
 3.9

Pb/206 Pb*
207

Pb/238 U*
206

457.3
478.3
468.3
469.3
470.7
478.3
481.7
474.7
482.0
477.3
482.0
473.9
494.7
485.2
492.8
473.1
476.9
248.0
414.4
122.0
117.1
153.3
127.6
94.4
121.2
136.0
87.1
93.4
167.7
151.9
79.4
231.2
313.0
287.4
9.0
4.8
17.6
21.8
23.1
23.6
19.6
18.7
23.3
19.6
25.5
15.9
19.0
25.3
9.3
7.9
9.3

206

Pb*/  (%)
U*
235

207

 (%)
Pb*/
Th*
232

208

U/  (%)
Pb**

238

206

Pb/  (%)
Pb(tot)

207

206

U/  (%)
Pb(tot)

238

206

Apparent ages (Ma)


Pb*
(ppm)

206

Th/238 U
232

Th (ppm)
U (ppm)
Pbc
206

Grain/spot

Table 2. Summary of SHRIMP U/Pb monazite data for sample A 39-01 (Nogol Metamorphic Complex)

Pringles Metamorphic Complex. The zircons from sample A 5601 are generally clear, with some inclusions, but of variable habit
ranging from elongate prismatic to more oblate rounded forms
(Fig. 4e). SEM CL imaging shows well-developed oscillatory
zoning in most grains. Fragmented, inherited cores are observed,
and many zircons have a small rim of unzoned zircon, denoting
a complex geological history.
Twenty-four analyses were carried out on 19 zircons, with the
younger portion of the data plotted on a conventional UPb
Wetherill concordia plot (Fig. 8, Table 1). The results of the U/Pb
analyses show a complex pattern as a result of inherited time
increments and new growth of rims during high-temperature
metamorphism. Although inheritance is common the appropriate
cores are difficult to identify. The oldest xenocrystic core
analysed yields an Archaean age of about 2.7 Ga, whereas
Mesoproterozoic ages were recorded in two cases. The majority
of the analyses were sited in the strongly zoned, magmatic zircons
(Fig. 8, dark grey error ellipses) but despite the moderate U (and
low Th) contents in this zircon type, many of the analyses show
extensive Pb loss. Therefore, calculating an age for these zircons
is difficult, but a number of the least discordant analyses do plot
as a group on the concordia, yielding an age of 545  8 Ma (95%
confidence, Fig. 8). Two analyses (Table 1, 19.2 and 19.3) of what
appears to be a core give slightly older apparent ages.
Attempts to date the small enveloping rims were not satisfactory because of severe Pb loss. Low Th/U ratios in the range
0.030.05 could indicate a metamorphic origin. A second
concordia age based on the oldest and least discordant rim
analyses (n 5) gives an age 498  10 Ma (95% confidence, Fig.
8). In light of the available geochronological and structural data,
no meaning could be attributed to one younger and near perfectly
concordant rim analysis at c. 400 Ma (Table 1, analysis 4.1).

0.688
0.932
0.487
0.389
0.573
0.450
0.385
0.370
0.346
0.293
0.355
0.409
0.526
0.474
0.695
0.698
0.784

Error corr.

208

Age (Ma)

Pb/232 Th*

974

Conlara Metamorphic Complex. The zircons from sample A 2501 are generally anhedral to subhedral, with most grains showing
some degree of rounding that could be of either sedimentary or
metamorphic origin (Fig. 4f). The SEM CL imaging shows the
grains to be composed predominantly of zoned fragments
representing probably inherited cores, surrounded by very thin
overgrowths. In many instances these overgrowths are dark in
CL, with thin outer surfaces of bright CL colours, which could
represent a late-stage reworking or recrystallization of the zircon.
These rims are too small to analyse.
Twenty-one analyses were carried out on 17 grains (Table 1)
to clarify the full history of this rock. The data are presented in
the form of a Wetherill UPb concordia plot (Fig. 8). It is clear
that most of the zircons from this rock are inherited from a range
of sources. Some have elevated common Pb contents of a few
per cent (e.g. analyses 3.1 and 12.1 in Table 1). The oldest grain
measured (analysis 16.1 in Table 1) gives an age close to 2 Ga,
but the majority were derived from Late Mesoproterozoic and
Neoproterozoic sources, as documented by age clusters at
962  17 Ma (n 5, 95% confidence) and 631  15 Ma (n 3,
95% confidence). Establishing a maximum age for the sedimentation of this rock is difficult. The youngest cluster of analyses
(n 4) gives a weighted mean 206 U/238 Pb age of 587  7 Ma
(2). Two of these analyses were sited in zircon rims, and two in
zoned parts of larger crystals. One younger age obtained from a
rim (analysis 8.2 in Table 1) has relatively high U, high common
Pb and most probably suffered Pb loss, making this result highly
suspect. Similar findings have been reported for the adjacent
basement block of the Sierras de Cordoba (Schwartz & Gromet
2004; Escayola & Pimentel 2006).

FA M AT I N I A N G E O DY NA M I C E VO L U T I O N O F G O N DWA NA

975

Fig. 7. Wetherill concordia diagram and weighted mean of 208 Pb/232 Th ages from SHRIMP U/Pb monazite data for sample A 39-01. Light and dark grey
error ellipses are used to calculate the two concordia ages. White error ellipses are excluded from the dataset. U/Pb and Pb/Th ages are indistinguishable
within error.

Fig. 8. Wetherill concordia diagrams and cumulative probability plots of SHRIMP U/Pb zircon data for the analysed metasedimentary samples from the
Pringles and Conlara Metamorphic Complexes. A 56-01: dark grey error ellipses represent core data of magmatically grown zircons that were used to
calculate the concordia age. White error ellipses (metamorphic grown rims) and light grey error ellipses (cores with extensive Pb loss) are used to
calculate the discordia age of the granulite-facies metamorphism. A 25-01: dark error ellipses refer to analysed cores. White error ellipses are excluded
from further inspection because of the scarcity of data and rather high common Pb. The cumulative probability plots show the Proterozoic and Archaean
inheritance within the zircon cores.

976

A. STEENKEN ET AL.

with micro-inclusions rather than to the host-mineral (Siegesmund


et al. 2007). However, based on the linearly correlated uranogenic
Pb release pattern and the establishment of statistically welldefined correlation lines (in fact, isochrons with low MSWD
values), we emphasize that inclusions in the host phases have the
same average age as their hosts and, consequently, the ages are
interpreted as closely post-dating the peak (granulite) metamorphic conditions during which the garnets and the staurolite
were formed. For sample A 56-01 (garnet) from the granulitefacies metasediments of the Pringles Metamorphic Complex a
PbSL age of 452  19 Ma was obtained. This is significantly
younger than the zircon concordia age of 498  10 Ma, calculated
on the zircon rims of the same sample (Fig. 8). The 207 Pb/206 Pb
age of 460  12 Ma from garnet of sample A 14-02 from a
leucocratic monzogranite of the Nogol Metamorphic Complex is
consistent with previous assumptions concerning the peak of
Famatinian metamorphism (Gonzalez et al. 2002a, 2004). It is
within error indistinguishable from the U/Pb and Pb/Th ages at
about 478 Ma of our sample A 39-01.
The staurolite separate of sample A41/42-03 yield a welldefined 206 Pb/207 Pb age of 428  17 Ma. This age appears to be
too young to match the expected time interval of Famatinian peak
metamorphism. Nevertheless, it is consistent with a previously
reported three-point Sm/Nd isochron of 434  12 Ma for a solidstate deformed mafic rock in the vicinity of the sample site
(Steenken et al. 2006) that is interpreted to correspond to the latest
stage of Famatinian amphibolite-facies metamorphism. The relation of this age to the latest stage of Famatinian metamorphism is
possible because of the observed retrogression of the staurolite at
the microscopic scale as well as because formation and diffusion
within staurolite could occur at temperatures as low as c. 520 8C
(Bucher & Frey 1994).

Pb/Pb stepwise leaching (PbSL) data


Pb/Pb stepwise leaching data for the two garnet and one staurolite
separate are summarized in Table 3, and Figure 9 depicts the
Pb-isotopic progressions of successive leach steps (17) in
uranogenic Pb (207 Pb/204 Pb 206 Pb/204 Pb space; Fig. 9ac) and
thorogenic Pb (208 Pb/204 Pb 206 Pb/204 Pb space; Fig. 9d-f ). All
errors are quoted at the 2 level (95% confidence limits). The first
leach steps (a 12:1 mixture of 1.0N HBr and 1.5N HCl) selectively
extracted the most primitive Pb established by the data in all three
samples, whereas steps (3), (4) and (5) (4N HBr, 2 3 8N HBr)
leached relatively radiogenic Pb, as indicated by higher isotopic
ratios (Table 3, Fig. 9). The linearity of garnet and staurolite leach
steps in the uranogenic plots indicates two-component mixing of
common and radiogenic Pb reservoirs from within the host phases
themselves and/or between the host phases and their inclusions.
The slopes defined by leach steps on the uranogenic diagrams
(Fig. 5ac) yield apparent 207 Pb/206 Pb isochron ages of
452  19 Ma (MSWD 0:92) for sample A 56-01, 460  12 Ma
(MSWD 1:7) for sample A 14-02, and 428  17 Ma
(MSWD 1:4) for sample A 41/42-03. The corresponding thorogenic diagrams for these leach steps (Fig. 9d-f) show that a phase
of relatively high 208 Pb/206 Pb (Th/U) was preferentially leached in
steps (3) and (4) in sample A 14-02, whereas in the other two
cases a more or less linear release spectrum of radiogenic
uranogenic and thorogenic Pb is indicated by the correlation lines
in the respective diagrams (Fig. 9e and f). We cannot observe the
tendency for the presence of a highly uranogenic Pb in the final
(step (7)) leaches of all samples, as would be expected in the case
where microscopically small zircon inclusion were present (they
would be strongly attacked by HF applied during the final step).
From experience in previous PbSL experiments applied to garnet
and staurolite (Frei & Kamber 1995; Schaller et al. 1997; Dahl &
Frei 1998), in which strong HBr selectively attacked microscopically small inclusions of phosphates, based on the combined
thorogenic and uranogenic character of the released radiogenic Pb,
we favour a derivation of the radiogenic Pb in the present
experiments from monazite inclusions, and not from loosely
bound radiogenic Pb leached from the garnet lattices. Similar, age
information from PbSL experiments on staurolite is associated

Discussion
Timing of sedimentation and provenance of the basement
units of the Sierra de San Luis
The detrital zircon ages of sample A25-01 from the Conlara
Metamorphic Complex allow a discussion of the provenance of

Table 3. Summary of PbSL experiments for the samples A 56-01 (grt), A 14-02 (grt) and A 41/42-03 (st)
Pb/204 Pb

 2

1
2
3
4
5
6
7

18.921
21.197
56.092
621.928
275.356
24.581
21.947

0.032
0.046
0.268
12.803
10.322
0.091
0.157

15.689
15.832
17.716
49.479
30.292
16.030
15.850

30 min
1h
3h
6h
17 h
17 h
96 h

1
2
3
4
5
6
7

19.566
41.501
646.356
563.421
52.088
34.710
44.660

0.027
0.244
48.101
34.027
0.588
0.126
0.350

20 min
1h
3h
12 h
48 h
12 h
24 h

1
2
3
4
5
6
7

19.389
21.910
234.807
284.994
71.406
47.411
25.879

0.032
0.045
2.300
2.674
0.865
0.453
0.034

Sample

Phase

Acid*

Time

A 56-01
A 56-01
A 56-01
A 56-01
A 56-01
A 56-01
A 56-01

grt
grt
grt
grt
grt
grt
grt

mix
1N HBr
4N HBr
8.8N HBr
8.8N HBr
HF conc.
HF conc.

30 min
1h
3h
6h
17 h
17 h
96 h

A14-02
A14-02
A14-02
A14-02
A14-02
A14-02
A14-02

grt
grt
grt
grt
grt
grt
grt

mix
1N HBr
4N HBr
8.8N HBr
8.8N HBr
HF conc.
HF conc.

A41/42-03
A41/42-03
A41/42-03
A41/42-03
A41/42-03
A41/42-03
A41/42-03

st
st
st
st
st
st
st

mix
1N HBr
4N HBr
8.8N HBr
8.8N HBr
14N HNO3
residue

Leach
step

206

207

Pb/204 Pb

 2

0.028
0.036
0.085
1.019
1.137
0.065
0.114

40.020
46.551
179.027
2279.903
935.189
39.680
39.043

0.093
0.108
0.866
46.959
35.064
0.205
0.281

0.8292
0.7469
0.3158
0.0796
0.1100
0.6521
0.7222

0.0003
0.0003
0.0002
0.0001
0.0002
0.0011
0.0005

2.1151
2.1961
3.1917
3.6659
3.3963
1.6143
1.7789

15.674
16.959
50.818
46.272
17.370
16.565
17.000

0.023
0.100
3.783
2.795
0.197
0.060
0.134

38.540
47.125
707.088
869.085
58.088
39.004
39.133

0.059
0.279
52.624
52.509
0.657
0.143
0.307

0.8011
0.4086
0.0786
0.0821
0.3335
0.4772
0.3807

0.0003
0.0002
0.0001
0.0001
0.0002
0.0002
0.0002

15.691
15.844
27.593
30.439
18.623
17.292
16.089

0.027
0.033
0.276
0.286
0.227
0.166
0.023

38.851
41.626
482.669
610.431
116.121
80.482
38.562

0.068
0.090
4.736
5.740
1.409
0.772
0.066

0.8093
0.7231
0.1175
0.1068
0.2608
0.3647
0.6217

0.0003
0.0003
0.0002
0.0001
0.0004
0.0003
0.0003

Pb/204 Pb

r1 206 Pb=204 Pb v. 207 Pb/204 Pb error correlation (Ludwig 1990); r2 206 Pb=204 Pb v.
*Mix denotes a mixture of 1N HBr and 1.5N HCl in the proportion 12:1.

Errors are two standard deviations absolute (Ludwig 1990).

 2

208

208

207

Pb/206 Pb

 2

Pb/204 Pb error correlation (Ludwig 1990).

208

Pb/206 Pb

 2

r1

r2

0.0032
0.0014
0.0019
0.0018
0.0022
0.0058
0.0012

0.978
0.984
0.994
1.000
0.999
0.911
0.995

0.752
0.962
0.992
1.000
1.000
0.721
0.995

1.9697
1.1355
1.0940
1.5425
1.1152
1.1237
0.8762

0.0009
0.0007
0.0008
0.0026
0.0006
0.0005
0.0004

0.976
0.996
1.000
1.000
0.998
0.996
0.997

0.957
0.995
1.000
1.000
0.999
0.993
0.998

2.0038
1.8999
2.0556
2.1419
1.6262
1.6976
1.4901

0.0008
0.0011
0.0008
0.0010
0.0008
0.0011
0.0015

0.983
0.981
0.981
0.998
0.993
0.996
0.954

0.973
0.965
0.999
0.999
0.999
0.998
0.801

FA M AT I N I A N G E O DY NA M I C E VO L U T I O N O F G O N DWA NA

977

Fig. 9. Isochron plots presenting the results


from PbSL experiments. Left column shows
Pb-isotopic progression of successive leach
steps (17) in uranogenic Pb, and the Pbisotopic progression of thorogenic Pb is
shown in the right column for comparison.
(a,d) Garnet from sample A 56-01, a
granulite-facies rock from the Pringles
Metamorphic Complex. (b,e) Garnet from
sample A 14-02, a pervasively solid-state
deformed leucomonzogranite from the
Nogol Metamorphic Complex. (c,f)
Staurolite from sample A 41/42-03, an
amphibolite-facies mica schist from the
Pringles Metamorphic Complex.

its magmatic precursors. The recorded ages (n 20) define at


least three distinct peaks (Fig. 8). (1) The latest Ediacaran
population can be separated into two subordinate peaks at
587  7 Ma and 631  15 Ma, representing the final amalgamation of southwestern Gondwana during the pan-AfricanBrasiliano Orogeny (Leite et al. 2000; Janasi et al. 2001). (2) A
significant contribution to the inheritance is indicated by an Early
Neoproterozoic cluster at 962  17 Ma, with some older analyses
covering the typical Grenvillian spectra. (3) One analysis yielded
a Palaeoproterozoic age (c. 2.0 Ga) only. Almost identical zircon
inheritance within the metasediments of the northwestern ranges
of Cordoba was demonstrated (Schwartz & Gromet 2004) and
correlated with the Puncoviscana Formation based on additional
Nd and Pb isotopic data (cf. Bock et al. 2000). Geochemical and
isotopic studies of the Conlara basin deposits (Steenken et al.
2004; Zimmermann 2005) confirm the connection with the
Puncoviscana Formation. This suggests that the protolith of the
basement of both sierras might have belonged to the same basin
during the Ediacaran, when the Puncoviscana sediments were
deposited. To explain the ubiquitous Grenvillian inheritance of
the sediments a probable source area within the cratonic South
America was proposed (Schwartz & Gromet 2004). Increasing

evidence on the Grenvillian signature within the basement of


southwestern Gondwana (e.g. Wareham et al. 1998) shades doubt
on the uniqueness of the Laurentia Grenvillian source but rather
suggests an appropriate source component within the Gondwanan
basement. The inherited age spectra can be clearly distinguished
from those of the Cuyania Terrane (Finney et al. 2005), which
show a significant Early Mesoproterozoic inheritance.
Detrital zircons within the Pringles Metamorphic Complex
(sample A56-01, Fig. 6) provide less detailed information
concerning their provenance. A maximum U/Pb age for the
sedimentation at 545  8 Ma (n 9) is regarded as the maximum deposition age. Older age contributions to the entire
inheritance are subordinate. Two analysed grains contributed
either a Mesoproterozoic or a Late Archaean increment to the
dataset of this sample (Fig. 8). A population of inherited zircons
in the northern stock of the Paso del Rey granite shows a
comparable peak of concordant U/Pb ages at 547  10 Ma.
The Pampean orogen is bracketed between 530 and 515 Ma
(Rapela et al. 1998a,b; Stuart-Smith et al. 1999), although
Ediacaran crystallization ages have been reported for some calcalkaline igneous rocks (Llambas et al. 2003). A Pampean
structural evolution and metamorphism of the Pringles basement

978

A. STEENKEN ET AL.

was proposed by von Gosen & Prozzi (1998), von Gosen


(1998a,b) and von Gosen et al. (2002). Gonzalez et al. (2002a)
reinterpreted the detrital zircon ages at c. 530 Ma (Sims et al.
1998) as a metamorphic age. In conrast, our results favour an
exclusively Pampean source, which argues against a Pampean
structural evolution and metamorphism for the Pringles basement.
Sims et al. (1998) had already suggested a restricted Pampean
source area. New structural data for a previous Pampean
deformation have not been reported (Whitmeyer & Simpson
2004; Steenken et al. 2006). The geochemical characterization of
andesitic and acidic contributions in the sediments is suggestive
of a continental island-arc or an active margin setting (Lopez de
Luchi et al. 2003) that developed to the west of the Pampean
orogen (Hauzenberger et al. 2001; Steenken et al. 2006). In
contrast, Chernicoff & Ramos (2003) seized the possibility of the
development of a foreland basin. Predominantly Pampean detrital
zircon ages were reported for other localities in the Sierras
Pampeanas, such as, for instance, in the basement of the Sierra
Famatina (Fig. 1; Collo et al. 2005), characterizing the Pampean
orogen as an important barrier during the early to mid-Cambrian.

Igneous and metamorphic history of the basement units of


the Sierra de San Luis
Pringles Metamorphic Complex. It is generally accepted that the
magmato-metamorphic evolution of the Pringles Metamorphic
Complex essentially took place during the Famatinian cycle
throughout the Ordovician. The implementation of SHRIMP and
PbSL analysis will help to define distinct episodes of this history.
Five zircon rim analyses including two low Th/U spots define
a concordant U/Pb age at 498  10 Ma (sample A 56-01) that
provides a good constraint on the timing of the granulite-facies
metamorphism. This age agrees with the earliest evidence of
magmatism in the Sierra de San Luis (Sims et al. 1998; von
Gosen et al. 2002; Steenken et al. 2006) but differs significantly
from the Mid-Ordovician time constraint on the end of granulitefacies metamorphism based on SHRIMP U/Pb monazite (and
few zircon) ages (Sims et al. 1998). Only two new zircon
analyses yield ages at c. 460 Ma (Fig. 8). Complementary 207 Pb/

206
Pb garnet ages within the same sample indicate that the Pb
isotope equilibration took place during the Mid- to Late Ordovician (452  19 Ma), defining the lower limit of the metamorphism (Fig. 6). Cooling below c. 580 8C at c. 454  18 Ma was
demonstrated by K/Ar hornblende dating (Steenken et al. 2006).
As there are no petrological data indicating a two-phase metamorphic history (Hauzenberger et al. 2001), the new data suggest
that the high-grade metamorphism lasted for c. 46 Ma.
Little is understood about the meaning of the young 207 Pb/
206
Pb age of 428  17 Ma. A possible resetting of the Pb isotope
system as a result of the Devonian compression as a result of the
approaching Chilenia Terrane seems unlikely. Hitherto, only
resetting of the Ar system in micas has been reported (Sims et
al. 1998; Steenken et al. 2004). The connection of the age with a
SilurianDevonian accretion of the Cuyania Terrane against the
Pampean Terrane as proposed by Keller et al. (1998) and Keller
(1999) appears to be unlikely in view of the increasing evidence
for a Mid-Ordovician collision (Casquet et al. 2001; Astini &
Davila 2004). The most likely explanation for the age of 428 Ma
is probably related to the re-equilibration of the staurolite during
slow cooling after the peak metamorphism. This protracted
cooling can find a probable explanation in the differential
exhumation of the basement as proposed by Steenken et al.
(2004, 2006).
Post-dating the onset of high-grade metamorphism the Pringles
Metamorphic Complex was intruded by tonalitic to granitic melts
that exhibit distinct structural relations with their host. A group
of granodioritic to tonalitic intrusions is restricted to the lowgrade metamorphic successions of the San Luis Formation.
These were previously considered to predate the Famatinian
orogeny (Llambas et al. 1998). Available SHRIMP U/Pb ages
for two of these intrusions, i.e. the Bemberg and Tamboreo
plutons at 468  6 and 470  5, respectively (Stuart-Smith et al.
1999), are insignificantly younger than the new age of the Las
Verbenas tonalite at 478  4 (Fig. 6), increasing the evidence
that all of these tonalitic to granodioritic stocks were emplaced
during a narrow period during the Mid-Ordovician (Fig. 6).
Structural investigations (Steenken et al. 2006) and microstructures denote an emplacement synchronous with the development

Fig. 10. Chart illustrating the


tectonometamorphic and igneous history of
the Sierra de San Luis within the frame of
the Pampean and Famatinian orogenic
cycles.

FA M AT I N I A N G E O DY NA M I C E VO L U T I O N O F G O N DWA NA

of the first metamorphic fabric in the hosting San Luis Formation


(Fig. 10). This foliation is considered to be contemporary with
the S2 foliation in the higher-grade metamorphic rocks of the
Pringles Metamorphic Complex (see von Gosen & Prozzi 1998;
Steenken et al. 2006). Therefore, SHRIMP U/Pb zircon ages
provide at least the upper time constraint for the ubiquitous D2
event in the central basement unit of the Sierra de San Luis.
The new results for the northern stock of the Paso del Rey
granite (Fig. 6) also point to an Early to Mid-Ordovician
emplacement (456  30 Ma) of this sheet-like intrusion.
Although the calculated error of the isochron on the magmatic
rims is high, it is in accordance with a previously presented Rb/
Sr errorchron and 207 Pb/206 Pb evaporation experiments in the
range of 452491 Ma (Llambas et al. 1991; Steenken et al.
2006). In conclusion, on the available geochronological information, it seems likely that the intrusion of the granite took place
during the Famatinian amphibolite- or granulite-facies metamorphic and deformational episode. This assumption is favoured
also by the fact that the foliated and concordant granite parallels
the S2 compressional fabrics. A possible Neoproterozoic age of
the crustal-derived granitoids of the Pringles Metamorphic Complex as was suggested by conventional U/Pb dating (von Gosen
et al. 2002) appears unlikely in view of the new data.
Another outcome from U/Pb SHRIMP dating with respect to
the crystallization ages of the various intrusions is that, despite
their variable degree of pervasive deformation, all the crystallization ages fall into a narrow time gap within the Early to MidOrdovician. This apparent contradiction is most probably explained by differences in magma viscosity and rheological
aspects. The low water content in a tonalitic melt will open the
possibility for the little fractionated melt to reach higher crustal
levels, where it is exposed to different deformation mechanisms
compared with the deeper-seated Paso de Rey granite. This
assumption is substantiated by the fact that the tonalitic to
granodioritic intrusions are restricted to the low-grade metamorphic belts of the San Luis formation. The contemporaneous
intrusion of the two granitoid suites in the same geotectonic
setting was demonstrated by Lopez de Luchi et al. (2006). A
similar study, based on geochemical and geochronological
results, was presented by Pankhurst et al. (2000) for the
tonalitetrondhjemitegranodiorite, I and S-type plutonic
suites of the Sierra de Chepes. Those workers demonstrated that,
in spite of different origins of the melts, all the intrusions
invaded the Pampean basement within a narrow time frame
between 468 and 499 Ma.
Nogol Metamorphic Complex. A uncertain pre-Famatinian deformational and metamorphic history for this complex was
proposed based on structural interference patterns (von Gosen &
Prozzi 1998; Gonzalez et al. 2002a, 2004; Sato et al. 2003b) and
one reported conventional U/Pb age at c. 554  5 Ma for a
crustal derived granitoid in the north of the complex (Vujovich
& Ostera 2003). The new 208 Pb/232 Th SHRIMP results from the
monazites of a migmatite from the northern part of the complex
(sample A 39-01: 478  4 Ma) point to a Famatinian (i.e. Early
Ordovician) metamorphic peak (Fig. 6). No inheritance of a
previous (Pampean?) metamorphism was found. The Famatinian
metamorphism is causal for the development of the migmatites.
Microstructures within these migmatites record a first planar
metamorphic fabric (S1 ) that was subsequently overgrown by
large muscovite plates (Fig. 10). The migmatites interfere with
the resumption of compression and folding around steeply
inclined fold axes during D2 (see von Gosen & Prozzi 1998).
Consequently, the establishment of the pervasive fabrics has to

979

be attributed to Famatinian compression, instead of a Pampean


origin as was proposed by von Gosen & Prozzi (1998). Ductile
folding without development of a new metamorphic plane but
with complete resetting of the U/Th system of monazite, as
would be required to maintain the interpretation of the Pampean
origin of fabrics, turns out to be highly unlikely.
Famatinian metamorphism was recorded by garnets from the
leucocratic monzogranite (sample A 14-02) of the southern part
of the Nogol Metamorphic Complex. The 207 Pb/206 Pb age of
460  12 Ma is in agreement with previous results on the
geochronological history (Fig. 6; Gonzalez et al. 2002a, 2004).
In contrast to the analysed sample from the northern part, the
regional NNE-trending planar fabrics would correspond to the
Famatinian compression (Sato et al. 2003b; Gonzalez et al.
2004) and the age cannot be used to rule out a possible Pampean
evolution within this area of the complex, although undeniable
absolute time constraints for the Pampean evolution have not
been reported yet.
Conlara Metamorphic Complex. The structural evolution of the
Conlara Metamorphic Complex was previously considered by
e.g. Lopez de Luchi (1986), Llaneza & Ortz Suarez (2000) and
Lopez de Luchi & Cerredo (2001). According to the first two
studies, the banding of the widespread fine-grained gneisses of
this complex is the result of a D2 axial-plane foliation. Folded
xenoliths of these banded gneisses are observed in the pervasively solid-state deformed granitoid intrusions such as the El
Salado granodiorite in the vicinity of the El Penon granite (Fig.
6). Microstructures within the latter intrusion denote continuity
from magmatic to high-temperature solid-state deformation, indicating the synkinematic emplacement of the pluton with respect
to D3. Geochronological constraints suggest a late Cambrian
(497  8 Ma) emplacement of the El Penon pluton, marking the
earliest stage of Famatinian compression within this basement
domain (Fig. 10). Consequently, it may be proposed that the D1
D2 history took place during the Pampean Orogeny (see
Steenken et al. 2005). This contrasts the assumption that the
entire evolution of the Sierra de San Luis could be related to
post-Pampean events (Whitmeyer & Simpson 2004).

Conclusions
The combination of SHRIMP U/Pb zircon and monazite dating
and PbSL experiments on garnet and staurolite with structural
observations at the macroscopic to microscopic scale provides
crucial absolute time constraints as well as substantiation of
previous results on the structural and metamorphic evolution of
the basement complexes of the Sierra de San Luis. This
evolution is the result of multiple depositions of psammopelitic
sediments involved in the distinct compressional events (i.e. the
Pampean, Famatinian and Achalian tectonic cycles) that occurred
on the Palaeozoic margin of southwestern Gondwana.
The crystallization ages of the various Ordovician granitoid
intrusions allow no separation into a pre- and syn-orogenic group
with respect to the Famatinian orogenic cycle. It has been shown
that the S2 fabric formation within the Pringles Metamorphic
Complex (i.e. the S1 axial-plane foliation within the phyllites of
the San Luis Formation) corresponds to the emplacement of the
Ordovician tonalites and granites during the Early to MidOrdovician. Their synkinematic nature has been inferred from
the continuity of (sub-)magmatic to high-temperature solid-state
microstructural indications. Differences in the structural integration with their host have to be related to variations in magma
viscosity and rheological behaviour of the host rocks.

980

A. STEENKEN ET AL.

Within the Pringles basement the granulite- to amphibolitefacies metamorphism might have lasted for c. 46 Ma, starting at
498 Ma as suggested by U/Pb zircon data. The capture of the
lower limit of the metamorphism at c. 428 Ma by PbSL staurolite
data is probably related to Pb diffusion during amphibolite-facies
conditions. The younger metamorphic peak in the Nogol
Metamorphic Complex is constrained from 208 Pb/232 Th data on
monazite at c. 480 Ma and from PbSL garnet data at c. 460 Ma.
This metamorphic peak corresponds to the D2 folding within this
complex, emphasizing an entirely Famatinian history of the
complex. No geochronological evidence for a pre-Famatinian
evolution of the Nogol Metamorphic Complex was found.
The Early Ordovician metamorphic peak within the basement
domains of the Sierra de San Luis predates the metamorphic peak
in the eastern sector of the Cuyania Terrane (Casquet et al.
2001), which was considered to slightly post-date the collision of
this terrane with Gondwana. Therefore, magmatism and metamorphism within the Sierra de San Luis are instead related to
foreland compression (i.e. closure of an extensional basin in the
west of the Pampean orogen as a result of the continuous
subduction), and there is no need to invoke a stage of crustal
thickening related to continental collision in the area of San Luis.
Within the Conlara Metamorphic Complex the structural
relation of the El Penon pluton with its host denotes the preFamatinian evolution of the basement. The appearance of folded
xenoliths (D3 ) demands a pre-Famatinian (i.e. Pampean) deformation and metamorphism of the eastern basement domain of
the Sierra de San Luis. Sub-magmatic to high-temperature solidstate microstructures within the granite point to its synkinematic
emplacement with respect to D3.
Provenance results indicate clearly separated sedimentary
basins for the psammopelitic precursors of the Conlara and
Pringles Metamorphic Complexes. The detrital zircon cores of
the metasediments of the latter denote an early to mid-Cambrian
deposition. The inheritance pattern indicates a source on the
nearby Pampean orogen. These sediments were accommodated in
an extensional basin on the western outboard of the Pampean
block that formed as a result of the resumption of convergence
and subduction along the margin of Gondwana.
Inherited zircon cores within the Conlara Metamorphic Complex indicate a Gondwana provenance. Deposition of the clastic
sequence took place after 587 Ma during the Ediacaran, and most
probably represents higher-grade metamorphic equivalents of the
Puncoviscana Formation, suggesting that the proto-Andean basement of a part of the Eastern Sierras Pampeanas belonged to a
common sedimentary basin at this time.
We are grateful for the German Science Foundation (DFG) Grant Si 438/
16-1 that funded our research project in central Argentina. A.S. is grateful
for the Feodor-Lynen Fellowship V.3/FLF/1116298 granted by the Alexander von Humboldt-Foundation. The support by the Danish Natural
Science Research Council (SNF) via grant 56943 is gratefully acknowledged by R.F. For valuable and fruitful discussion of the analytical data
we thank R. Armstrong (Canberra). The manuscript was substantially
improved by reviews by V. Ramos (Bs. As.) and an anonymous reviewer.

Appendix
SHRIMP U/Pb dating
All zircons and monazites were mounted in epoxy at the
Research School of Earth Sciences (RSES, Canberra), together
with the RSES reference zircons FC1 and SL13 and the
Thompson Mine WB.T.329 monazite standard (age 1768 Ma, U
concentration 2100 ppm). Photomicrographs in transmitted and

reflected light were taken of all zircons and monazites. These,


together with SEM CL and BSE images, respectively, were used
to decipher the internal structures of the sectioned grains and to
target specific areas within the accessories (i.e. metamorphic
rims or inherited cores) using a 46 nA primary O2 ion beam
with an c. 25 m diameter spot.
The majority of the UPb zircon analyses were carried out in
a single session on the SHRIMP II with a follow-up session on
the SHRIMP RG (sample A 59-01). For the zircon calibration
the Pb/U ratios were normalized relative to a value of 0.1859 for
the 206 *Pb/238 U ratio of FC1 reference zircons, equivalent to an
age of 1.099 Ga (Paces & Miller 1993). U and Th concentrations
were determined relative to the SL13 standard. The error in the
standard calibration was 0.25% on the SHRIMP II and 0.12% on
the SHRIMP RG.
For the monazite, secondary ions were analysed at a resolution
of c. 5100 with the data collected through six scans of the
masses 203 CePO2, 204 Pb, background, 206 Pb, 207 Pb, 208 Pb, 238 U,
248
ThO and 254 UO. To reduce any isobaric interference on mass
204
Pb, 50% energy filtering was applied to the secondary beam.
Seventeen standard analyses were carried out, producing a 2
error of the (weighted) mean Pb/U calibration of 0.64%. Correction for Th/U fractionation was calculated directly using the
method of Williams (1998), and all other data reduction utilized
the SQUID software of Ludwig (2000).
Uncertainties given for individual analyses (ratios and ages) are
at the 1 level; however, uncertainties in any calculated weighted
mean ages or concordia ages (Ludwig 1998) are reported as 95%
confidence limits (unless stated otherwise) and include the
uncertainties in the standard calibrations where appropriate.
Concordia plots, regressions and weighted mean age calculations
were carried out using Isoplot/Ex 3.0 (Ludwig 2003).

Pb stepwise leaching (PbSL)


Pb/Pb stepwise leaching experiments (PbSL) were applied to one
staurolite and two garnet separates. Slightly modified procedures
of Frei & Kamber (1995) were applied, in an effort to date the
formation ages of these important mineral phases. Mineral
separates were obtained by standard techniques using jaw crusher
and sieve kit. The 150250 m sieve fractions were then purified
by handpicking followed by repeated rinsing in deionized water,
and 200 mg of these materials were transferred to 7 ml Savillex1
screw-cap beakers for step-leaching. Successive 120 8C acidleach steps (seven in total) involving various mixtures of HBr,
HNO3 and HF were performed on each separate, to extract Pb
selectively from the phases. The leaching schemes are given in
Table 3. Purified Pb extracts were mounted on Re filaments and
Pb isotopic ratios were determined by thermal ionization mass
spectrometry (TIMS) at the University of Copenhagen. A similar
PbSL technique was previously applied to garnet in metapelite
(Dahl & Frei 1998). All PbSL ages are discussed in the text and
depicted graphically at the 2 level (95% confidence limits).
Fractionation for Pb was controlled by repeated analysis of the
NBS 981 standard and amounted to 0.103  0.007% per a.m.u.
(2; n 5) relative to the values proposed by Todt et al. (1993).
Procedural blanks for Pb remained below 87 pg, an amount that
insignificantly affects the isotopic data for the samples.

References
Acenolaza, F.G., Miller, H. & Toselli, A.J. 2002. ProterozoicEarly Paleozoic
evolution in western South Americaa discussion. Tectonophysics, 354,
121137.

FA M AT I N I A N G E O DY NA M I C E VO L U T I O N O F G O N DWA NA
Astini, R.A. 1996. Las fases diastroficas del Paleozoico medio en la Precordillera
del oeste argentinoevidencias estratigraficas13th. In: 13th Congreso Geologico Argentino & 3th Congreso de Exploracion de Hidrocarburos, Actas. 5,
509526.
Astini, R.A. 1998. Stratigraphical evidence supporting the rifting, drifting and
collision of the Laurentian Precordillera terrane of western Argentina.
In: Pankhurst, R.J. & Rapela, C.W. (eds) The Proto-Andean Margin of
Gondwana. Geological Society, London, Special Publications, 142, 1133.
Astini, R.A. & Davila, F.M. 2004. Ordovician back arc foreland and Ocloyic
thrust belt development on the western Gondwana margin as a response to
Precordillera terrane accretion. Tectonics, 23, doi:10.1029/2003TC001620.
Astini, R.A. & Thomas, W.A. 1999. Origin and evolution of the Precordillera
terrane of western Argentina: a drifted Laurentian orphan. In: Ramos, V.A. &
Keppie, J.D. (eds) LaurentiaGondwana Connections before Pangea. Geological Society of America, Special Papers, 336, 120.
Astini, R.A., Benedetto, J.L. & Vaccari, N.E. 1995. The early Paleozoic
evolution of the Argentine Precordillera as a Laurentian rifted, drifted, and
collided terrane: a geodynamic model. Geological Society of America
Bulletin, 107, 253273.
Bock, B., Bahlburg, H., Worner, G. & Zimmermann, U. 2000. Tracing crustal
evolution in the Southern Central Andes from Late Precambrian to Permian with
geochemical and Nd and Pb isotope data. Journal of Geology, 108, 515535.
Bucher, K. & Frey, M. 1994. Petrogenesis of Metamorphic Rocks, 6th. Springer,
Berlin.
Casquet, C., Baldo, E., Pankhurst, R.J., Rapela, C.W., Galindo, C.,
Fanning, C.M. & Saavedra, J. 2001. Involvement of the Argentine
Precordillera terrane in the Famatinian mobile belt: UPb SHRIMP and
metamorphic evidence from the Sierra de Pie de Palo. Geology, 29, 703706.
Chernicoff, C.J. & Ramos, V.A. 2003. El basamento de la sierra de San Luis:
nuevas evidencias magneticas y sus implicancias tectonicas. Revista de la
Asociacion Geologa Argentina, 58, 511524.
Collo, G., Astini, R.A., Cawood, P.A. & Buchan, C. 2005. Preliminary detrital
ages and stratigraphy for the basement in Famatina, proto-Andean margin of
Gondwana Gondwana. In: Pankhurst, R.J & Veiga, G.D. (eds) Gondwana
12, Abstracts. 108.
Dahl, P.S. & Frei, R. 1998. Step-leach Pb-Pb dating of inclusion-bearing garnet
and staurolite, with implications for Early Proterozoic tectonism in the Black
Hills collisional orogen, South Dakota, United States. Geology, 26, 111114.
Dalla Salda, L.H., Lopez de Luchi, M.G., Cingolani, C. & Varela, R. 1998.
LaurentiaGondwana collision: the origin of the FamatinianAppalachians
Orogenic Belt. In: Pankhurst, R.J. & Rapela, C.W. (eds) The ProtoAndean Margin of Gondwana. Geological Society, London, Special Publications, 142, 219234.
Escayola, M.P. & Pimentel, M. 2006. A Neoproterozoic back-arc basin:
SHRIMP UPb and Sm/Nd isotopic evidence from the Eastern Pampean
Ranges, Argentina. Geology, in press.
Finney, S., Peralta, S., Gehrels, G. & Marsaglia, K. 2005. The Early
Paleozoic history of the Cuyania (greater Precordillera) terrane of western
Argentina: evidence from geochronology of detrital zircons from Middle
Cambrian sandstones. Geologica Acta, 3, 339354.
Frei, R. & Kamber, B.S. 1995. Single mineral PbPb dating. Earth and Planetary
Science Letters, 129, 261268.
Frei, R., Villa, I.M. & Nagler, T.F. et al. 1997. Single mineral dating by the
PbPb step-leaching method: assessing the mechanisms. Geochimica et
Cosmochimica Acta, 61, 393414.
Gonzalez, P.D., Sato, A.M., Basei, M.A.S., Vlach, S.R.F. & Llambas, E.
2002a. Structure, metamorphism and age of the PampeanFamatinian
orogenies in the western Sierra de San Luis. In: Cabaleri, N., Linares, E.,
Lopez de Luchi, M.G., Ostera, H. & Panarello, H. (eds) 15th Congreso
Geologico Argentino, Actas. 2, 5156.
Gonzalez, P. D., Sato, A. M. & Llambas, E. J. 2002b. The komatiites and
associated mafic to ultramafic metavolcanic rocks of western Sierra de San
Luis. In: Cabaleri, N., Linares, E., Lopez de Luchi, M.G., Ostera, H. &
Panarello, H. (eds) 15th Congreso Geologico Argentino, Actas. Calafate,
Argentina, 2, 8790.
Gonzalez, P.D., Sato, A.M., Llambas, E.J., Basei, M.A.S. & Vlach, S.R.F.
2004. Early Paleozoic structural and metamorphic evolution of western Sierra
de San Luis (Argentina), in relation to Cuyania Accretion. Gondwana
Research, 7, 11571170.
Hauzenberger, C.A., Mogessie, A. & Hoinkes, G. et al. 2001. Metamorphic
evolution of the Sierras de San Luis, Argentina; granulite facies metamorphism related to mafic intrusions. Mineralogy and Petrology, 71, 95126.
Janasi, V.A., Leite, R.J. & van Schmus, W.R. 2001. UPb chronostratigraphy of
the granitic magmatism in the Agudos Grandes Batholith (west of Sao Paulo,
Brazil)implications for the evolution of the Ribeira Belt. Journal of South
American Earth Sciences, 14, 363376.
Jezek, P., Willner, A.P., Acenolaza, F. & Miller, H. 1985. The Puncoviscana
trough: a large basin of Late Precambrian to Early Cambrian Age on the

981

Pacific edge of the Brazilian shield. Geologische Rundschau, 74, 573584.


Jordan, T.E. & Allmendinger, R.W. 1986. The Sierras Pampeanas of Argentina:
a modern analogue of Rockey Mountain foreland deformation. American
Journal of Science, 286, 737764.
Keller, M. (ed.) 1999. Argentine Precordillera: Sedimentary and Plate Tectonic
History of a Laurentian Crustal Fragment in South America. Geological
Society of America, Special Papers, 341.
Keller, M., Buggish, W. & Lehnert, O. 1998. The stratigraphic record of
Argentine Precordillera and its plate tectonic background. In: Pankhurst,
R.J. & Rapela, C.W. (eds) The Proto-Andean Margin of Gondwana.
Geological Society, London, Special Publications, 142, 3556.
Kruhl, J.H. 1996. Prism- and basal-plane parallel subgrain boundaries in quartz: a
microstructural geothermobarometer. Journal of Metamorphic Geology, 14,
581589.
Leite, J.A.D., Hartmann, L.A. & Fernandes, L.A.D. et al. 2000. Zircon UPb
SHRIMP dating of gneissic basement of the Dom Feliciano Belt, southernmost Brazil. Journal of South American Earth Sciences, 13, 739750.
Llambas, E.J., Cingolani, C.A. & Varela, R. et al. 1991. Leucogranodioritas
sincinematicas ordovcicas en la Sierra de San Luis, Republica Argentina. In:
9th Congreso Geologico Chileno, Actas. 187191.
Llambas, E.J., Sato, A., Ortiz Suarez, A. & Prozzi, C. 1998. In: Pankhurst,
R.J. & Rapela, C.W. (eds) The Proto-Andean Margin of Gondwana.
Geological Society, London, Special Publications, 142, 325341.
Llambas, E.J., Gregori, D., Basei, M.A.S., Varela, R. & Prozzi, C. 2003.
Ignimbritas riolticas neoproterozoicas en la Sierra Norte de Cordoba:
evidencia de un arco magmatico temprano en el ciclo Pampeano? Revista de
la Asociacion Geologia Argentina, 58, 572582.
Llaneza, G. & Ortz Suarez, A. 2000. Geologa y petrografa del Granito El
Penon (Provincia de San Luis) y su relacion con el metamorfismo y la
deformacion. In: 9th Congreso Geologico Chileno, Actas. 1, 639643.
Lopez de Luchi, M. G. 1986. Geologa y petrologa del basamento de la Sierra de
San Luis, region del Batolito de Renca. PhD thesis, Universidad de Buenos
Aires.
Lopez de Luchi, M.G. & Cerredo, M.E. 2001. Submagmatic and solid-state
microstructures in La Tapera pluton. San Luis, Argentina. In: Cortes, J.M.,
Rossello E. & Dalla Salda, L. (eds) Advances en Microtectonica.
Comision de Tectonica de la Asociacion Geologica Argentina (ComTec) y
Grupo de Trabajo en Microtectonica. Publicaciones Especiales de la
Asociacion Geologia Argentina, Serie D, 5, 121126.
Lopez de Luchi, M.G., Rapalini, A.E., Rossello, E. & Geuna, S.E. 2002. Rock
and magnetic fabric of the Renca batholith (Sierra de San Luis, Argentina):
constraints on its emplacement. Lithos, 61, 161186.
Lopez de Luchi, M.G., Cerredo, M.E., Siegesmund, S., Steenken, A. &
Wemmer, K. 2003. Provenance and tectonic setting of the protoliths of the
metamorphic complexes of Sierra de San Luis. Revista de la Asociacion
Geologia Argentina, 58, 525540.
Lopez de Luchi, M.G., Rapalini, A.E., Siegesmund, S. & Steenken, A. 2004.
Application of magnetic fabrics to the emplacement and tectonic history of
Devonian granitoids in Central Argentina. In: Martn-Hernandez, F.,
Luneburg, C., Aubourg, C. & Jackson, M. (eds) Magnetic Fabric:
Methods and applications. Geological Society, London, Special Publications,
238, 447474.
Lopez de Luchi, M.G., Siegesmund, S., Wemmer, K., Steenken, A. &
Naumann, R. 2006. Paleozoic tectonomagmatic evolution of the Sierra de
San Luis (Sierras Pampeanas, Argentina). Journal of South American Earth
Sciences, in press.
Ludwig, K.R. 1998. On the treatment of concordant uraniumlead ages.
Geochimica et Cosmochima Acta, 62, 665676.
Ludwig, K.R. 1990. ISOPLOT: a plotting and regression program for radiogenic
isotope data for IBM-PC compatible computers. US Geological Survey OpenFile Report 88-557.
Ludwig, K.R. 2000. SQUID 1.00, A Users Manual. Berkeley Geochronology
Center Special Publications, 2.
Ludwig, K.R. 2003. Isoplot/Ex Version 3.00: a Geochronological Toolkit for
Microsoft Excel. Berkeley Geochronology Center Special Publications, 4.
Martino, R.D. 2003. Ductile deformation shear belts at Pampean Ranges near
Cordoba. Revista de la Asociacion Geologia Argentina, 58, 549571.
Paces, J.B. & Miller, J.D. 1993. Precise UPb ages of Duluth Complex and
related mafic intrusions, Northeastern Minnesota: geochronological insights
to physical, petrogenic, paleomagnetic and tectonomagmatic processes
associated with the 1.1 Ga midcontinent rift system. Journal of Geophysical
Research, 98B, 1399714013.
Pankhurst, R.J., Rapela, C.W., Saavedra, J., Baldo, E., Dahlquist, J.,
Pascua, I. & Fanning, C.M. 1998. The Famatinian magmatic arc in the
central Sierras Pampeanas: an Early to Mid-Ordovician continental arc on the
Gondwana. In: Pankhurst, R.J. & Rapela, C.W. (eds) The Proto-Andean
Margin of Gondwana. Geological Society, London, Special Publications, 142,
343367.

982

A. STEENKEN ET AL.

Pankhurst, R.J., Rapela, C.W. & Fanning, C.M. 2000. Age and origin of coeval
TTG, I- and S-type granites in the Famatinian belt of NW Argentina.
Transactions of the Royal Society of Edinburgh, Earth Sciences, 91, 151168.
Paterson, S.R., Fowler, T.K., Schmidt, K.L., Yoshinobu, A.S., Yuan, E.S. &
Miller, R.B. 1998. Interpreting magmatic fabric patterns in plutons. Lithos,
44, 5382.
Prozzi, C.R. & Ramos, G. 1988. La Formacion San Luis. Primeras Jornadas de
Trabajo de Sierras Pampeanas (San Luis, 2426 Agosto, 1988), Abstracts.
Universidad Nacional de San Luis, San Luis.
Quenardelle, S. & Ramos, V.A. 1999. Ordovician western Sierras Pampeanas
magmatic belt: record of Precordillera accretion in Argentina. In: Ramos,
V.A. & Keppie, J.D. (eds) LaurentiaGondwana Connections before Pangea.
Geological Society of America, Special Papers, 336, 6386.
Ramos, V.A. 1988. Late ProterozoicEarly Paleozoic of South Americaa
collisional history. Episodes, 11, 168174.
Ramos, V.A. & Basei, M.A.S. 1997. The basement of Chilenia: an exotic
continental terrane to Gondwana during the early Paleozoic. In: Bradshaw,
J.D. & Weaver, S.D. (eds) Terrane Dynamics97, International Conference
on Terrane Geology. Conference Abstracts. University of Canterbury,
Christchurch, 97, 140143.
Ramos, V.A., Jordan, T.E., Allmendinger, W., Mpodozis, C., Kay, S.M.,
Cortes, J.M. & Palma, M. 1986. Paleozoic terranes of the Central
ArgentineChilean Andes. Tectonics, 5, 855880.
Rapalini, A.E., Astini, R.A. & Conti, C.M. 1999. Paleomagnetic constraints on
the tectonic evolution of Paleozoic suspect terranes from southern South
America. In: Ramos, V.A. & Keppie, J.D. (eds) LaurentiaGondwana
Connections before Pangea. Geological Society of America, Special Papers,
336, 171182.
Rapela, C.W. 2000. The Sierras Pampeanas of Argentina: Paleozoic building of
the southern Proto-Andes. In: Cordani, U.G., Milani, E.J., Thomaz-Filho,
A. & Campos, D.A. (eds) Tectonic Evolution of South America. 31th
International Geological Congress, Ro de Janeiro. 381387.
Rapela, C.W., Pankhurst, R.J., Casquet, C., Baldo, E., Saavedra, J. &
Galindo, C. 1998a. Early evolution of the Proto-Andean margin of South
America. Geology, 26, 707710.
Rapela, C.W., Pankhurst, R.J., Casquet, C., Baldo, E., Saavedra, J.,
Galindo, C. & Fanning, C.M. 1998b. The Pampean Orogeny of the
southern proto-Andes; Cambrian continental collision in the Sierras de
Cordoba. In: Pankhurst, R.J. & Rapela, C.W. (eds) The Proto-Andean
Margin of Gondwana. Geological Society, London, Special Publications, 142,
181217.
Sato, A.M., Gonzalez, P.D. & Sato, K. 2001. First indication of Mesoproterozoic age from the western basement of Sierra de San Luis, Argentina. In:
Editor, A. (ed.) 3rd South American Symposium on Isotope Geology,
Extended Abstracts. 6467.
Sato, A.M., Gonzalez, P.D., Basei, M.A.S., Passarelli, C.R., Tickyj, H. &
Ponce, J.M. 2003a. The Famatinian granitoids of southwestern Sierra de San
Luis, Argentina. In: Editor, A. (ed.) 4th South American Symposium on
Isotope Geology, Short Papers. 290293.
Sato, A.M., Gonzalez, P.D. & Llambas, E.J. 2003b. Evolution of the
Famatinian orogen in the Sierra de San Luis: arc magmatism, deformation,
and low to high-grade metamorphism. Revista de la Asociacion Geologica
Argentina, 58, 487504.
Sato, A.M., Gonzalez, P.D. & Basei, M.A.S. 2005. Los ortogneises granodiorticos del complejo metamorfico Nogol, Sierra de San Luis. In: Editor, A.
(ed.) 16th Congreso Geologico Argentino. 1, 4148.
Schaller, M., Steiner, O., Studer, I., Frei, R. & Kramers, J.D. 1997. Pb
stepwise leaching (PbSL) dating of garnet - addressing the inclusion problem.
Swiss Bulletin of Mineralogy and Petrology, 77, 113121.
Schwartz, J.J. & Gromet, L.P. 2004. Provenance of a late Proterozoicearly
Cambrian basin, Sierras de Cordoba, Argentina. Precambrian Research, 129,
121.
Siegesmund, S., Steenken, A., Lopez de Luchi, M.G., Wemmer, K.,
Hoffmann, A. & Mosch, S. 2004. The Las Chacras-Potrerillos Batholith:
Structural Evidences on its Emplacement and Timing of the Intrusion.
International Journal of Earth Sciences, 93, 2343.
Sims, J.P., Skirrow, R.G., Stuart-Smith, P.G. & Lyons, P. 1997. Informe
geologico y metalogentico de las Sierras de San Luis y Comechingones
(provincias de San Luis y Cordoba), 1:250 000. IGRM, SEGEMAR, Buenos
Aires, Anales, 28.
Sims, J.P., Ireland, T.R., Camacho, A., Lyons, P., Skirrow, R.G., StuartSmith, P.G. & Miro, R. 1998. UPb, ThPb and ArAr geochronology
from southern Sierras Pampeanas, Argentina: implications for the Palaeozoic
tectonic evolution of the western Gondwana margin. In: Pankhurst, R.J. &
Rapela, C.W. (eds) The Proto-Andean Margin of Gondwana. Geological

Society, London, Special Publications, 142, 235258.


Sollner, F., de Brodtkorb, M.K., Miller, H., Pezzutti, N.E. & Fernandez,
R.R. 2000. UPb zircon ages of metavolcanic rocks from the Sierra de San
Luis, Argentina. Revista de la Asociacion Geologica Argentina, 55, 1522.
Steenken, A., Lopez de Luchi, M.G., Siegesmund, S., Wemmer, K. & Pawlig,
S. 2004. Crustal provenance and cooling of the basement complexes of the
Sierra de San Luis: an insight into the tectonic history of the proto-Andean
margin of Gondwana. Gondwana Research, 7, 11711195.
Steenken, A., Lopez de Luchi, M.G., Martino, R.D., Siegesmund, S. &
Wemmer, K. 2005. SHRIMP dating of the El Penon granite: a time marker at
the turning point between the Pampean and Famatinian cycles within the
Conlara Metamorphic Complex (Sierra de San Luis; Argentina)? In:
Llambas, E., Barrio, R. de, Gonzalez, P. & Leal, P. (eds) 16th
Congreso Geologico Argentino. 1, 889896.
Steenken, A., Siegesmund, S., Wemmer, K. & Lopez de Luchi, M.G. 2006.
Time constraints on the Famatinian and Achalian structural evolution of the
basement of the Sierra de San Luis (Eastern Sierras Pampeanas, Argentina).
Journal of South American Earth Sciences, in press.
Stewart, J.H., Gehrels, G.E., Barth, A.P., Link, P.K., Christie-Blick, N. &
Wrucke, C.T. 2001. Detrital zircon provenance of Mesoproterozoic to
Cambrian arenites in the western United States and northwestern Mexico.
Geological Society of America Bulletin, 113, 13431356.
Stuart-Smith, P.G., Camacho, A. & Sims, J.P. et al. 1999. Uraniumlead
dating of felsic magmatic cycles in the southern Sierras Pampeanas,
Argentina: implications for the tectonic development of the proto-Andean
Gondwana margin. In: Ramos, V.A. & Keppie, J.D. (eds) Laurentia
Gondwana Connections before Pangea. Geological Society of America,
Special Papers, 336, 87114.
Thomas, W. & Astini, R. 2003. Ordovician accretion of the Argentine
Precordillera terrane to Gondwana: a review. Journal of South American
Earth Sciences, 16, 6779.
Thomas, W.A., Astini, R.A. & Bayona, G. 2002. Ordovician collision of the
Argentine Precordillera with Gondwana, independent of Laurentian Taconic
orogeny. Tectonophysics, 345, 131152.
Todt, W., Cliff, R.A., Hanser, A. & Hofmann, A.W. 1993. Recalibration of
NBS lead standards using a 202 Pb/205 Pb double spike. Terra Abstracts, 5, 396.
von Gosen, W. 1998a. The phyllite and micaschist group with associated
intrusions in the Sierras de San Luis (Sierras Pampeanas/Argentina)
structural and metamorphic relations. Journal of South American Earth
Sciences, 11, 79109.
von Gosen, W. 1998b. Transpressive deformation in the southwestern part of the
Sierra de San Luis (Sierras Pampeanas, Argentina). Journal of South
American Earth Sciences, 3, 233264.
von Gosen, W. & Prozzi, C.R. 1996. Geology, structure and metamorphism in
the area south of La Carolina (Sierra de San Luis, Argentina). In: 13th
Congreso Geologico Argentino & 3rd Congreso de Exploracion de Hidrocarburos, Actas. 2, 301314.
von Gosen, W. & Prozzi, C.R. 1998. Structural evolution of the Sierra de San
Luis (Eastern Sierras Pampeanas, Argentina): implications for the ProtoAndean margin of Gondwana. In: Pankhurst, R.J. & Rapela, C.W. (eds)
The Proto-Andean Margin of Gondwana. Geological Society, London, Special
Publications, 142, 235258.
von Gosen, W., Loske, W. & Prozzi, C.R. 2002. New isotopic dating of intrusive
rocks in the Sierra de San Luis (Argentina): implications for the geodynamic
history of the Eastern Sierras Pampeanas. Journal of South American Earth
Sciences, 15, 237250.
Vujovich, G.I., & Ostera, H.A. 2003. Evidence of the Pampean cycle in the
basement of the northwestern sector of the Sierra de San Luis. Revista de la
Asociacion Geologa Argentina, 58, 541548.
Wareham, C.D., Pankhurst, R.J., Thomas, R.J., Storey, B.C., Grantham,
G.H., Jacobs, J. & Eglington, B.M. 1998. Pb, Nd, Sr isotope mapping of
Grenville-age crustal provinces in Rodinia. Journal of Geology, 106,
647659.
Whitmeyer, S.J. & Simpson, C. 2004. Regional deformation of the Sierra de San
Luis, Argentina: implications for the Paleozoic development of western
Gondwana. Tectonics, 23, doi:10.1029/2003TC001542.
Williams, I.S. 1998. UThPb geochronology by ion microprobe. In: McKibben,
M.A., Shanks, W.C.P. III & Ridley, W.I. (eds) Applications of Microanalytical Techniques to Understanding Mineralizing Processes. Reviews in
Economic Geology, 7, 135.
Zimmermann, U. 2005. Provenance studies of very low- to low-grade metasedimentary rocks of the Puncoviscana Formation in Northwest Argentina. In:
Vaughan, A.P.M., Leat, P.T. & Pankhurst, R.J. (eds) Terrane Processes
at the Margins of Gondwana. Geological Society, London, Special Publications, 246, 381416.

Received 3 May 2005; revised typescript accepted 14 March 2006.


Scientific editing by Mike Villeneuve

Das könnte Ihnen auch gefallen