You are on page 1of 28

Downloaded from gsabulletin.gsapubs.

org on January 20, 2011

Geological Society of America Bulletin


Tectonomagmatic evolution of Western Amazonia: Geochemical
characterization and zircon U-Pb geochronologic constraints from the
Peruvian Eastern Cordilleran granitoids
Aleksandar Miskovic, Richard A. Spikings, David M. Chew, Jan Kosler, Alexey Ulianov and Urs
Schaltegger
Geological Society of America Bulletin 2009;121;1298-1324
doi: 10.1130/B26488.1

Email alerting services

click www.gsapubs.org/cgi/alerts to receive free e-mail alerts when new


articles cite this article

Subscribe

click www.gsapubs.org/subscriptions/ to subscribe to Geological Society of


America Bulletin

Permission request

click http://www.geosociety.org/pubs/copyrt.htm#gsa to contact GSA

Copyright not claimed on content prepared wholly by U.S. government employees within scope of
their employment. Individual scientists are hereby granted permission, without fees or further
requests to GSA, to use a single figure, a single table, and/or a brief paragraph of text in subsequent
works and to make unlimited copies of items in GSA's journals for noncommercial use in classrooms
to further education and science. This file may not be posted to any Web site, but authors may post
the abstracts only of their articles on their own or their organization's Web site providing the posting
includes a reference to the article's full citation. GSA provides this and other forums for the
presentation of diverse opinions and positions by scientists worldwide, regardless of their race,
citizenship, gender, religion, or political viewpoint. Opinions presented in this publication do not reflect
official positions of the Society.

Notes

2009 Geological Society of America

Downloaded from gsabulletin.gsapubs.org on January 20, 2011

Tectonomagmatic evolution of Western Amazonia:


Geochemical characterization and zircon U-Pb geochronologic
constraints from the Peruvian Eastern Cordilleran granitoids
Aleksandar Mikovi1,*,, Richard A. Spikings1, David M. Chew2, Jan Koler3, Alexey Ulianov4, and Urs Schaltegger1
1

Department of Mineralogy, Earth Sciences Section, University of Geneva, 13 rue des Marachers, CH-1205 Geneva, Switzerland
Department of Geology, Trinity College Dublin, College Green, Dublin 2, Ireland
3
Department of Earth Sciences, University of Bergen, Allegaten 41, N 5007 Bergen, Norway
4
Institute of Mineralogy and Petrography, University of Lausanne, BFSH 2, CH-1015 Lausanne, Switzerland
2

ABSTRACT
The results of a coupled, in situ laser
ablationinductively coupled plasmamass
spectrometry (LA-ICP-MS) U-Pb study on
zircon and geochemical characterization of
the Eastern Cordilleran intrusives of Peru
reveal 1.15 Ga of intermittent magmatism
along central Western Amazonia, the Earths
oldest active open continental margin. The
eastern Peruvian batholiths are volumetrically dominated by plutonism related to
the assembly and breakup of Pangea during the Paleozoic-Mesozoic transition. A
Carboniferous-Permian (340285 Ma) continental arc is identified along the regional
orogenic strike from the Ecuadorian border
(6S) to the inferred inboard extension of
the Arequipa-Antofalla terrane in southern
Peru (14S). Widespread crustal extension
and thinning, which affected western Gondwana throughout the Permian and Triassic
resulted in the intrusion of the late- to posttectonic La MercedSan Ramn-type anatectites dated between 275 and 220 Ma, while
the emplacement of the southern Cordillera
de Carabaya peraluminous granitoids in the
Late Triassic to Early Jurassic (220190 Ma)
represents, temporally and regionally, a separate tectonomagmatic event likely related to
resuturing of the Arequipa-Antofalla block.
Volcano-plutonic complexes and stocks associated with the onset of the present Andean
cycle define a compositionally bimodal

*Current address: Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, 77 Massachusetts Avenue (54-1224), Cambridge,
Massachusetts 02139, USA.

E-mail: miskovic@mit.edu

alkaline suite and cluster between 180 and


170 Ma. A volumetrically minor intrusive
pulse of Oligocene age (ca. 30 Ma) is detected
near the southwestern Cordilleran border
with the Altiplano. Both post-Gondwanide
(30170 Ma), and Precambrian plutonism
(6911123 Ma) are restricted to isolated
occurrences spatially comprising less than
15% of the Eastern Cordillera intrusives.
Only one remnant of a Late Ordovician intrusive belt is recognized in the Cuzco batholith
(446.5 9.7 Ma) indicating that the Famatinian arc system previously identified in Peru
along the north-central Eastern Cordillera
and the coastal Arequipa-Antofalla terrane
also existed inboard of this parautochthonous crustal fragment. Hitherto unknown
occurrences of late Mesoproterozoic and
middle Neoproterozoic granitoids from the
south-central cordilleran segment define
magmatic events at 691 13 Ma, 751 8 Ma,
985 14 Ma, and 10711123 23 Ma that
are broadly coeval with the Braziliano and
Grenville-Sunss orogenies, respectively.
Our data suggest the existence of a continuous orogenic belt in excess of 3500 km along
Western Amazonia during the formation of
Rodinia, its early fragmentation prior to
690 Ma, and support a model of reaccretion
of the Paracas-Arequipa-Antofalla terrane to
western Gondwana in the Early Ordovician
with subsequent detachment of the Paracas
segment in form of the Mexican Oaxaquia
microcontinent in Middle Ordovician. A tectonomagmatic model involving slab detachment, followed by underplating of cratonic
margin by asthenospheric mantle is proposed
for the genesis of the volumetrically dominant
Late Paleozoic to early Mesozoic Peruvian
Cordilleran batholiths.

INTRODUCTION
Whereas the Cretaceous to recent Andean
orogenic cycle is well characterized (e.g.,
Ramos and Aleman, 2000), our knowledge of
the early Phanerozoic and Proterozoic evolution of the Andes becomes increasingly fragmentary with age due to paucity of exposed
lithologies. The problem is less pronounced
along the Peruvian segment of the orogen
where a lacuna in the ubiquitous Cenozoic volcanic cover is interpreted to have resulted from
the flat-slab subduction of the Nazca ridge during the Neogene (Jaillard et al., 2000). Batholiths of the Eastern Cordillera of Peru straddle
the tectonic boundary between the Western
Amazonian tectonic provinces of San Ignacio
(1.571.24 Ga) and Sunss (1.190.92 Ga;
Cordani and Sato, 2000) and parautochthonous
to allochthonous crustal domains (1.91.8 Ga
Arequipa-Antofalla; 150 Ma Olmos-Amotape
terrane), thus providing a continuous record of
the nature and rate of crustal growth at a longlived cratonic margin. Despite its fortuitous
setting, however, the timing of magmatism in
the central Andes is relatively poorly understood with most of the geochronological work
to date relying heavily upon whole-rock Rb-Sr
and K-Ar techniques, both of which are known
to yield ambiguous dates due to low retention
temperatures and the possibility of isotopic
disturbance by subsequent thermal episodes
(Dodson, 1973). This is a particularly acute
problem in Peru when we consider ca. 150 Ma
of uninterrupted subduction during the last
Andean orogenic cycle (Benavides, 1999).
We use a combination of in situ U-Pb geochronology, major- and trace-element geochemical characterization of plutonic rocks along the
1400 km of the orogenic strike of the Eastern

GSA Bulletin; September/October 2009; v. 121; no. 9/10; p. 12981324; doi: 10.1130/B26488.1; 12 figures; 1 table; Data Repository item 2009052.

1298

For permission to copy, contact editing@geosociety.org


2009 Geological Society of America

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru

Foreland

Huancabamba
deflection

Ventuari
Rio
Negro

o
60 Maroni

10o

or

i dg
e

Pacific
Ocean

600 km
F

Na
zc
aR

300

Co

20

di

San
Ignacio
I

-fo

rd

ld

ill

be

l
e t
l l Altiplano r a
e
ra

50 W

Central
Amazon

The Peruvian Andes have been historically


divided into six linear geological provinces
striking parallel to the Pacific coast. From west
to east, these are (1) Coastal forearc, (2) Western Cordillera, (3) Puna-Altiplano, (4) Eastern
Cordillera, (5) Sub-Andean fold and thrust belt,
and (6) Foreland basin (Fig. 1; Jaillard et al.,
2000). The Peruvian Eastern Cordillera is a
1400-km-long belt straddling the east-verging
thrust and fold belts of the Maran, Ticlio, and
Maazo complexes to the west and a 120- to
250-km-wide zone of deformed Mesozoic and
Tertiary foreland sedimentary rocks to the east
(Fig. 2). Its basement lithologies are predominantly comprised of the Paleozoic Maran
greenschist facies metasediments, which underwent four successive stages of deformation
(Bard et al., 1974; Mgard, 1978; Zeil, 1983).
There is a general trend of increasing metamorphic grade to the south resulting in a transition
from gray phyllites with subordinate graphitemica schist intercalations in the north to polydeformed paraschists and granulitic gneisses

O-A

s
ru
th
n
ea rn
nd te
-A Eas
rn
ste
We

Morphogeological Units

70

ub

GEOLOGICAL SETTING

80o

Cordillera of Peru to construct a detailed chronologic framework, and identify tectonic regimes
that shaped the central proto-Andean margin of
Amazonia. By relating the secular changes in
magma composition to the tectonomagmatic
cycles of continental assembly and breakup
over the past 1.1 Ga along the western Rodinia
and Gondwanaland, we can holistically test the
current geodynamic scenarios for the evolution
of the Amazonian shield, with particular focus
on the poorly understood breakup of Rodinia
(Cordani et al., 2003; Loewy et al., 2003; Meert
and Torsvik, 2003; Fuck et al., 2008; Li et al.,
2008; Ramos, 2008).
Our data demonstrate the existence of a composite continental margin heavily dominated
by three distinct intrusive pulses related to the
assembly and breakup of Gondwana (middle
Carboniferous to Late Triassic), together with
volumetrically subordinate plutons emplaced
during the initiation of the modern Andean
cycle of subduction in the Early Jurassic. Plutonic remnants belonging to the early Paleozoic
Famatinian, middle Neoproterozoic Braziliano,
and the late Mesoproterozoic Sunss orogens
are located in south-central Peru and indicate
the presence of a periodically reworked latest
Mesoproterozoic crust, 200 km east of the confirmed limit of the Arequipa terrane and 225 km
from the present-day coast, making them the
westernmost exposures of the Amazonia craton
in South America.

Juruena
Sunss

Brazilides

Su
ns
s

So
Francisco
Puna
aulacogen

Arica bend

A-A

Sierras Pampeanas 0.53-0.48 Ga


Olmos-Amotape 0.16-0.15 Ga FMB
Chilenia Terrane 0.46-0.36 Ga
Arequipa-Antofalla 1.8-1.9 Ga
Famatinia Mobile Belt 0.48-0.43 Ga CT

Phanerozoic platform
sediments

SP

Rio
de la
Plata

Atlantic
Ocean

Figure 1. Map of the central South American continent with the generalized major tectonic
provinces and ages of their most recent tectonothermal episodes. Black triangles representing volcanic centers of the Northern and Central Volcanic Zone are superimposed on the six
morphogeological belts of Peru. Modified after Tassinari and Macambira (1999), Cordani
and Sato (2000), and Loewy et al. (2004).

located 90 km east of Hunuco in the central


segment (Cardona et al., 2006). Whereas U-Pb
geochronology on detrital zircons from the
Maran transect between 6S and 10S effectively constrains the depositional age of the
sedimentary protolith between the Early Ordovician and Carboniferous (Chew et al. 2007),
the timing of emplacement of the numerous
batholiths that intrude the metasedimentary
substrate and form a 1400-km-long intrusive
chain between the two Andean oroclines, the
Huancabamba deflection and Arica bend, is
poorly understood.
Regional Basement
Despite lacking direct geochronological
evidence, the Central Andes of Peru have
been inferred to rest upon the westernmost
Amazonian craton, which formed parts of the
long-lived Mesoproterozoic Sunss orogen
(1.20.95 Ga) that resulted from the collision
of Laurentia with Amazonia (Fig. 1; Ramos
et al., 1986; Sadowski and Bettencourt, 1996;
Tosdal, 1996; Jaillard et al., 2000). The parautochthonous granulitic basement of the southwestern Arequipa-Antofalla terrane (Shackleton
et al., 1979) was initially thought to represent
the apex of a Laurentian promontory (together
with Labrador, Greenland, and Scotland) that

was subsequently detached from Laurentia and


incorporated into the Grenville-Sunss orogen
(Dalziel et al., 1994). Its overall allochthonous
character to cratonic South America was reconfirmed based on U-Pb zircon geochronology
and Pb isotope systematics (Loewy et al., 2004),
although the exact relationship to Laurentia
remains unclear. The tip of the Huancabamba
deflection along the northwestern coast of Peru
is dominated by the Amotape continental parautochthonous block and floored by oceanic
crust where the intervening Lancones synclinorium separates the Amotape complex from the
Loja-Olmos massif (Fig. 1; Mourier et al., 1988;
Spikings et al., 2005). A notable exception to the
dominantly Proterozoic basement architecture
of Peru is the isotopically juvenile root of the
Lima segment of the Peruvian Coastal batholith, which does not yield Amazonian basement
signatures (Mukasa and Tilton, 1985; Macfarlane et al., 1990; Petford et al., 1996). Together
with the absence of inherited zircon ages in
the Western Cordillera between Chimbote and
Pisco (Mukasa, 1986), the juvenile isotopic
ratios have been explained as manifestation of a
fundamental change in the age and nature of
basement rock, and the absence of cratonic crust
has been proposed west of the Eastern Andean
Cordillera between 7S and 14S (Beckinsale
et al., 1985; Polliand et al., 2005).

Geological Society of America Bulletin, September/October 2009

1299

AMOT

APE

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.

80 S
O

78

76

74

72

70 W

OL

FORELAND
BALSAS

0
SITABAMBA
R
oM

WE

ara

ER

HUANUCO

ST

ru
Pe

CO
LIMA

12 S

SAN RAMON

TARMA
TICLIO

CARRIZALES

HUANCAYO

Early Jurassic fold / quartz syenites

HUANCAVELICA

Late Triassic Carabaya granitoids


Triassic Mitu Gr volcano-sediments

PARACAS
N

NI
S
BA

Tit
i

ca

Arequipa

id

ge

A-A TERRANE

Tarma
San Ramon

WESTERN CORDILLERA

YA

LIMBANI

TH

ca

BA

COASA

I
OL
TH

az

Lima

COASTAL ACCR. WEDGE

RA

ANO SANTA ROSA

LA

Cuzco

CA

ALTIPL

CO

10 km

PTP PAUCARTAMBO
Transform
CO
RD
. DE

ALLINCAPAC
QUEROBAMBA
(SUCRE)

Paleozoic basement
(Maraon phyllites, paraschists)
Neo/Mesoproterozoic granites

ABANCAY
DEFLECTION

SA

Permo-Triassic granodiorites/
S-type granites-monzogranitoids
Carbo-Permian rhyodacites/
I-type diorites - granitoids
Ordovician- Early Silurian granitoids

BRAZIL

ALTIPLANO

Sira
Anticline

EASTERN CORDILLERA THRUST-FOLD BELT

ca

IA

h
nc

ER

e
Tr

ILL

CR. PASCO

LIV

ile

SIRA

OXA PAMPA

RD

Ch

10 S

400 km

300

200

WESTERN AMAZONIA

ga
lla
ua

Trujillo

100

BO

oH
Ri

PATAZ

B
Ucayali Basin

FORELAND

Figure 2. Geologic map of the Eastern Cordilleran plutonic belt of Peru illustrating the presently known extent of tectonic domains and
locations of the sampled intrusives. The upper Rio Maran of the northern Eastern Cordillera marks the inferred boundary between
Western Amazonia and craton-free zone underlying much of the coastal Western Cordillera (Haeberlin, 2002). The cratonic edge east
of the thickened Altiplano crust is suggested to lie along the Mitu Group (Gr.) basin (Sempere et al., 2002), while the eastern limit of the
Arequipa-Antofalla terrane is currently defined by 206Pb/ 204Pb isotopic contrast from the Neogene volcanic centers sampling the southern
Peruvian crust (Mamani et al., 2005). The E-W fault system subparallel to the Abancay deflection at the latitude of the Paracas Peninsula
(14S) bounds the AA terrane from the north. The NNE-striking tectonic lineament east of the Lancones basin in the NW coastal Peru
demarcates the extent of the Loja-Olmos terranes (Litherland et al., 1994). Geological cross section modified after Mgard (1967).

1300

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru
Neoproterozoic Evolution
Following a poorly defined, ca. 300 Ma,
post-Sunss tectonic lull, the breakup of Rodinia
in the middle Neoproterozoic led to the opening of the Iapetus Ocean by westward drift of
Laurentia leaving the Arequipa craton attached
to Amazonia (Bond et al., 1984; Cawood et al.,
2001). Subsequent westerly shift of the spreading center is interpreted to have resulted in the
generation of intervening oceanic crust between
both the Arequipa-Antofalla and Pampean terranes and mainland Amazonia that was consumed by eastward subduction of proto-Pacific
lithosphere during the Pampean orogeny in the
Middle Cambrian (Rapela et al., 1998). Dacitic
dikes that crosscut layered amphibolites of
the north Chilean segment of the ArequipaAntofalla block were dated at 635 Ma, and are
currently the only local temporal constraints
on the timing of Arequipa detachment (Loewy
et al., 2004). During the extensional phase, a
passive margin developed along the western
proto-Gondwanan margin while rifting propagated northward into present Bolivia in the
late Neoproterozoic times, eventually forming
the epeiric Puncoviscana depocenter for thick
marine sediments behind the parautochthonous
Arequipa-Antofalla craton (Keppie and Bahlburg, 1999). Together with the Maran River
lineament of north-central Peru, the paleo-suture
between the Arequipa block and ancestral Amazonia constitutes a zone of crustal weakness that
was inherited from the Sunss-aged orogens and
was periodically reactivated during the Phanerozoic (e.g., Permo-Triassic extension and deposition of the volcano-sedimentary Mitu Group;
Kontak et al., 1985, 1990; Forsythe et al., 1993;
Sempere et al., 2002).
Phanerozoic Evolution
Beginning in the Early Cambrian, the western margin of Amazonia was again characterized by compressional tectonics due to global
plate reorganization and the final amalgamation
of Gondwana (Cawood and Buchan, 2007). The
collision of the Pampean terrane and reaccretion
of the Arequipa-Antofalla crustal block against
the Ro de la Plata and Amazonian cratons,
respectively, in the Early Cambrian closed the
Puna aulacogen (Ramos, 2008). The resultant
Sierras Pampeanas orogeny of NW Argentina
was accompanied by low-grade metamorphism
and deformation of shallow marine sediments,
followed by emplacement of the post-orogenic,
calc-alkaline Santa Rosa granitoids, and
coeval rhyodacitic pyroclastics (534523 Ma;
Adams and Miller, 2007) as incipient magmatic responses to a collapsing orogen (Rapela

et al., 1998). Although no clear chronometric


evidence of the Pampean orogen presently
exists in Peru, the diachronous rift-drift transition recorded in clastic platform sedimentary
rocks at the latitude of the northern ArequipaAntofalla terrane suggests the end of an extensional regime by the Late Cambrian (Sempere,
1995). Magmatism soon resumed in the Ordovician along the frontal Arequipa terrane giving
rise to the coastal San Nicols batholith (Fig. 2),
which is interpreted to represent the root of a
continental arc (i.e., Faja Eruptiva Occidental) resulting from eastward subduction of the
proto-Pacific oceanic crust beneath western
Gondwana (Ramos, 1988; Mukasa and Henry,
1990). A recent, single grain, U-Pb zircon geochronological survey of the San Nicols intrusives (Loewy et al., 2004) yielded Ordovician
intrusive ages between 468 and 440 Ma that are
typical of the Famatinian orogenic cycle, which
developed along the eastern Sierras Pampeanas
(Pankhurst and Rapela, 1998). The apparent
eastward transposition of the Ordovician igneous belt north of the Arequipa-Antofalla terrane
in Peru is suggested by the presence of Latest
Ordovician granodioritic gneisses within the
Ro Maran valley east of Cajamarca (Fig. 2).
These are interpreted to mark the presence of
either an original embayment along the western
Gondwanan margin (Chew et al., 2007), or a
gap left by removal of Oaxaquia microcontinent
leaving behind a speculative Paracas terrane,
which collided with western Gondwana during
the Famatinian cycle (Ramos, 2008).
A conspicuous lack of magmatic activity
during much of Late Silurian and Devonian
in the central Andes, as well as the absence of
Grenvillian terranes along the margin from the
Arequipa massif to Chibcha terrane in Colombia has been interpreted as evidence for the
development of a passive margin west of the
Famatinian arc resulting from strike-slip detachment of the northern segment of the ArequipaAntofalla (i.e., Oaxaquia) block during either
the Devonian (Mgard, 1973; Bahlburg and
Herv, 1997) or the Permo-Triassic extension
(Haeberlin et al., 2004). Resumption of arc
activity in the Early Mississippian was first
recorded by Mgard et al. (1971) and Dalmayrac
et al. (1980), who described arc magmatism in
the context of an Eohercynian orogeny, but the
arc activity was only recently documented with
certainty in the Balsas-Callangate-Pataz auriferous province of the northern Eastern Cordillera,
based on 40Ar/39Ar geochronology of plutonic
host rocks and ore deposits (Schreiber, 1989;
Snchez, 1995; Haeberlin et al., 2004; Mikovi
et al., 2005). These intrusives were emplaced
subparallel to the Rio Maran crustal lineament (Fig. 2), and appear to have undergone a

phase of regional uplift in the Early Pennsylvanian, associated with the regional orogenic
type Au-Ag mineralization (Haeberlin, 2002).
Late Permian to Early Jurassic lithospheric
thinning in Peru and Bolivia resulted in emplacement of the central San RamonLa Merced-type
monzogranitoids, associated with localized migmatization (Soler, 1991). The plutonism predated
or was contemporaneous with synrift deposition
of calc-alkaline, Mitu Group bimodal volcanics
and continent-derived sediments in transcurrent
half grabens and pull-apart basins that opened
along the western margin of the orogen (Figs. 1
and 2; Sempere et al., 2002). The genetically
similar but younger pulse of Late Triassic,
syntectonic, peraluminous granitic plutonism
occurred in a localized transpressional setting
in the southern Cordillera de Carabaya, and was
broadly coeval with mantle-derived Mitu Group
alkali basalts and shoshonites of the Cuzco basin
(Kontak et al., 1985, 1990). With the onset of the
modern Andean tectonic regime, and renewed
easterly subduction of the protoPacific plate
below the western South American margin in the
Early Jurassic, the newly formed continental arc
developed along the present-day Coastal (Western) Cordillera, thus leaving the Eastern Cordillera in a backarc position, which consequently
affected both the volume and type of plutonism
during the past 150 Ma. Except for the easternmost porphyritic monzonite and granitoid
stocks of the Miocene metallogenic belt along
the central and northern Western Cordillera of
Peru (Noble and McKee, 1999), the inferred
Early Jurassic age for the peralkaline Allincapac
volcano-plutonic complex in the Cordillera de
Carabaya of southern Peru (Fig. 2; Kontak et al.,
1990) makes it the only volumetrically significant intrusive locality of the Andean cycle in the
Peruvian Eastern Cordillera. In summary, alternating passive margin sedimentation, volcanic
arcs and rift-related magmatism, including tectonothermal events associated with both strikeslip displacements and crustal shortening over
the past 1.0 Ga, has led to a composite magmatic
belt displaying complex lateral and along-strike
variations throughout the Peruvian cordilleras.
Previous Geochronology
A compilation of radiometric ages from
the Eastern Andes of Peru (Jacay et al., 1999)
reveals that the majority of the reported dates
are based either on whole-rock Rb-Sr and K-Ar
isotopic analyses of biotite and K-feldspar, or
U-Pb chronometry of bulk-zircon separates with
relatively few, high-precision, single-grain U-Pb
zircon or 40Ar/39Ar dates. Moreover, previous
work was mainly of local focus, and no attempts
were made at a regionally integrated survey. A

Geological Society of America Bulletin, September/October 2009

1301

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.
review of the available data indicates three main
pulses of intrusive activity, separated in time and
space; from north to south these are Carboniferous, Permian to Triassic, and Late Triassic belts
(Petersen, 1999). Intrusive activity in the northern
segment of the Eastern Cordillera was initially
dated by Snchez (1983, 1995), who reported
K-Ar dates of 346.7 7.3 Ma and 329 10 Ma
for the Balsas and Callangate monzogranitic plutons. Recent 40Ar/39Ar dating of the auriferous
Pataz Batholith granodiorites by Haeberlin et al.
(2004) confirmed the earlier 40Ar/39Ar geochronology of Schreiber et al. (1990), with plateau
ages ranging between 329.2 1.4 Ma (biotite)
and 319 3.2 Ma (hornblende). Emplacement of
the east-central leucogranitoids was crudely constrained by bulk-zircon U-Pb dating of Lancelot
et al. (1978) and Dalmayrac et al. (1980), and
K-Ar work on biotites by Soler (1991) from the
Equiscocha (253 11 Ma), San Ramon (246
10 Ma), La Merced (238 10 Ma), and TalhuisCarrizal plutons (245 11 Ma and 233 10 Ma).
The youngest pre-Andean intrusives in the Peruvian Eastern Cordillera were identified in the
southern segment of the Cordillera de Carabaya,
where S-type monzogranitoids yielded K-Ar
(biotite and muscovite) and whole-rock Rb-Sr
ages between 225 15 Ma in the Coasa Batholith at 14S, and the southernmost Limacpampa
pluton was dated at 199 10 Ma (15S; Clark
et al., 1990; Kontak et al., 1990).

tite, hornblende granodiorites, and minor granites. Fragmented diorite dikes and mafic magma
pulses intrude most of the granodioritic rocks
and occur together with abundant microgranular
enclaves as well as partially fused xenoliths of
the Maran Complex. There is widespread textural evidence supporting the coexistence and
mingling of compositionally contrasting magmas (Fig. 3A). Similar intrusive facies prevail
across the belt at 8S with the upper Huallaga
River granitoids exhibiting variable proportions

of amphibole and biotite. A feature unique to the


northeastern margin of the Eastern Cordilleran
intrusive belt is the ubiquitous presence of NWtrending bimodal dikes of the Permo-Triassic
Mitu Group.
Central Eastern Cordillera
The central intrusive belt, including the
northern Altiplano, was sampled along four
transects between the latitudes of 10S and

SAMPLING
Five orthogonal and three margin-parallel
sampling transects were conducted throughout the Eastern Cordillera of Peru between the
latitudes of 6S and 15S. Traverses varied in
length between 70 and 250 km and spanned the
total orogenic strike length of 1300 km (Fig. 2).
Northeastern Cordillera
The northern sector was surveyed along two
traverses covering: (1) the Balsas-Callangate
plutons (6S to 7S), and (2) the intrusive
belt between the Pataz Batholith, bounded
by the Maran River to the west and the Ro
Huallaga to the east, between 7S and 9S. The
Balsas, Callangate, and Pataz intrusive complexes are emplaced subparallel to a NNWtrending fault zone associated with the upper
Maran River valley and locally intrude the
Ordovician Maran biotite schists and phyllites,
Contaya Formation meta-arenites and Ordovician to Devonian Vijus metabasalts. The plutonic
rocks define a compositional spectrum from
medium- to coarse-grained, hornblende-bearing
diorites, through amphibole-rich tonalites into
volumetrically dominant, medium-grained bio-

1302

Figure 3. Field photographs illustrating the typical mineralogical and structural features of
the Eastern Peruvian intrusives; (A) mingling of microdioritic and granodioritic magmas
in the calc-alkaline Pataz Batholith (8S), (B) mineralogically monotonous monzogranite of
the San Ramon pluton (11S) cut by up to 1-m-thick aplite dikes, (C) a partially absorbed
metasedimentary xenolith (restite) in the area of high-grade migmatization along the Chanchumayo River segment of the San RamonLa Merced batholith, (D) medium-grained
Machu Picchu biotite granite (13S), (E) a close-up photograph of the Allincapac complex
nepheline syenite (13.5S) with arfvedsonite and biotite as the principal Fe-Mg phases,
(F) Coasa Batholith K-feldspar megacrystic leucogranites (14S) displaying the classic
horse tooth texture.

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru
14S. Progressively from north to south,
these are (1) OxapampaCerro de Pasco,
(2) TarmaLa Merced, (3) Satipo-Huancayo,
and (4) Huancavelica-Sucre (Fig. 2). Despite
the textural heterogeneity that is locally manifested by widespread migmatization along the
Rio Chanchumayo, or the existence of a WNWoriented fabric found in the Talhuis pluton as
well as in the Querobamba intrusive suite of
the eastern Sucre region, the intrusive facies
of the central Cordilleran segment exhibit overall uniform silica enrichment compared to the
northern Peruvian plutonic belt. The NNWoriented, San RamonLa Merced batholith is
the type locality of Permo-Triassic plutonism in
east-central Peru (Mgard, 1978; Soler, 1991). It
comprises an 80-km-long and in places 10-kmwide intrusive belt composed of characteristically pink-colored, K-feldspar megacrystic,
biotite, and subordinate amphibole-bearing
granodiorites to leucogranites (Fig. 3B) that are
emplaced into late Carboniferous metasediments
of the Tarma-Copacabana Group, which are
locally strongly deformed. The interior granitoid
facies of each plutonic body include mica-rich
restites, together with subordinate mafic microgranular enclaves, whereas the volumetrically
minor biotite quartz monzonites occur along the
margins of individual plutonic bodies and entrain
ubiquitous, partially fused, metasedimentary
xenoliths (Fig. 3C). The NNE-striking, subhorizontal aplite and fine-grained monzonite dikes
indiscriminately traverse the intrusive suites but
are conspicuously absent from those that display strong structural fabrics. The mineralogical trends across the orogenic strike, from east
to west, are characterized by an increase in the
modal amount of amphibole accompanied by
a decrease in K-feldspar and a larger proportion of microdioritic enclaves. In contrast to the
northern segment, the central intrusives of the
Eastern Cordillera are bound inboard and outboard by contiguous, transtensional grabens
that accumulated interbedded Mitu Group
bimodal volcanics (rhyodacitic pyroclastics and
minor pillow basalts), together with subarkosic
arenites during inferred Triassic synorogenic,
continental rifting (Sempere et al., 2002).

tion (Fig. 2), corresponds with the appearance


of the thickened Altiplano crust to the west that
is interpreted as a collage of lithospheric structural blocks that are stacked vertically along
E-Wdipping, doubly vergent, Miocene thrusts
(Allmendinger et al., 1997). We have surveyed
the discontinuously outcropping intrusive facies
along a strike-parallel traverse over 8300 km2
starting from the northern Cuzco batholith, in
between the two principal transform lineaments.
The traverse passes through the western margin
of a poorly delineated intrusive belt near Paucartambo, via the central San Gabn pluton, and
into the southernmost segment represented by
the Cordillera de Carabaya specifically, the western Ayapata suite of the Coasa pluton (Fig. 2;
Kontak et al., 1990). The Cusco and Paucartambo batholiths are composed of texturally
homogeneous, medium-grained, biotite-bearing,
alkali feldspar granites (Fig. 3D) intruded into
the Paleozoic basement phyllites that are equivalent to the Maran basement in the northeastern
Peruvian Cordillera. There is a gradual increase
in the proportion of modal mica as well as the
first appearance of muscovite in the granitoid
assemblages southward of Paucartambo. The
centrally located San Gabn biotite granite displays numerous, 3- to 5-cm-long mafic enclaves
and mica-rich restites. Of particular importance
is the occurrence of peralkaline intrusives such
as the Nevado de Allincapac volcano-plutonic
complex covering 350 km2. It is composed of
porphyritic leucite, nepheline phonolites, and
peralkaline volcanic breccias that are intruded
by the coarse-grained, nepheline-bearing, alkali
feldspar syenite core (Fig. 3E). The central
and southern Carabaya intrusive suite is predominantly composed of medium-grained to
K-feldspar megacrystic, two-mica monzogranites (Fig. 3F), granodiorites, and muscovite
leucogranites, occasionally containing cordierite
and locally displaying intermediate to mafic
facies (diorite to quartz diorite). The Carabaya
granitoids exhibit a sporadic magmatic fabric and are cut by sinuous and discontinuous
biotite quartz diorite dikes that are themselves
traversed by the host granitoids indicating synintrusive emplacement (Pitcher, 1997).

Southeastern Cordillera

RESULTS

An overall sinistral offset along the


ENE-WSWtrending Patacancha-TamburcoPuyentimari transform fault system during
Permo-Triassic rifting is considered responsible for a drastic change in strike of the Peruvian Eastern Cordillera (Carlotto et al., 2006).
The resultant displacement of the intrusive belt
and the Mitu graben by 200 km eastward at the
latitude of 13S, known as the Abancay deflec-

Zircon Characterization
A total of 738 grains extracted from 60
intrusives were imaged and examined for
morphology and internal textures. Representative subsets were analyzed for trace-element
content and were dated by the U-Pb method.
Ages are shown in Table 1. The results of
individual U-Pb spot analyses are presented

in the GSA Data Repository (Table DR11).


The analyzed crystals are colorless and transparent ranging in size from 50 to 250 m and
with ratios of length to width between 1:1 and
3:1. In cathodoluminescence (CL) images,
95% of zircons exhibit well-developed, yet
frequently blurred or convoluted oscillatory
growth zoning, characteristic of a magmatic
origin (photomicrograph insets in Figs. 6M,
6S, and 6X; Corfu et al., 2003). The zoned
rims tend to discordantly overgrow CL-bright,
oscillatory-zoned cores. In 25% of the zircon
population, thin rims mantle patched or sectorzoned xenocrystic cores that are characterized
by moderate CL intensities. Recrystallization
and abundant mineral inclusions or melt trails
are observed in less than 8% of the imaged
grains. All identified domains within texturally
complex zircons such as cores, rims, or sectors
were analyzed by separate laser traverses.
U-Pb Geochronology
The U-Pb ages from the intrusive rocks
of the Eastern Peruvian Cordillera fall into
more than six distinct groups, with the geographical and temporal distribution presented
in Figures 4 and 5. Plots based on 476 concordant analyses are presented in Figure 5.
The identified age populations are Oligocene
(2834 Ma; n(grains) = 8), Early to Middle Jurassic (167174 Ma; n(grains) = 12), Late Triassic
to Early Jurassic (178217; n(grains) = 48);
Permian-Triassic (ca. 220270 Ma; n(grains) =
126), Carboniferous-Permian (ca. 275360
Ma; n(grains) = 207); Ordovician (ca. 440
530 Ma; n(grains) = 9); middle Neoproterozoic
Cryogenian (ca. 650770 Ma; n(grains) = 24),
and a broadly defined Late Mesoproterozoic
early Neoproterozoic (Stenian-Tonian) age
span between 960 and 1200 Ma (n(grains) = 25).
An additional 17 zircons, with ages clustering
around 1300 Ma but also extending to 1700 Ma,
do not correspond to pluton crystallization
ages, but are instead interpreted as an inherited
Mesoproterozoic component (Fig. 5A).
Middle JurassicOligocene
The middle Cenozoic (Andean) ages clustering near 30 Ma were obtained from two
granodiorites (SCAM-04 and 08) in the southcentral Huancavelica region of the Eastern
Cordillera (Fig. 6A). The Oligocene zircons
are long prisms displaying blurred primary
1
GSA Data Repository item 2009052, In situ
zircon U-Pb isotope data and whole-rock geochemistry of the Peruvian Eastern Cordilleran intrusives, is available at http://www.geosociety.org/pubs/
ft2008.htm or by request to editing@geosociety.org.

Geological Society of America Bulletin, September/October 2009

1303

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.

Sample
SCAM-04
SCAM-08
NAM-11a
SAM-21
SAM-20

Locality
ColcabambaHuancavelica
ColcabambaHuancavelica
Huayillas-Pataz
Chillacori-Puno
Chillacori-Puno

TABLE 1. SUMMARY OF THE LA- ICP-MS IN SITU ZIRCON U-PB GEOCHRONOLOGY


OF THE EASTERN PERUVIAN CORDILLERA PLUTONS (CUMULATIVE CONCORDIA AGES)
Latitude
Lithology
n (analyses)
Concordia age
Inheritance
(2)
(2)
(Ma)
(Ma)
S
12.5
Granodiorite
19
29.37
0.78

Common Pb

Pb loss

Moderate

Minor

12.5

Granodiorite

15

31.36

0.64

94

Minor

8.2
13.8
13.8

19
11
2

172.9
184.1
195

2.0
3.7
11

Minor

11

190.3

2.5

Minor

Ayapata-Carabaya

13.8

Quartz syenite
Nepheline syenite
Nepheline
monzosyenite
Granite

San Gabn
E. CoasaCarabaya

13.7
13.8

Granite
Monzodiorite

12
7

191.2
207.0

3.5
3.4

281

20

Minor

E. CoasaCarabaya

14.0

Monzodiorite

13

208.4

4.9

294

25

Minor

S. San Gabn

13.8

12

216.1

3.1

274

13

Minor

San Ramn
San Ramn
La Merced

11.3
11.3
11.2

11
17
8

223
265.5
250.0

12
7.3
5.9

279

268

21

12

Moderate
Strong

CAM-16
CAM-24

Satipo
Sacsacancha

11.0
11.5

13
10

256.3
260

4.8
18

309

Strong
Minor

CAM-33
CAM-35
CAM-39

Oxapampa
Oxapampa
Rio Huachn

10.7
10.8
10.7

13
17
9

238.0
246.8
227.4

3.6
4.6
3.8

351

25

Minor
Minor

Minor

CAM-45a
NAM-05
SCAM-01
SCAM-02
SCAM-22
CAM-19a
CAM-40
COCA
268
CAM-45c
SAM-08
CAM-10

Paucartambo-Pasco
SE Pataz
Ayacucho
Ayacucho
Ayacucho
Mariposa-Junin
Rio Huachn
C. CoasaCarabaya

10.7
8.1
13.3
13.0
13.4
11.4
10.7
14.0

Alkali feldspar
granite
Granite
Granite
Alkali feldspar
granite
Granite
Alkali feldspar
granite
Granite
Granite
Alkali feldspar
granite
Granite
Granite
Quartz monzonite
Granite
Granite
Granodiorite
Granodiorite
Granodiorite

13
7
6
21
7
19
18
5

258.4
244.5
243.4
217.8
238.4
254.5
260.9
227.7

4.4
3.3
6.7
3.5
4.9
4.2
4.2
5.6

288

289
291

7
35

Strong
Moderate
Moderate
Minor

Minor

Minor

Paucartambo-Pasco
Urubamba
San Ramn

10.7
13.3
11.3

CAM-41
COCA
362
COCA
362
COCA
358
SAM-17
SCAM-06

Rio Huachn
Limbani-Carabaya

COCA
302
SAM-22a
COCA
262
COCA
269
COCA
298
CAM-11a
CAM-12
CAM-15

CAM-03
CAM-26
CAM-30
CAM-49a
CAM-52
CAM-54
CAM-54
CAM-55b
CAM-57
NAM-02a
SAM-09
SAM-12a
CAM-04
CAM-04
AM-80
NAM-28a
CAM-44a
NAM-27a
NAM-30
NAM-18
NAM-22

1304

14
8
2

253.6
284.8
259.7

6.3
4.6
8.1

296

14

Minor

Strong

10.7
14.2

Tonalite
Quartz syenite
Alkali feldspar
granite
Granite
Granite

18
10

255.6
227.4

7.4
5.4

402
278

11
28

Minor

Limbani-Carabaya

14.2

Granite

468

23

Aricoma-Carabaya

14.1

Granite

227.4

4.2

245

12

Minor

Santa RosaPuno
ColcabambaHuancavelica
Parcamayo

13.3
12.4

Rhyolitic tuff
Granite

8
8

226
257.6

10
8.6

Strong

Minor

11.3

18

315.2

4.3

Moderate

Minor

Huancayo
Parihuanca
Huachn
Tingo Maria
Upper Rio Huallaga
valley
Upper Rio Huallaga
valley
Nuevo Progresso
Nuevo Progresso
SE Pataz
Urubamba

11.8
12.0
10.7
9.2
8.7

Alkali feldspar
granite
Granite
Granite
Granite
Granite
Granite

5
15
15
17
17

292
284
309.4
313.4
293.3

20
15
4.0
5.2
5.0

343
349
363
513

24
7
24
17

Strong
Moderate
Moderate
Moderate

Minor

8.7

Granite

934

26

8.6
8.5
8.1
13.3

16
10
12
5

316.7
304.5
301
291.5

5.9
7.2
5.2
5.8

497
351

23
8

Strong
Moderate

Machu Picchu
Parcamayo
Parcamayo
Central Pataz
San Vincente
Amazonas
Junin
Balsas
Golln-Callangate
West Balsas
East Balsas

13.2
11.3
11.3
7.8
7.0

Granite
Granite
Quartz monzonite
Alkali feldspar
granite
Granite
Granite
Granite
Monzogabbro
Tonalite

14
19

5
16

324.1
317.4

333.2
313.5

5.3
4.4

7.7
4.5

376
687

7
27

Minor

Moderate

10.8
7.0
7.2
6.9
6.8

Granodiorite
Granodiorite
Granodiorite
Granodiorite
Granodiorite

9
17
15
17
17

303.8
313.9
313
320.0
309.0

5.3
4.3
4.3
3.8
4.0

354

1303

35

Minor

Minor

(Continued )

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru
TABLE 1. SUMMARY OF THE LA- ICP-MS IN SITU ZIRCON U-PB GEOCHRONOLOGY
OF THE EASTERN PERUVIAN CORDILLERA PLUTONS (CUMULATIVE CONCORDIA AGES) (CONT.)
Sample
Locality
Latitude
Lithology
n (analyses)
Concordia age
Inheritance
Pb loss
(2)
(2) Common Pb
(Ma)
(Ma)
S
SAM-04a East Cusco batholith
13.2
Alkali feldspar
10
446.5
9.7
508
12

granite
SAM-04a East Cusco batholith
13.2
Alkali feldspar

1110
26

Minor
granite
CAM-22
Carrizales
11.5
Quartz monzonite
16
752
21
1011
34
Moderate
Minor
CAM-23
Carrizales
11.5
Alkali feldspar
7
691
13

Minor

granite
SCAM-18 Querobamba-Sucre
13.9
Alkali feldspar
13
751.7
8.1

Minor
Minor
granite
CAM-17
Satipo
11.3
Tonalite
18
985
14
1149
17

CAM-17
Satipo
11.3
Tonalite

1226
18

CAM-18
Mariposa-Junin
11.4
Alkali feldspar
17
1071
23
1305
33

Minor
granite
CAM-18
Mariposa-Junin
11.4
Alkali feldspar

1668
28

Moderate
granite
SCAM-17 Querobamba-Sucre
14.0
Granite
8
1123
13
1200
38

Note: The concordia ages are calculated using the routines of Ludwig (2003) and following Th and U decay constants of Steiger and Jger (1977). See Figure 2 for
sample locations. LA-ICP-MSLaser ablationinductively coupled plasmamass spectrometry.

zoning developed around rare planar cores.


Only one inherited zircon was identified with
a U-Pb concordia age of 94 Ma. The Middle
Jurassic grains were extracted from two geographically dispersed localities at the southern
and northern extremes of the study area (Figs.
6B and 6C). The concordia age recorded by a
quartz syenite sample collected at the margin
of the northern Cordilleran Pataz batholith
(Huayillas; NAM-11a), and two SiO2 undersaturated nepheline (monzo)syenitic intrusives
of the southern Allincapac volcano-plutonic
complex in the southernmost Cordillera de
Carabaya (SAM-20, 21) yielded Middle to
Early Jurassic ages between 172.9 2.0 and
184195 Ma, respectively. Whereas the zircons from the north exhibit well-preserved
oscillatory growth zones within large and
euhedral crystals, their southern equivalents
from the peralkaline Allincapac suite take the
form of irregularly truncated crystal fragments
of larger zircon grains. Not surprisingly, both
of the two zircon populations are devoid of
xenocrystic cores, which tend to be dissolved
in Zr undersaturated magmas such as the host
syenites (Watson, 1996; Baker et al., 2002).
Late Triassic
The Late Triassic ages were obtained from
peraluminous, highSiO2 (monzo)granites of
the central and northern Cordillera de Carabaya,
namely the Coasa, Ayapata, and San Gabn
plutons (Fig. 4). Concordia ages from samples
COCA 302, 262, 269, 298, and SAM-22a show
both minor Pb loss and the presence of elevated
levels of common Pb. The ages obtained straddle
the Triassic-Jurassic boundary from 190.3 2.5
to 216.1 3.1 Ma (Figs. 6D6F). A unique feature of this zircon group is widespread evidence
for post-magmatic, solid-state recrystallization
represented by ubiquitous blurring and occa-

sionally convoluted zoning within otherwise


euhedral prismatic crystals. The CL-dark rims
occasionally have sporadic bright domains that
are subparallel to the growth zoning and are
likely due to a combination of metamictization
by U-induced radiation and localized recrystallization (Kempe et al., 2000; Nasdala et al.,
2002). The inherited component from moderately to strongly recrystallized, CL-bright,
xenocrystic cores defines an early Permian age
interval between 274 and 281 Ma.
Permo-Triassic
The emplacement of voluminous, partially
migmatized granitoids within the south-central
segment of the Eastern Cordillera of Peru is constrained to the Permian and Triassic (223 12 Ma
to 284.8 4.6 Ma). Although encompassing
more than 65 Ma and stretching over 800 km of
orogenic strike, from the central Peruvian district of Hunuco to the southernmost Cordillera
de Carabaya (Aricoma pluton) (Fig. 2), 80%
of the Permo-Triassic plutonism by area was
emplaced within a geographically restricted belt
between 10S and 12S with a peak in magmatic
activity occurring from 240 to 260 Ma (Figs.
6G6L). Zircon grain morphology is dominated
by euhedral and fragmented prisms that have
preserved primary, oscillatory growth zones.
Exceptions to the dominant magmatic growth
pattern are zircons from three granitoids of the
southern Cordillera de Carabaya (COCA 268,
358, and 362). Here, as in the Late Triassic magmatic episode, uniformly low CL rims surround
partially recrystallized cores and show widespread convoluted zoning. Overall, in situ U-Pb
isotopic analyses reveal the minor to moderate
presence of common Pb. Half of the analyzed
grains display CL-bright xenocrystic cores, with
the dominant peak in the inherited age spectrum
clustering between 288 and 296 Ma (Table 1).

Late OrdovicianEarly Permian


The Carboniferous to early Permian ages are
recorded by the compositionally heterogeneous
and regionally most extensive intrusive belt
of the Eastern Cordillera of Peru. This array
of plutons dominates the orogenic margin
in its northern segment between the cities of
Bolvar and Hunuco (6S to 10S) but also
extends farther southward, west of the PermoTriassic domain, as a narrow suite of plutons
and stocks that ultimately terminates with
the outer Cuzco batholith at 13.5S (Figs. 2
and 3). A 5 Ma lacuna at the MississippianPennsylvanian boundary separates two dominant magmatic pulses. The shorter, early Carboniferous intrusive episode progressively
increases in number of concordant analyses
from 350 to 325 Ma before a sharp decrease,
reemergence at 315 Ma, and final termination
at 285 Ma (Figs. 6M6R). Furthermore, the
diachronous nature of Carboniferous-Permian
plutonism is confirmed by the progressively
southward-younging trend from an average
intrusive age of 314.5 Ma in the Pataz region
to 299.9 Ma near Cuzco (Fig. 4). The pattern
of decreasing crystallization ages from north
to south roughly corresponds to a similar trend
in the ages of inherited zircons. Namely, the
Balsas granodiorite (NAM-22) at the latitude
of 7S retains a Mesoproterozoic age as old as
1303 35 Ma, while CAM-54 and CAM-04
granites of the central Ro Huallaga valley
(9S) and Parcamayo district (11S), respectively, record Neoproterozoic dates of 934 26
and 687 27 Ma. Subhedral to euhedral, occasionally fragmented zircon prisms show mostly
pristine to faintly blurred primary zoning around
CL-bright cores. An age of 293.3 5.0 Ma
from the Huallaga granite (CAM-54) was
obtained on zircons that have clearly undergone a phase of metamorphic overprint, as

Geological Society of America Bulletin, September/October 2009

1305

AMOT

APE

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.

80 S
O

78

OL320.0

N-18
N-27a

N-22

333.2 7.7
o

74

309.0 4.0
3.8

313.9 4.3

76

R
ar o
a
n

AM-4-80

N- 28a
N -30

313.5 4.5
313.0 4.3
244.5 3.3

N-5

304.5 7.2

Trujillo
N-2a
301.0 5.2 N-11a C55b

C-57
C-54

72

70 W

WESTERN
AMAZONIAN
TECTONIC
PROVINCES

293.3 5.0

Hu R
al io
la
ga

ru
Pe

172.9 2.0
316.7 5.9

313.4 5.2

100

200

300

400 km

C-52

ile

Ch

258.4 4.4,253.6 6.3,227.4 3.8


309.4 4.0
C -45a,c 238.0 3.6
10 S
BRAZIL
C -39
246.8 4.6
260.9 4.2 C-49a
C-33
256.3
4.8
C-35 255.6 7.4
C-40
303.8 5.3
C-16
C41
250.0 5.9
C-44a
C15 C-17,18
317.4 4.4
C12 C19a
C-4,3
985 14,1071 23
315.2 4.3
C-10,11
C
C-22,23
254.5 4.2
259.7 8.1,223 12 265.5 7.3 C-26 C24
o
752 21,691 13 291.5 5.8
12 S
C-30
260
18 324.1 5.3
292 20 SC-6
LIMA
284.8 4.6
SC-4
284 15
SC-8
S-12a
SC-2
191.2 3.5 190.3 2.5
S-9
257.6 8.6 SC-1
Early Jurassic fold / quartz syenites
S-8
S-22a
217.8 3.5
S-4a
COCA 302 208.4 4.9
SC-22
Cuzco
243.4 6.7
COCA 298
Late Triassic Carabaya granitoids
COCA269
S-20,21
446.5 9.7
238.4 4.9 SC-18
COCA 262
COCA 268
Permo-Triassic granodiorites/
SC-17
COCA 358
207.0 3.4
216.1 3.1
S-type granites-monzogranitoids
751.7 8.1
COCA 362
199.1 7.4
S-17
Carbo-Permian rhyodacites/
1123 13 184.1 3.7
235.6 6.1
I-type diorites - granitoids
226 10 T 227.4 5.4
iti
Ordovician- Early Silurian granitoids
N
227.7 5.6
ca
az
ca
ca
Paleozoic basement
R
Arequipa
(Maraon phyllites, paraschists)
id
ge
A-A TERRANE
Neo/Mesoproterozoic granites
o

ch
en
Tr

SA

LA

CO

NI

BO

LIV

TH
LI

IA

HO

T
BA

Figure 4. Regional distribution of the intrusive ages along the Peruvian Eastern Cordillera with the errors presented as 2. Sample
name abbreviations: NNAM; CCAM; SCSCAM. Two Oligocene ages for SCAM-04 and SCAM-08 of 29.37 Ma and 31.36 Ma,
respectively, are not shown.

indicated by the presence of thin blurred rims


mantling large, CL-bright xenocrystic cores.
A single, Late Ordovician concordia age of
446.5 9.7 Ma is recorded by a marginal alkali
feldspar granite (SAM-04a) of the eastern Cuzco
batholith. Isotopic analyses of CL-dark cores
revealed a bimodal inheritance at 508 12 Ma
and 1110 26 Ma (Fig. 6S). All of the Cambrian ages identified in the Eastern Cordillera
are inherited zircons sampled by Pennsylvanian or Ordovician intrusions.

1306

Neoproterozoic
Aside from the parautochthonous ArequipaAntofalla basement rocks (Loewy et al., 2004),
concordant Precambrian ages have been recovered along the central segment of the protoAndean margin of Amazonia from two disparate
localities in the south-central Eastern Cordillera
of Peru. The emplacement of a quartz monzonite
(CAM-22) and alkali feldspar granites (CAM-23)
in the Carrizales and Querobamba (SCAM-18)
plutons has been constrained to 752 21,

691 13, and 751.7 8.1 Ma, respectively


(Figs. 6T and 6U). Zircons exhibit minor to moderate presence of common lead and appear to
have experienced only minor Pb loss. Morphologically, the Neoproterozoic zircons from the
Carrizales region comprise crystalline fragments
and structureless xenocrystic cores where the primary oscillatory growth zones have been largely
obliterated. The Querobamba sample, however,
preserves the original, albeit somewhat blurred,
pattern of growth within euhedral prisms.

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru
45
40

Gondwanides

35

n = 476
278
Bin width: 20 Ma
Mean efficiency: 68.7 %

Frequency

30
25
20

Serra da
Nova Brasilndia /
Providncia /
Sunss belts
E. Braziliano /
PampeanLomas
(1.07 - 1.19 Ga)
Pan-African
Famatinian
Aguape belt
San Igncio Meneches
orogeny
orogeny
suite
(0.96-92 Ga)
orogeny
0.65 0.77 Ga)
(0.63-0.75
(0.44-0.53 Ga)
(Sunss orogeny) (1.32 -1.34 Ga) (1.52-1.57 /
1.66-1.69 Ga)

Andean
cycle

15
10
5
0
0

100 200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500 1600
206

Pb/ 238U age (Ma)

16

B
n = 393
217
Bin width: 5 Ma
Mean efficiency: 33.5%

14
12

Frequency

Carbo-Permian Gondwanides
(Pangea assembly)

Permo-Triassic Gondwanides
(Pangea break-up)

10

Late Triassic
(Cordillera de Carabaya)

Early Jurassic

6 (Andean back arc)


4
2
0
150

175

200

225

250
206

Pb/

275

300

325

350

238

U age (Ma)

Figure 5. Zircon 206Pb/ 238U age histograms summarizing the known intrusive episodes represented by the granitoids of the Eastern Cordillera of Peru; (A) cumulative
Mesoproterozoic-Oligocene diagram, (B) blow-up of the volumetrically most dominant plutonism of Gondwanide age (350160 Ma). Empty bins represent crystallization ages; black
boxes correspond to the ages of inherited zircons.

Late Mesoproterozoic
Closely associated with the Neoproterozoic
granitoids are partially foliated and compositionally more diverse late Mesoproterozoic
to earliest Neoproterozoic intrusives. These
include: (1) the Satipo tonalite (CAM-17),
which yields an age of 985 14 Ma, with a
strong inherited component between 1150 and
1250 Ma, (2) the Mariposa alkali feldspar granite (CAM-18) with a Mesoproterozoic age of
1071 23 Ma and exhibiting bimodal inheritance recorded by xenocrystic cores at 1305 Ma
and 1668 Ma, and (3) the Querobamba gran-

ite (SCAM-17), emplaced at 1123 13 Ma,


with loosely defined inheritance clustering at
1200 Ma (Figs. 6V6X). In cathodoluminescence images, the Mesoproterozoic grains display ubiquitous dark rims and moderately to
strongly recrystallized, CL-bright cores.
Whole-rock Geochemistry and
Tectonic Affinity
Whole-rock, major-, minor-, and traceelement data are presented in Table DR2 (see
footnote 1). The Early Jurassic Allincapac

nepheline monzosyenites are the only sampled


intrusives from the Eastern Peruvian Andes
that are silica undersaturated and lack normative quartz. In the standard International Union
of Geological Sciences (IUGS) classification of
Streckeisen (1974), the majority of the Carboniferous-Permian and half of the Late Triassic Cordillera de Carabaya plutonic rocks have modal
mineralogies corresponding to granites, while
the Permo-Triassic plutons together with three
Neoproterozoic samples straddle the boundary
between the granite and alkali-feldspar granite
fields (Fig. 7). Both of the Oligocene (Andean)
samples are quartz syenites, while the Neoproterozoic intrusives plot within the alkali feldspar granite field. The most diversified intrusive
suite is comprised of the Ordovician samples,
which span the modal spectrum from quartz
monzonites through granodiorites to alkali feldspar granites. According to a recently proposed
geochemical classification scheme for granitic
rocks based on major-element chemistry (Frost
et al., 2001), the Carboniferous-Permian, Ordovician, and the Late Triassic granitoids are all
magnesian, but transgress different boundaries
between the calcic, calc-alkaline, and alkalicalcic differentiation trends, respectively. They
span a wider compositional spectrum (60
77 wt% SiO2) than the predominantly alkalicalcic, magnesian to ferroan Neoproterozoic
and Permo-Triassic plutons (7078 wt% SiO2;
Fig. 8A). The calcic Oligocene quartz syenites
together with the strongly alkalic and ferroan
Early Jurassic nepheline syenites, define endmember compositions on this modified alkalilime index of Peacock (1931; Fig. 8B). In terms of
aluminum saturation (Maniar and Piccoli, 1989),
the Eastern Peruvian granitoids are peraluminous
to mildly metaluminous (A/CNK = 0.951.13).
The exceptions are the strongly peraluminous
Late Triassic Carabaya suite (A/CNK = 0.98
1.42), volumetrically minor Oligocene stocks
that show a metaluminous character (A/CNK =
0.910.97), and the metaluminous to peralkaline Early Jurassic Allincapac Group nephelinebearing (alkali feldspar) syenites (Fig. 9). Peculiar features of both the Early Ordovician and
Late Triassic granites are their anomalously high
K2O/Na2O ratios. Taken together with the cationic classification of de la Roche et al. (1980),
which relates changes in silica saturation,
Fe/(Fe + Mg) ratio and plagioclase composition during differentiation to tectonomagmatic
settings of emplacement (Fig. 10), the Eastern
Cordilleran plutons can be grouped into: (1) Neoproterozoic anorogenic and transitional late- to
post-orogenic granitoids of Permo-Triassic age;
(2) Cordilleran-type, continental-arc related
intrusives from the Carboniferous to Permian,
and Late Ordovician and late Mesoproterozoic

Geological Society of America Bulletin, September/October 2009

1307

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.

SCAM-08

NAM-11a

100 m

100 m
ierror ellipses are 2

ierror ellipses are 2

SAM-21

COCA-302

100 m
100 m

ierror ellipses are 2

ierror ellipses are 2

COCA-262

COCA-298

.
100 m

ierror ellipses are 2

.F

100 m

100 m

ierror ellipses are 2

Figure 6 (on this and following three pages). Chronologically arranged concordia diagrams for selected samples from the Figure 4. Dashed
lines correspond to the inherited zircon component. Cathodoluminescence (CL) images of zircon crystals illustrate the respective mineral
domains analyzed. MSWDMean square of weighted deviates.

1308

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru

CAM-11a

CAM-33

50 m

100 m

ierror ellipses are 2

ierror ellipses are 2

NAM-05

CAM-35

.
50 m

100 m

ierror ellipses are 2

ierror ellipses are 2

CAM-40

CAM-45c

.
100 m

50 m
100 m

ierror ellipses are 2

ierror ellipses are 2

Figure 6 (continued).

Geological Society of America Bulletin, September/October 2009

1309

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.

NAM-02a

CAM-54

ierror ellipses are 2


100 m
100 m

ierror ellipses are 2

CAM-03

NAM-30

.
100 m

100 m
50 m

ierror ellipses are 2

ierror ellipses are 2

NAM-18

SAM-12a

100 m

100 m

ierror ellipses are 2

ierror ellipses are 2

Figure 6 (continued).

1310

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru

SAM-04a

CAM-23

ierror ellipses are 2

50 m

100 m
100 m
ierror ellipses are 2

SCAM-18

CAM-17

100 m
100 m

ierror ellipses are 2

ierror ellipses are 2

CAM-18

SCAM-17

.
100 m

100 m

100 m

error ellipses are 2

ierror ellipses are 2

Figure 6 (continued).

Geological Society of America Bulletin, September/October 2009

1311

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.

Modal Q (vol.%)

Oligocene (Andean)
E. Jurassic(N)
E. Jurassic (S)
L. Triassic

Granite

Permo-Triassic
Carbo-Permian

Granodiorite

Alk.-feldspar
Granite

Late Ordovician
Early Ordovician
Neoproterozoic
Mesoproterozoic

Modal A
(vol.%)
Neph.-bear
Alk.-feldspar
Syenite

Qtz. Syenite
Neph.-bear
Syenite

Qtz. Monzonite

Qtz.
Monzodiorite

Neph.-bear
Monzonite

Modal P
(vol.%)

Neph.
Monzosyenite

Nepheline
Syenite

Modal F (vol.%)
Figure 7. Modal composition of the Eastern Peruvian Cordilleran intrusives plotted on
the International Union of Geological Sciences (IUGS) classification diagram (Streckeisen,
1974). Modal proportions were determined by staining 10 10 cm polished rock slabs
for K-feldspar and plagioclase with Amaranth (C20HllN2Na3OlOS3) and Na-cobalt nitrite
[Na3CO (NO2)6] respectively. Slabs were digitally scanned and an image processing software
JMicroVision created by Nicols Roduit was used for calculating the modal surface areas
that were projected into the ternary Qtz-Ab-Plag space.

ably higher Rb/Ba values than the rest of the


Carboniferous-Permian plutonic belt (Figs. 4
and 11E). Given that the ratio of Rb/Ba
decreases with increasing whole-rock SiO2 content, this trend cannot be explained by simple
fractional crystallization but likely reflects an
increased proportion of K-bearing phases in
the source. The most recent (Oligocene) intrusive pulse is uniquely LILE enriched, with
mildly depressed Nb-Ta concentrations indicative of a contaminated I-type magma that was
emplaced marginally inboard of the principal
arc batholiths (Walawender et al., 1990).
DISCUSSION
Due to protracted and sometimes overlapping orogenic episodes that shaped the central
Andes, the interpretation of their Precambrian
evolution is fragmentary and often speculative.
Nevertheless, to shed additional light on the
evolution of the Peruvian segment of the western Gondwanan margin, we place the newly
constructed geochronological framework for
the Eastern Cordillera of Peru in a wider context
of magmatism along the proto-Andean margin
of Amazonia, and ultimately link it to tectonic
cycles affecting the western edge of the craton
since the Mesoproterozoic.
Mesoproterozoic (Sunss-Grenvillian)

periods, (3) Late Triassic syncollisional plutons and an Early Ordovician post-collisional
intrusive, and (4) Early Jurassic within-plate,
peralkaline plutons forming the cores of alkaline volcanic complexes, and usually located in
backarc settings. The metaluminous granite porphyries of the modern Andean orogenic cycle
that were emplaced in the Oligocene are magnesian and calcic, and exhibit cationic ratios that
overlap those associated with continental arc
settings (Fig. 10).
Trace-element data from the Eastern Peruvian intrusives broadly corroborate tectonic
regimes inferred from the major-element
chemistry. Although primarily dependent on
the sources and crystallization history of the
melt (Frost et al., 2001), trace-element compositions of granitoids have long been used
as first-order tectonic discriminators for granitoids (Harris et al., 1986; Barbarin, 1999).
The oceanic plagiogranite normalized, multi
trace-element plots (Pearce et al., 1984) for the
Peruvian granitoids reveal a lack of heavy rareearth element (HREE) fractionation characteristic of a garnet-dominated source, and display
an overall positive correlation between the
extent of the large-ion lithophile to high fieldstrength element enrichment (LILE/HREE)

1312

and known episodes of subduction-related magmatism (Fig. 11). Negative Nb-Ta anomalies
are observed in both the late Mesoproterozoic
plutons and within the Carboniferous-Permian
suite but are also mildly present in the Late
Ordovician and Oligocene intrusive rocks
(Figs. 11A, 11E, 11F, and 11H). In contrast,
the Permo-Triassic, Early Ordovician quartz
monzonite, and especially Neoproterozoic
(monzo)granitoids, exhibit strong Ba/Th and
Ba/Rb anomalies characteristic of anorogenic
magmatism, in addition to the lack of a typical high LILE/HFSE trace-element pattern
associated with subduction (Figs. 11D, 11F,
and 11G). However, such time-dependent
tectonomagmatic classification is somewhat
complicated by the regional variability within
coeval intrusive suites along the proto-Andean
margin. For example, uniformly elevated HFS
elements that are characteristic of rift- or
backarc-related plutonism displayed by the
southern, Early Jurassic, peralkaline nepheline
syenites are markedly different from a mild
subduction-zone, trace-element signature of
the contemporaneous quartz syenites from the
northern Pataz region (Fig. 11B). On a more
local scale, the early Mississippian, northern
Balsas-Callangate pluton shows consider-

Two granitoids (CAM-17 and 18) from the


central Satipo-Mariposa transect (Fig. 2) represent previously unrecognized, late Mesoproterozoic to early Neoproterozoic magmatic
occurrences in the Peruvian Andes outside of the
known extent of the Arequipa-Antofalla terrane.
Although limited to a few samples, the peraluminous, highSiO2, calc-alkaline to alkali-calcic
granitoids exhibit uniform Nb-Ta anomalies,
characteristic of subduction-related intrusive
suites. The 1071 23 Ma and 1123 13 Ma
ages hence suggest an active segment of
north-central Amazonia during the peak of the
1.30.9 Ga Rondonia-Sunss orogen, and provide direct evidence for the existence of autochthonous South America within 150 km from
the eastern limit of the Arequipa-Antofalla
block (Fig. 2). Blurred and partially convoluted zircon rims from the late Mesoproterozoic
Querobamba granite (SCAM-17; Fig. 6X) of
the southwestern Peruvian Cordillera (Fig. 2),
which yield a concordia age of 1123 12 Ma
and mantle extraction ages of 1200 Ma from
inherited cores, also exhibit extreme U enrichments. They are seen as reflecting a period of
post-emplacement metamorphism coeval with
both syntectonic orthogneisses from the eastern
Bolivian Nova Brasilndia belt (U-Pb zircon

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru

1.0

A-type
Lachlan FB

Peraluminous
high-SiO2
granites

Global A-type granitoids

0.8
0.7

Ferroan

0.6
Cordilleran I-type
granitoids

0.5

S-type
Lachlan FB

Magnesian

0.4
12
A-type Lachlan FB

B
8

Na2O+K2OCaO

Figure 8. Major-element chemistry of the Eastern Peruvian


Cordillera intrusives plotted
on the geochemical classification diagrams for granitoids
of Frost et al. (2001): (A) FeO
number vs. wt% SiO2 diagram showing the boundary
between ferroan and magnesian plutons based on a global
compilation of 175 A-type, 344
Mesozoic Cordilleran I-type
granitoids, and 95 peraluminous leucogranites, as well as
the Lachlan Fold Belt (LFB),
A (n = 67), I (n = 1155) and
S-type granitoids (n = 720;
Landenberg and Collins, 1996).
Shaded and stippled areas outline 95% of the analyses and the
bold dashed line is the boundary between calc-alkaline and
tholeiitic magmas as defied by
Miyashiro et al. (1970). Wholerock ferrous iron determined
by 2,2 bipyridine [(C5H4N)2]
complex solution photometry.
(B) Modified alkali-lime index
of Peacock (1931) showing
ranges for the alkalic, alkalicalcic, calc-alkaline, and calcic
rock series. For any given suite,
the SiO2 value where the modified Na2O + K2O CaO index
(MALI) equals 0 corresponds
to the original alkali-lime index
of Peacock.

Fe # (FeO/FeO+MgO)

0.9

Global A-type granitoids

Increasing
f(H2O)
in source

Alkalic

lcic

Peraluminous
high-SiO2 granites
S-type
Lachlan FB

ca
ali-

Alk

0
Ca

lc-

a
alk

lin

-4

Calcic

Cordilleran I-type
granitoids

Oligocene (Andean)

Carbo-Permian

E.Jurassic(S)

Late Ordovician

E. Jurassic(N)

Early Ordovician

L. Triassic

Neoproterozoic

Permo-Triassic

Mesoproterozoic

-8
50

60

70

80

SiO2 (wt%)

ages of 1113 56 Ma and 1110 8 Ma; Rizzotto, 1999), and a Ro Pichari granulite-grade,
garnet-bearing charnockite from southeastern
Peru that yielded an upper intercept U-Pb zircon
age of 1140 30 Ma (Dalmayrac et al., 1988).
Taken together, the Peruvian Mesoproterozoic intrusives represent the northern extension
of the 1.100.92 Ga collisional Sunss province

in eastern Bolivia (Boger et al., 2005), and may


temporally link the 1098 9 Ma gneissic basement inliers of the eastern Colombian Andes
(Garzn Massif; Restrepo-Pace et al., 1997) with
the 1.21.0 metamorphosed Arequipa-Antofalla
basement (Martignole and Martelat, 2003) and
ultimately, the 1.2 Ga granulite-grade metamorphic basement of the Argentinean western Sierras

Pampeanas (Casquet et al., 2006). Consequently,


a contiguous orogenic belt must have existed for
over 3500 km along the Amazonian margin during the assembly of Rodinia, involving a continental collision with the Laurentian Llano belt,
including Labrador, and Baltica between ca. 1080
and 970 (Fig. 12A; Hoffman, 1991; Sadowski
and Bettencourt, 1996; Sadowski, 2002).

Geological Society of America Bulletin, September/October 2009

1313

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.

3.0
Excess H2O melting of mafic source /
melting (semi)pelitic source

2.5

Al/(Na+K)

Peraluminous

2.0
Metaluminous

1.5

1.0

Peralkaline

0.5
0.6

0.7

0.8

0.9

1.0 1.1 1.2


Al/(Ca+Na+K)

1.3

1.4

1.5

Figure 9. Aluminum saturation index (ASI) plot for the Eastern Cordillera plutons of Peru
(Maniar and Piccoli, 1989). Al/Na + K and Al/Ca + Na + K are defined as molecular ratios
and take into account the presence of apatite so that rocks with ASI>1.0 are corundum normative and are termed peraluminous (Zen, 1988). For weakly peraluminous rock additional
Al-bearing in addition to feldspars may be Al-biotite, but more strongly peraluminous granitoids are usually characterized by muscovite, magmatic cordierite, and garnet or Al2SiO5
polymorph, depending on pressure of formation. Intrusives with ASI<1.0 and Na + K<Al
are labeled metaluminous and contain excess Ca usually accommodated by hornblende and
Ca-rich augite but lack either muscovite or Fe-Mg pyroxene. Water-saturated partial melts
of a mafic source tend to be increasingly peraluminous with an increase in H2O content due
to feldspar destabilization (Ellis and Thompson, 1986). Peralkaline rocks have Na + K>Al,
which results in crystallization of leucite and nepheline, with the excess Na being accommodated in Fe-Mg silicates such as sodic amphiboles and jadeite. See Figures 7 and 8 for
the symbol key.

Neoproterozoic (pre-Braziliano)
Three middle Neoproterozoic A-type quartz
monzonite and alkali feldspar granites from
this study are the first identified outcrops of
Neoproterozoic granitoid plutonism along the
western margin of ancestral Amazonia. The
752691 Ma, mostly metaluminous, alkalic to
alkali-calcic, K-feldspar-bearing granitoids are
found in spatially distinct localities proximal to
the Sunss-age plutonic rocks (Fig. 4), and exhibit
high LILE and HFSE abundances characteristic
of anorogenic, A-type plutonism (Bonin, 2007;
Figs. 10 and 11). Other intrusives of similar ages
have been identified in the northern segment
of the parautochthonous Arequipa-Antofalla
terrane of southern Peru, where zircons from
syntectonic dacitic dikes that cut the gneissic
basement yield lower concordia intercepts at
635 5 Ma (Loewy et al., 2004). Middle Neo-

1314

proterozoic ages were also recorded by A-type


orthogneisses intruded into basement of the
Laurentian Precordillera (Cuyania) Terrane in
northwestern Argentina (774 6 Ma; Baldo
et al., 2006), and a nephelinite-carbonatite
body reported from a Grenville age terrane in
the Sierra de Maz of the western Sierras Pampeanas (570 Ma; Casquet et al., 2008). Here,
middle Neoproterozoic magmatism was classified as anorogenic and attributed to a phase of
extensional tectonics during the proto-Iapetus
rifting of Rodinia. The Peruvian granitoids,
however, mark the northernmost and one of the
oldest Neoproterozoic occurrences of the preBrazilianoaged anorogenic magmatism along
the western margin of ancestral South America.
This activity might have commenced as early as
ca. 800 Ma in the form of synrift, ultrapotassic
mafic dikes and HFSE-enriched alkaline lava
flows underlying the intracontinental Puncovis-

cana basin of northwestern Argentina (Omarini


et al., 1999). It complements the 765680 Ma
ages from the Appalachian Blue Ridge A-type
granites and bimodal volcanics of Tollo et al.
(2004), predates the Laurentia-Gondwana separation at ca. 615570 Ma sensu stricto (Cawood
et al., 2001), and overlaps with the initiation of
Cawoods (2005) rift-to-drift related sedimentary sequences along the Pacific margin of protoGondwana, where the opening of the Iapetus
Ocean is better constrained. Taken cumulatively,
the timing of emplacement of anorogenic intrusives in southwestern Amazonia suggests a
protracted period of discrete, north-migrating
extensional events that might have heralded
the incipient fragmentation of Rodinia and formation of the proto-Iapetus oceanic crust earlier in Neoproterozoic than the conventionally
accepted 620550 Ma rifting of the Iapetan margin of Laurentia (Hoffman, 1991; Cawood et al.,
2001). This contention is further supported by
recent paleomagnetic models and tectonic syntheses that invoke an early breakup of Western
Rodinia by, or prior to 725 Ma, involving the
separation of Australia and Eastern Antarctica
from Western Laurentia and consequent opening
of the paleo-Pacific (Panthalassic) ocean (Powell
et al., 1993; Li et al., 2008). It appears that asynchronous northward separation of present-day
Western Amazonia similarly took place along
the Appalachian margin of then southern, and
presently, Eastern Laurentia (Hartz and Torsvik,
2002; Fig. 12B). Moreover, the distribution of
the Peruvian intrusives places spatial constraints
on the Laurentiaproto-Gondwana conjugate
margin during the Neoproterozoic. In particular, it gives credence to the recently proposed
paleogeographic reconstructions where Western
Amazonia, Arequipa-Antofalla, western Sierras
Pampeanas, and East Laurentia are juxtaposed
within Rodinia (Loewy et al., 2003).
Early Paleozoic (Famatinian)
The extent of early Paleozoic plutonism
within the Eastern Peruvian Andes in our study
is limited to a single alkali feldspar granite
(SAM-04a) from the eastern margin of the
Cuzco batholith, which yields a Late Ordovician
crystallization age of 446.5 Ma, with significant
Mid-Cambrian inheritance (508 Ma; Table 1).
The central Peruvian Carboniferous plutons also
contain abundant inherited Late Ordovician and
Late Cambrian zircons (446 Ma; 490514 Ma).
These ages are broadly coeval with those from
the Sitabamba granodioritic orthogneisses identified in the north-central segment of the Eastern
Cordillera (445.9 2.4 Ma; 442.4 1.4 Ma),
the nearby Balsas paragneiss (477.9 4.3 Ma),
and the central Cordilleran Pacococha quartz

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru

R2 = 6Ca + 2Mg +Al (millications)

1500

Continental arc
Post-collisional
uplift

1000
Peralkaline
intrusives

Late orogenic

500
Anorogenic
Post-orogenic

0
0

1000
2000
3000
R1 = 4Si 11[Na+K] 2[Fe+Ti] (millications)

Figure 10. Geotectonic discrimination plot for the Peruvian Eastern Cordilleran granitoid
rocks based on de la Roche et al. (1980) multicationic R1R2 diagram (after Batchelor and
Bowden, 1985). Symbols are the same as those in Figures 7, 8, and 9.

monzonite dated at 474.2 3.4 Ma (Chew


et al., 2007). Whereas the orthogneisses and the
Cuzco alkali granite exhibit high LILE/HFSE
ratios coupled with profound Nb-Ta anomalies,
the Pacococha quartz monzonite shows none
(Fig. 11). Its trace-element pattern is, however,
characteristic of late- to post-collisional, calcalkaline suites akin to the Cordillera de Carabaya
granitoids. Combined with the refined ages for
the emplacement of the arc-derived San Nicols
batholith between 473 3 and 464 4 Ma
(Mukasa and Henry, 1990; Loewy et al.,
2004), the early Paleozoic age data imply that
the northern projection of the Famatinian arc
system of NW Argentina extended along the
Arequipa-Antofalla terrane, whereas the Latest
Ordovician arc granitoids lay inboard throughout the Eastern Cordilleran of Peru up to 6S.
Although a detailed paleogeographic reconstruction is outside the scope of our study,
this poorly understood time period deserves
a closer tectonodynamic examination. Here a
model is offered for the apparent early Paleozoic orogenic bifurcation, by relating the
timing and spatial distribution of granitoid
emplacement to metamorphism recorded in the
basement lithologies.
In Peru, the Paracas fault boundary (Fig. 2)
sharply truncates the Arequipa-Antofalla terrane at 14S (Dalmayrac et al., 1980) with a
lack of the Grenvillian basement ages to the

north, interpreted by Ramos (2008) to represent


a separate crustal block, i.e., the Paracas terrane. This terrane would account for the Early
Ordovician compressional event in northeastern
Peru, and possibly simplifies the outline of the
central Gondwanan paleoPacific margin that
does not require an embayment north of the
Paracas Peninsula, as suggested by Chew et al.
(2007). The presence of any significant amount
of sialic basement beneath the Peruvian Western
Cordillera since at least the Middle Cretaceous,
however, is refuted on the grounds of the isotopically extremely primitive intrusives of the
Coastal Batholith and the Cordillera Blanca of
Peru, a dominantly mafic underlying crust with
average density of ca. 3.0 g/cm3 and a complete
absence of zircon inheritance (Couch et al., 1981;
Polliand et al., 2005; de Haller et al., 2006). In
our model, somewhat analogous to the Argentinean Puna (Ramos, 1988), the Early Ordovician Peruvian margin experienced re-docking
of the parautochthonous Paracas-ArequipaAntofalla terrane against the Gondwanan
margin dominated by the protoArica bend
(Fig. 12C). We interpret the Vijus Group calcalkaline andesites overlying the Maranon Complex metapelites in the northern Pataz region of
the Eastern Cordillera (Haeberlin, 2002) as evidence for a Late Cambrian arc developing on a
thinned forearc crust similar to the Faja Eruptiva
de la Puna Oriental of NW Argentina (Ramos,

1988). The resuturing of Paracas-ArequipaAntofalla terrane closed the marginal basin that
hosted the pre-Famatinian Old Maran sediments of Chew et al. (2007), and resulted in the
emplacement of peridotite lenses as thrust slices
within thermally overprinted Maran metasediments (Haeberlin, 2002). To the south, the
original west Gondwana (passive) metasedimentary sequences inboard of the present
Arequipa terrane are probably only preserved
as thrust slices flanking the inverted Triassic rift
from Ayacucho to the northern Bolivian border.
Following the collision, a new continental arc
would have been established upon the mature
and thickened Paracas-Arequipa-Antofalla crust
at ca. 473464 Ma, in response to westward
jumping of the subduction zone over the newly
accreted block (Fig. 12D). Granitoids intruded
into the western margin of Famatina, in response
to subduction of the intervening Iapetus oceanic
lithosphere during the approach of the Cuyania
terrane farther south (Ramos, 2004), correspond
to arc-derived granites of the San Nicols batholith emplaced along the coastal Arequipa block
and would have likely extended northward
along the suspect Paracas terrane. We suggest
that a major shift in plate motions that produced
a rapid northward drift and clockwise rotation
of the Antofalla segment during the Middle
Ordovician (>12 cm/a; Forsythe et al., 1993),
effectively ceased arc activity, and impinged
the Paracas-Arequipa-Antofalla block upon the
south Eastern Cordillera due to the shape of
the Arica orocline. The resultant detachment
of the Oaxaquia terrane left behind the (Paracas)
mafic and isotopically primitive lower crustal
substrate that has been modeled by Couch et al.
(1981) and Polliand et al. (2005) as 3.0 g/cm3
dense basaltic underplate. The northeastward
transport might have occurred along a strikeslip fault system that is represented by the
600-km-long Ro Maran crustal lineament
(Fig. 12E). The present-day location of Paracas
terrane as the Oaxaquia-Acatln microcontinent in Mexico is suggested on the basis of the
early Paleozoic faunal and paleomagnetic correlations with Gondwanan of NW Argentina
(Snchez Zavala et al., 1999; Keppie et al., 2008).
The 20 Ma magmatic gap in the Late Ordovician was characterized by turbiditic sedimentation within newly formed depocenters that were
underlain by isotopically juvenile lower crust,
both north of the Arequipa (Contaya Formation, Dalmayrac et al., 1980; Young Maran;
Chew et al., 2007), and behind the rotated Antofalla block (Puna turbidites; Bahlburg and Herv,
1997). In the terminal phase of the Maran
orogeny, and following a reversal in subduction
vectors, the trench stepped in to the modern-day
position along the Peruvian coastline and the

Geological Society of America Bulletin, September/October 2009

1315

Rock/Ocean Ridge Granite

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.

100

.A

Oligocene
(Andean)

.B

E. Jurassic
Cordill. Carabaya (south)
Pataz (north)

10

Rock/Ocean Ridge Granite

0.1

.C
100

Late Triassic

.D

Permo-Triassic

10

Rock/Ocean Ridge Granite

0.1
100

.E

Carbo-Permian

.F

Balsas-Callangate (north)

Late Ordovician
Early Ordovician

10

Rock/Ocean Ridge Granite

0.1
100

.G

Neoproterozoic

.H

L. Mesoproterozoic

10

0.1
K RbBa Th Ta Nb Ce P Hf Zr SmTi Y Yb K RbBa Th Ta Nb Ce P Hf Zr SmTi Y Yb
Figure 11. Ocean-ridge, granite normalized, selected trace-element patterns for the Eastern Peruvian
Cordillera intrusives through time. Normalizing values taken from Pearce et al. (1984).

1316

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru
arc reestablished along the length of the Eastern Cordillera from 7S to 14S (Fig. 12F), as
indicated by the Cuzco batholith sample and
possibly comagmatic, calc-alkaline Ollantaytambo pyroclastic volcanics (Bahlburg et al.,
2006). The arc-derived granitoids in the northern Eastern Cordillera subsequently recorded
a high-grade metamorphic event equivalent to
mid-crustal conditions of 750 C and 12.8 kbar
(Chew et al., 2005). According to 40Ar/39Ar ages
from metamorphic micas (Cardona et al., 2006),
this last orogenic episode of the Early Phanerozoic along this segment of the proto-Andean
margin may be only slightly younger than the
442 Ma emplacement age, or alternatively it
may be related to the Devonian Chanic orogeny
that accompanied docking of Chilenia against
Cuyania (Keppie and Ramos, 1999).
Middle Paleozoic (Gondwanide)
Following the Devonian magmatic lull, a
continental arcderived intrusive belt formed
along the Peruvian margin of Gondwana in
the Early to Middle Mississippian (Fig. 12G).
Locally, this marked the onset of the panPacific Gondwanide Orogeny, a terminal phase
in the Neoproterozoic Terra Australis Orogen
(Cawood, 2005). Here, we address the entire
Gondwanide cycle (350190 Ma) as a genetically linked succession of magmatic flare-ups
that can be related to discrete steps in the final
agglomeration and initial breakup of the Pangea
supercontinent. The beginning of the orogenesis
in the Carboniferous, known as the Eohercynian
phase in Peru (Mgard et al., 1971), was characterized by the synchronous emplacement of
composite calc-alkaline hornblende, biotitebearing granodioritic and granitic batholiths
over 1200 km of strike of the Eastern Cordillera
(320.0 3.8 Ma at 6.9S and 324.1 5.3 Ma
at 13.2S). The similarity between the width of
the early Carboniferous belt (2030 km) and
modern arc plutonic tracts of central Chile and
the Aleutians, together with generally flat HREE
profiles, argue for intrusions within a relatively
thin crust (<40 km) that was underlain by a moderately steep, eastward-dipping subduction zone.
Volumetrically rivaled only by the subsequent
Permo-Triassic intrusive province of the central Eastern Cordillera (Fig. 2), these granitoids
feature an array of intrusive contacts indicative
of the interplay between mafic and felsic melts
(Fig. 3F). Their systematically depleted HFS
elements, but mature isotopic signatures and
the presence of upper crustal xenoliths point to
arc-derived magmas that extensively assimilated the basement country rocks (Macfarlane,
et al., 1999). An increased sample density due
to better exposure allowed for identification

of two separate pulses of magmatism from


the Carboniferous to the Permian (Fig. 5B), in
addition to revealing latitude-dependent compositional patterns (Fig. 11). The northern 313
320 Ma Balsas, Callangate, and segments of
the Pataz batholith (Fig. 2) display extreme
Rb/Ba ratios that cannot be explained by simple
fractional crystallization in a haplogranitic system (Harris and Inger, 1992; Patio Douce and
Beard, 1995) but likely reflect source parochialism. This increasing pelitic character of granitoid
sources northward along the axis of the Eastern
Cordillera is consistent with the changing basement sedimentary component from proximal
psammites to deeper marine shales caused by
tectonic subsidence of the Ordovician-Silurian
basin (Gohrbandt, 1992; Haeberlin, 2002). Furthermore, a 5 Ma gap in the Early Pennsylvanian
is contemporaneous with the 310.1312.9
3.0 Ma ages of high-grade metamorphism
recorded by foliated garnet-biotite paraschists in
the central Maran complex of the Eastern Cordillera (600 C and 11 kbar; Chew et al., 2005),
and broadly overlaps with the timing of the
high-pressure, low-temperature Toco tectonic
event that folded the Sierra del Tigre turbidites in
northern Chile (Bahlburg and Breitkreuz, 1991).
Although geochemical patterns from the Eastern Cordillera batholiths of Peru have recently
been interpreted as evidence for a possible Late
Pennsylvanian accretionary event (Mikovi et
al., 2005), no field evidence currently exists for
the presence of exotic crustal fragments north of
the Arequipa-Antofalla terrane, except for the
Early Cretaceous Amotape and Loja-Olmos terranes (Mourier, et al., 1988). Coupled with the
relatively short duration, the magmatic gap is
therefore compatible with models of increased
convergence rate (Ramos, 1988) that caused
compression accompanied by shallowing of the
subducted oceanic slab and an eventual pause in
arc activity. Conspicuously, this time period corresponds precisely to the onset of the OuachitaAlleghenian orogeny (Rast, 1984), which
marked the simultaneous collision between the
South American segment of northern Gondwana, Western Africa, and the Appalachian
Gulf Coast tract of Euramerica. Because the
continental crust is an excellent stress transmitter above its brittle-ductile transition (Lacombe,
2007), rigid Amazonia could have easily propagated the paleostress field orthogonally to the
Peruvian continental edge as it impacted southern North America at ca. 320 Ma, in a similar
fashion to what was recorded within the conjugate plate in mechanically twinned calcites
2100 km north from the Ouachita orogenic front
(Craddock et al., 1993). The sudden arrest of
northward-moving Gondwana would increase
plate coupling along the southwestern margin,

thus causing the Toco compression and resulting


in the trenchward motion of the overlying plate
that induced a temporary flat-slab geometry that
was responsible for the cessation of magmatism
(Kay and Abbruzzi, 1996; Fig. 12H).
The middle Permian to Late Triassic (Hercynian) stage of the Gondwanide orogeny in
Peru saw the emplacement of partially magmatized and compositionally restricted, highSiO2
granites slightly inboard of the Carboniferous
arc. The peak of 60 Ma of magmatism occurred
at 235240 Ma (Fig. 5B), during a period of
well-documented extensional block tectonics
and contemporaneous with the formation of the
Mitu basins (Mgard, 1978). These depocenters
consisted of a series of NW-SEtrending, faultcontrolled grabens, and initially accumulated
synrift continental epiclastics and progressively
more intercalated, bimodal peralkaline ignimbrites to tholeiitic basalts up the stratigraphic
column (Carlotto, 1998). Southward, Mitu
Group mafic volcanism changes compositionally from tholeiites to phonolites and tephrites
that crop out within the Putina Synclinorium
(Laubacher et al., 1988)a feature ascribed to
maturing and deepening of an aborted rift. Our
newly obtained age from the southern Santa
Rosa rhyolitic tuff (SAM-17) of 226 10 Ma,
together with the U-Pb age of 219.7 0.6 Ma
from a Mitu rhyolite in the central Tarma
province (Fig. 2; Chew et al., 2005), suggests
that rift-related volcanism in Peru during the
Paleozoic-Mesozoic transition is generally
younger than the bulk of the Permo-Triassic plutonism. Combined with a southward younging
trend of intrusive ages where no plutonic rock
older than 245 Ma outcrops south of 11.5S, this
indicates that plutonism played an important
role in controlling the pattern of crustal extension and rifting by thermally weakening the
lithosphere along the Eastern Cordillera. The
shift from the Carboniferous, highLILE/HFSE,
calc-alkaline, arc-derived granitoids to Permian
post-tectonic, highTh-Rb, alkali feldspar granites without significant HFSE depletions, and
ultimately the Triassic alkaline Mitu Group
bimodal volcanics (Figs. 9, 10, and 11) constitutes a classic magmatic succession associated
with regions undergoing lithospheric thinning
following extensional collapse (Jones, 1991;
Turner et al., 1999, Xu et al., 2007). Unlike the
coeval Choiyoi Granite Province in the Frontal
Cordillera of northern Chile, however, the termination of Carboniferous arc activity in Peru
and formation of post-tectonic granites cannot
be explained by the accretion of allochthonous
crust in the earliest Permian (i.e., terrane X),
as postulated by Mpodozis and Kay (1992). The
reasons for this are (1) absence of an identified
exotic terrane west of the Maran lineament in

Geological Society of America Bulletin, September/October 2009

1317

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.

.A

.B

L. Mesoproterozoic
(1.001.15 Ga)

Baltica

PAA

RN
nvil
Gre

W.
Africa

SU
?
?

Late Cambrian
(485500 Ma)

Iapetus
Ocean

W.
Africa

SL

AM

RN

.C

Neoproterozoic
(0.570.77 Ga)

SL
SF

SU
PAA

RP

le

Laurentia

AM

Laurentia

Baltica

Vijus
arc
Iapetus
Ocean

RN
SU

PAA

Laurentia

eanas
Pamp

Congo

Proto-Iapetus
Ocean

SF

.E

Middle Ordovician
(465480 Ma)
E. Iapetus
Ocean
Av
alo
Rheic nia

Pacococha/
Balsas

AM

BR SF

YM A SU
-A
OM

ia.a

atin

rc

Sedimentary platforms
Ophiolitic lenses

AM CENTRAL AMAZONIA

Magmatism arc/rift-extension

BR BRASILLIANO PROV.
C CHILENIA
F FAMATINIA

(MARONI,VENTUARI)

Plate vector

SU
AA

BR SF

BR SF

RP

SL

RP

PC

AM

RN

Panthalassic
Ocean

SL

AM

RN

Panthalassic
Ocean

RP

L. OrdovicianSilurian
(420445 Ma)
Sitabamba

Paracas

Contaya
turbidites

OM

am

A
-F
PA icolas

n
Sa

PC

E. Rheic Ocean

Maran
lineament

SL

SU

.F

Late Ordovician
(450460 Ma)

Ocean

RN

W. Iapetus
Ocean

RP

Congo

.D

BR SF

OM

RP

SL

AM

C
PC

OM
P
PAA
PC
PT

OLD MARAN
PAMPIA
PARACAS-AREQUIPA
PRECORDILLERA
PATAGONIA

RN
RP
SF
SL
SU

RIO NEGRO PROV.


RIO DE LA PLATA
SO FRANCISCO
SO LUIS
SUNSAS ARC / PROV.

Figure 12 (on this and following page). Schematic cartoon of the tectonic evolution of the present-day western proto-margin of the Amazonian craton. Boundaries of autochthonous tectonic provinces modified after Cordani and Sato (2000) and Tassinari et al. (2000). Dashed
areas encompass non-Gondwanan domains. Sequence of tectonic events for Mesoproterozoic are adopted and modified after Hoffman
(1991), Ramos and Aleman (2000), and Ramos (2008).

Figure 2, (2) no geochemical support for crustal


thickening beyond 32 km (i.e., 9 kbar) as indicated by the lack of significant Na enrichment
(Na2O = 3.54.5; Rapp and Watson, 1995),
and HREE fractionation in the Permo-Triassic
granitoids that would both be characteristic of

1318

a deep-sourced, garnet-bearing, residual mineralogy (Sm/Yb <2.5; Fig. 11), (3) decreasing Sr concentrations and larger Eu anomalies
(Eu/Eu* = 0.710.08) with increasing SiO2 content, combined with a lack of middle rare-earth
element (MREE) depletions (i.e., amphibole-

rich residue), implying a relatively dry,


feldspar-dominated source region at low pressures, (4) extensional relaxation following a
collision is expected to affect the buttressing
margin locally and cannot account for an episode
of craton-wide plutonism along Panthalassan

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru

.G

.H

Mississippian
(320345 Ma)

Ophiolitic lenses

Pennsylvanian
(310315 Ma)

Magmatism arc/rift-extension
Plate vector

N. America

A-

SU
A

Rheic
Ocean

BR SF

c
du
ub

RP

PC

RP

Africa

Panthalassic C
Ocean
PT

.J

PermianTriassic
(225280 Ma)

Africa

SL

RN AM

Euramerica

.K

SL

RN AM
SU

BR SF

BR SF

ar

RP

C
PT

A-A

Fm

PC

PaleoPacific
Ocean

SU

e
at

South
Africa

SL

RN AM
ol

RP

C. A
tla
Oce ntic
an

oc

PT

EarlyMid. Jurassic
(160180 Ma)

Allincapac

A-A

SUNSAS ARC / PROV.

Africa
Carabaya
anatexis

BR SF

A
-A

C
Panthalassic
Ocean

L. TriassicE. Jurassic
(190220 Ma)

BRASILLIANO PROV.
CHILENIA
FAMATINIA
OLD MARAN
PAMPIA
PARACAS-AREQUIPA
PRECORDILLERA
PATAGONIA
RIO NEGRO PROV.
RIO DE LA PLATA
SO FRANCISCO
SO LUIS

Ch

SU

N S slab
roll-back &
break off

BR
C
Iberia
F
OM
P
PAA
PC
PT
RN
RP
SF
SL
SU

Euramerica

Euramerica
Mitu Gr.

PC

PT

.I

BR SF

AA

n
tio

(MARONI,VENTUARI)

SL

AM

SU

s
at
Fl

AM CENTRAL AMAZONIA
Africa

RN
Panthalassic
Ocean

Sedimentary platforms

Ouachita Alleg

SL

AM

RN

ian

n
he

P
South
Africa

Pacific
Ocean

PT

PC

RP

South
Africa

PC

Figure 12 (continued).

Gondwana between 200 and 300 Ma (Kay et al.,


1989; Veevers et al., 1994).
An alternate hypothesis invoking slab shallowing has been proposed in several studies to
explain the diminishing volume of the magmaproducing mantle wedge below the Chilean segment of the (proto)Andean margin in the Early
Permian (Pankhurst et al., 1988), and Neogene
(Kay and Mpodozis, 2002). This model also
conflicts with the apparent normal crustal thicknesses, and fails to explain a number of other

magmatic and sedimentary patterns in Peru


during the Late CarboniferousEarly Permian:
(1) a minimal inboard shift in the locus of
magma production from the Carboniferous
arc to the adjacent Permian-Triassic granitoid
province, which is not expected if 60% of the
continental mantle is displaced up to 600 km
inland from the trench during a phase of shallow subduction (Kay and Abbruzzi, 1996),
(2) a lack of syndepositional deformation in
the early Permian shallow marine mudstones

of the Copacabana Group is incompatible with


strong plate coupling as the angle of subduction
falls below 10 (Dumitru et al., 1991; Gutscher,
2002), (3) the formation of post-tectonic granitoids preceded and overlapped with Triassic
extension (Ramos and Kay, 1991). Instead of a
collisional event or a scenario that appeals to tectonic underplating, we favor a model similar to
that proposed by Franzese and Spalletti (2001)
for the pre-rift evolution of the Neuqun Basin in
central Chile. We suggest that the transition from

Geological Society of America Bulletin, September/October 2009

1319

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.
the continental arc to post-tectonic magmatism
is related to the early Permian counterclockwise
block rotation observed by Rapalini (1998) in
the forearc of the Argentinean Ventana fold belt.
Development of strike-slip tectonics subparallel
to the protoAndean margin due to a change in
plate vector kinematics would have been accompanied by a strong reduction in the orthogonal
convergence vector, resulting in thermal weakening of the subducting slab below ~100 km depth
(Gerya et al., 2004). This mechanism could
account for progressively diminished volumes of
emplaced arc magmas from 300 Ma to 285 Ma
in Peru and agrees well with the 2030 Ma delay
necessary to trigger slab breakoff after the slowdown initiation (Wong et al., 1997; Arcay et al.,
2007; Vos et al., 2007). Operating in concert, the
northwestward-directed obliquity and concave
shape of the oroclinal margin generated a concomitant north-to-southdirected rollback of the
underlying plate, which in turn imposed a dextral strike-slip lithospheric wrenching along the
west-central Gondwanan margin of Pangea. This
explains the diachronous appearance of early
Mitu-type, transtensional duplexes from Peru,
via Bolivia to northern Chile and Argentina, and
the eventual cessation of arc activity due to slab
steepening and eventual detachment, represented
by the 280270 Ma minimum in the 206Pb/238U
zircon ages (Figs. 5B and 12I).
A relatively primitive isotopic signature of
Permo-Triassic plutonism in Peru (Nd = 2
to +2; Mikovi et al., 2007), and the anomalous enrichment in incompatible elements
(Th = 2050 ppm; Rb = 100400 ppm), suggest
a genetic scenario that is fundamentally different from Carboniferous arc magmatism. After
the ca. 1015 Ma required for the asthenospheric
mantle to occupy a newly formed gap left by the
detached slab, decompression melting generated
basaltic magmas at the base of the lower crust,
resulting in a vast region of extensive crustal
melting from 260 to 200 Ma along the western
Gondwana (Kay et al., 1989). Mantle advection locally provided a heat source for thermal
doming and the development of the Cusco-Puno
Swell outboard of the Eastern Peruvian Cordillera (Chavez et al., 1996). This rock uplift is
seen as a major late Carboniferous to Middle
Triassic depositional hiatus east of the rift axis
(Rakotosolofo et al., 2006). Instead of wholesale assimilation of an isotopically mature upper
crust by primitive (and wet) arc melts, the posttectonic granites appear to have formed during
progressively larger degrees of dehydration melting of an amphibolitic lower crust, driven by heat
and fluids supplied from basaltic underplates
(e.g., Skjerlie et al., 1993; Hansen et al., 2002;
Sisson et al., 2005). The buoyancy contrast,
which initially inhibited the rise of mafic melts,

1320

disappeared in the Middle Triassic when rifting eventually tapped the lower crustal magma
sources and gave rise to the bimodal Mitu Group
volcanics. The terminal phase of Mesozoic tectonism between 220 Ma and 190 Ma was characterized by thermal subsidence and formation of
the post-rift Pucar Group limestones landward
of the Cusco-Puno Swell (Rosas et al., 2007),
together with the anatexis of the preheated and
fertile uppermost crust beneath the Cordillera de
Carabaya (Fig. 12J). The reduced Nb-Ta anomalies of the Late Triassic granites relative to their
Early Permian protoliths, yet largely overlapping
Sr-Nd isotope systematics, are consistent with
melting of an upper crustal, arc-derived igneous
protolith contaminated by a substantial metasedimentary component (Mikovi et al., 2007).
The systematic pattern of plotting inside the
field of within-plate granites and post-tectonic
granitoids on the genetic classification diagrams
of Pearce et al. (1984) and Harris et al. (1986),
respectively, further suggests a late orogenic to
anorogenic granitoid suite (Fig. 10).
Jurassic to Miocene (Andean)
The transition to compressional tectonism
along the western South American margin most
likely occurred as a consequence of its westward
drift away from the locus of the Karoo plume,
recorded by progressively younging, silicic magmatic episodes in NE Patagonia (188187 Ma),
the northern Antarctic peninsula (172162 Ma),
and western Patagonia (157153 Ma; Pankhurst
et al., 2000). It heralded the ultimate dispersal
of Gondwana in the Cretaceous and was followed by the establishment of a continental
arc that has intermittently characterized the
modern Andean orogeny for the past 180 Ma.
Evidence for renewed Jurassic magmatic activity along coastal Peru comes from the southern ca. 177 Ma Chala basalts (Romeuf et al.,
1995), and fossiliferous volcano-carbonates of
the Pelado Formation (190170 Ma; Roperch
and Carlier, 1992). Our age data indicate that
backarc intrusives were emplaced in both the
northern Pataz area (NAM-11a) and the southern Allincapac province (SAM-21; Figs. 2 and
12K). Although coeval and confined to the
same segment of the arc architecture, the Early
to Middle Jurassic plutons clearly illustrate
dramatic differences in the composition of the
post-Triassic crust below the Peruvian Eastern Cordillera (Figs. 7, 8, and 11). The highly
HFSE-enriched, SiO2-undersaturated magmas
of the Allincapac complex probably formed by
small degrees of partial melting of the shallowest asthenospheric mantle below the thinned
Mitu basin crust, and markedly contrast not only
with the coeval Huayillas quartz monzonites at

8S, but also with all other plutonic rocks of


the Eastern Cordillera of Peru. However, their
northern equivalents reflect higher degrees of
interplay with a thicker crustal column under the
Eastern Cordillera between 6S and 12S (high
Rb/Nb; Fig. 11). Finally, the youngest intrusives
detected in our survey appear to be genetically
linked to continually changing plate convergence angles since the Eocene. The quartz
syenites of Colcabamba province belong to the
inner arc of the late Incaic orogeny, which was
the locus of widespread crustal anatexis (Clark
et al., 1984). Melting of previously hydrated
arc crust might have been induced by hot
asthenospheric mantle filling the gap left by the
Nazca plate, as it reverted from flat to a normal angle of subduction of ~30 that prevailed
through the Neogene (James and Sacks, 1999).
CONCLUSIONS
A U-Pb geochronological framework is
constructed for granitoid intrusives along the
Peruvian segment of the protoAndean margin
of Western Amazonia since 1.1 Ga. The newly
identified late Mesoproterozoic to early Neoproterozoic granitoids of the Eastern Cordillera
of Peru represent the westernmost extent of
autochthonous Amazonia in the north-central
Andes. They were emplaced between the peak
and waning stages of the Grenville-Sunss
orogeny. In combination with evidence from the
Argentinean Puncoviscana magmatism, the early
Pan-AfricanBraziliano-aged anorogenic plutonism along Western Amazonia in eastern Peru
(690750 Ma) suggests an active protoIapetus
margin along much of the western cratonic edge
as it rifted from eastern Laurentia during the
global breakup of Rodinia. Although it remained
a nonaccretionary margin during the Phanerozoic, the majority of plutonic rocks emplaced
along west-central Gondwana are directly
related to the assembly and ultimate fragmentation of Pangea (180330 Ma). Contrasting with
NW Argentina and Chile, where the Pampean
and Famatinian arcs were long lived and extensive, renewed convergent tectonism during the
early Paleozoic did not result in any significant
magmatic activity in Peru. Between the middle
Carboniferous and Early Jurassic, however, 85%
of the exposed Eastern Cordilleran batholiths
were emplaced. This period saw the establishment of a 1400-km-long, arc-related intrusive
belt that was succeeded by equally voluminous
Permo-Triassic, post-tectonic granitoids. Yet
another shift in the overall stress regime during the Late Triassic heralded the onset of the
modern Andean orogenic cycle that was initially
characterized by backarc plutonism as the focus
of arc activity shifted to the Western Cordillera.

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru
ACKNOWLEDGMENTS

Samples from the central Coasa, Aricoma, Limbani, and the southernmost Limacpampa plutons
were generously donated by Dr. Daniel Kontak and
Professor Allan Clark of Queens University, Canada.
Victor Carlotto and Jos Machar of the Peruvian
National Institute of Geology, Mines and Metallurgy
(INGEMMET) are thanked for the logistical support
and expertise in the field. We are particularly thankful to Ing. Seijas and Ing. Santillana of the Compaa
(CIA) Minera Poderosa for their continual logistical
support during the 20042006 field seasons. Massimo
Chiaradia provided many useful comments and advice
on granitoid petrology that helped focus and clarify
the ideas presented. This research was supported by
the Fonds Nationale Suisse de la Recherch Scientifiques (FNS) grant awarded to Urs Schaltegger.
APPENDIX
Zircons were extracted from 58 selected samples
ranging in weight between 1 and 5 kg following the
conventional mineral separation procedures using
gravimetric and magnetic methods including Wilfley
table (sub-300 micron, densest mineral concentrate),
heavy-liquid separation (diiodomethane; >3.32 g/cm2)
and Frantz magnetic separation (>1 A). Nonmagnetic,
euhedral and inclusion-free grains between 50 and
200 m in length were handpicked, and a representative set of 20 grains per sample was mounted in 1-cmthick epoxy blocks. These zircons were ground and
polished to reveal internal surfaces. They were subsequently carbon coated and imaged at the University
of Lausanne scanning electron microscopy (SEM)
facility using the CamScan MV 2300 scanning electron microanalyzer in the CL mode under the operating conditions of 15 kV of accelerating potential and
1520 nA beam current. Prior to isotopic analyses,
the sample mounts were ultrasonically cleansed in a
5 vol% HNO3 solution for 30 min and wiped clean.
LA-ICP-MS U-Pb Analytical Technique
U-Pb dating of zircons was conducted using a
Thermo-Finnigan Element2 sector field ICP-MS
coupled to a New Wave UP-213 Nd YAG laser at Bergen University following the procedure of Koler et al.
(2002). The sample introduction system was modified
to enable simultaneous nebulization of a tracer solution and laser ablation of the solid sample (Horn et al.,
2000). Natural Tl (205Tl/203Tl = 2.3871; Dunstan
et al., 1980), 209Bi and enriched 233U and 237Np (>99%)
were used in the tracer solution, and aspirated to the
plasma in an Ar-He gas mixture through an Apex
desolvation nebulizer and a T-piece tube attached to
the back of the plasma torch. Helium, as the sample
carrier gas, was fed to plasma via the T-piece tube
by a separate line. The laser was operated at a repetition rate of 10 Hz and constant beam diameter of
20 m resulting in an energy density of ~5.56 J/cm2.
An automated stage was used to produce a 100 m
linear raster. Typical acquisitions consisted of a 30 s
measurement of reference gas blank analytes (203Tl,
205
Tl, 209Bi, 233U, and 237Np), followed by a 150 s long
measurement of U and Pb signals from zircon along
with the continuous signal from the aspirated solution. The data were acquired in time-resolved, peakjumping, pulse-counting mode with 1 point measured
per peak for atomic masses of 202 (fly back), 203 and
205 (Tl), 206 and 207 (Pb), 209 (Bi), 233 (U), 237
(Np), 238 (U), 249 (233U oxide), 253 (237Np oxide), and
254 (238U oxide). Analyses of zircon reference material

91500 (1065 Ma; Wiedenbeck et al., 1995) systematically bracketed every 15 zircon unknowns. Raw data
were corrected for dead time of the electron multiplier
and processed off line in a spreadsheet-based program
(LAMDATE; Koler et al., 2002). Concordia diagrams were generated by the Microsoft Excelbased
tool, Isoplot version 3.41 (Ludwig, 2003). Data reduction included correction for gas blank, laser-induced
elemental fractionation of Pb and U, and instrument
mass bias (Koler and Sylvester, 2003). Formation
of oxides of U and Np was corrected by adding signal intensities at masses 249, 253, and 254 to the
intensities at masses 233, 237, and 238, respectively.
Although no common Pb correction was applied to
the data, great care was taken during the integration
of time-resolved signals in LAMDATE, and only the
most stable segments were selected, therefore largely
reducing the effect of 204Pb and Pb loss on the age calculation (Jackson et al., 2004). Furthermore, a graphical approach was employed whereby all data points
plotting to the right of apparent common Pb discordia
arrays in Tera-Wasserburg diagrams were interpreted
as having lost radiogenic Pb and were rejected (<3%
of the total of 770 analyses).
Bulk-Rock, Trace-Element Geochemistry
Selected minor- and trace-element concentrations were determined by X-ray fluorescence (XRF)
on pressed powder pellets using a Phillips PW 2400
sequential spectrometer equipped with a rhodium tube
at the University of Lausanne, Switzerland. A protocol developed by J. Michael Rhodes (University of
Massachusetts, USA) and based on modified procedures of Norrish and Chappel (1967) and Reynolds
(1967) was used for correction of nonlinear background, inter-element interferences, and variations in
mass absorption ( coefficients and/or Compton scatter). The trace-element accuracy by the XRF method
is 5%. Fragments of the Li2B4O7 diluted and fused
glass disks were analyzed for additional trace elements by laser ablationinductively coupled plasma
mass spectrometry (LA-ICP-MS) at the University of
Lausanne. The assembly consisted of a Perkin-Elmer
ELAN 6100 quadrupole ICP mass spectrometer
equipped with a dynamic reaction cell and coupled
to a Lambda Physik 193 nm EXCIMER (ArF) laser
that was fired at a frequency of 10 Hz delivering beam
energies between 140 and 200 mJ. The trace-element
data were acquired by averaging three 80120 m
analyses per disk (sample) over time intervals of 40
50 s on the peak transient signal and were repeatedly
normalized to the National Institute of Standards
and Technology (NIST) standard reference material
(SRM) 612 standard glass. Off-line data reduction of
time-resolved signals was performed on a Lotus 123
macrobased spreadsheet package (LAMTRACE)
written by Simon Jackson of Macquarie University,
Australia. The LA-ICP-MS REE data are accurate to
within 1% (La) or 6% (Lu) on the basis of duplicate analyses. We have chosen to use Y, Zr, Nb, Sr, Rb,
Ni, Cr, V, Ga, Co, and Sc obtained by XRF, and the
ICP-MS values for Ba, Ta, Hf, Cs, REE, Pb, Th, and U.
LA-ICP-MS Zircon Geochemistry
The selected trace-element (Sc, Ti, Rb, Sr, Y, Nb,
Ba, Hf, Ta, Pb, Th, and U) and REE contents of previously dated zircon domains were also analyzed
by the LA-ICP-MS technique at the University of
Lausanne. We modified the protocol used to acquire
trace-element data from fused disks by lowering the
repetition rate and energy output to 5 Hz and 120 mJ,

respectively, while probing polished grain surfaces


along 20-m-wide linear raster for 1530 s on the peak
transient signal. Data were normalized to the highly
doped NIST SRM 610 standard glass, and the off-line
data reduction was conducted by LAMTRACE similarly to the fused-disk procedure.
REFERENCES CITED
Adams, C., and Miller, H., 2007, Detrital zircon ages of the
Puncoviscana Formation of NW Argentina, and their
bearing on stratigraphic age and provenance: Kiel,
Germany, 20th Colloquium on Latin American Earth
Sciences, p. 6869.
Allmendinger, R.W., Jordan, T.E., Kay, S.M., and Isacks,
B.L., 1997, The evolution of the Altiplano-Puna plateau of the Central Andes: Annual Review of Earth and
Planetary Sciences, v. 25, p. 139174, doi: 10.1146/
annurev.earth.25.1.139.
Arcay, D., Doin, M.-P., Tric, E., and Bousquet, R., 2007,
Influence of the precollisional stage on subduction
dynamics and the buried crust thermal state: Insights
from numerical simulations: Tectonophysics, v. 441,
p. 2745, doi: 10.1016/j.tecto.2007.06.001.
Bahlburg, H., and Breitkreuz, C., 1991, Paleozoic evolution
of active margin basins in the southern Central Andes
(northwestern Argentina and northern Chile): Journal
of South American Earth Sciences, v. 4, p. 171188,
doi: 10.1016/0895-9811(91)90029-K.
Bahlburg, H., and Herv, F., 1997, Geodynamic evolution
and tectonostratigraphic terranes of northwestern
Argentina and northern Chile: Geological Society of America Bulletin, v. 109, p. 869884, doi:
10.1130/0016-7606(1997)109<0869:GEATTO>
2.3.CO;2.
Bahlburg, H., Carlotto, V., and Cardenas, J., 2006, Evidence
of Early to Middle Ordovician arc volcanism in the
Cordillera Oriental and Altiplano of southern Peru,
Ollantaytambo Formation and Umachiri beds: Journal
of South American Earth Sciences, v. 22, p. 5265, doi:
10.1016/j.jsames.2006.09.001.
Baker, D.R., Conte, A.M., Freda, C., and Ottolini, L., 2002,
The effect of halogens on Zr diffusion and zircon
dissolution in hydrous metaluminous granitic melts:
Contributions to Mineralogy and Petrology, v. 142,
p. 666678.
Baldo, E., Casquet, C., Pankhurst, R.J., Galindo, C., Rapela,
C.W., Fanning, C.M., Dahlquist, J.A., and Murra, J.,
2006, Neoproterozoic A-type magmatism in the Western
Sierras Pampeanas (Argentina): Evidence for Rodinia
break-up along a proto-Iapetus rift?: Terra Nova, v. 18,
p. 388394, doi: 10.1111/j.1365-3121.2006.00703.x.
Barbarin, B., 1999, A review of the relationships between
granitoid types, their origins and their geodynamic
environments: Lithos, v. 46, p. 605626, doi: 10.1016/
S0024-4937(98)00085-1.
Bard, J.P., Dalmayrac, B., Marocco, R., and Mgard, F.,
1974, Extension et caractres des roches mtamorphiques prcambriennes du Prou: Comptes Rendus
Hebdomadaires des Sances de lAcadmie des Sciences, Srie D, v. 278, p. 30353038.
Batchelor, R.A., and Bowden, P., 1985, Petrogenetic interpretation of granitoid rock series using multicationic
parameters: Chemical Geology, v. 48, p. 4355, doi:
10.1016/0009-2541(85)90034-8.
Beckinsale, R.D., Sanchez-Fernandez, A.W., Brook, M.,
Cobbing, E.J., Taylor, W.P., and Moore, N.D., 1985,
Rb-Sr whole-rock isochron and K-Ar age determinations for the Coastal Batholith of Peru, in Pitcher, W.S.,
Atherton, M.P., Cobbing, E.J., and Beckinsale, R.D.,
eds., Magmatism at a plate edge: The Peruvian Andes:
London, Blackie and Son, p. 177202.
Benavides, V., 1999, Orogenic evolution of the Peruvian
Andes: The Andean cycle, in Skinner, B.J., ed., Geology and ore deposits of the Central Andes: Society
of Economic Geologists Special Publication, v. 7,
p. 61107.
Boger, S.D., Raetz, M., Giles, D., Etchart, E., and Fanning,
C.M., 2005, UPb age data from the Sunsas region of
Eastern Bolivia: Evidence for the allochthonous origin
of the Paragua Block: Precambrian Research, v. 139,
p. 121146, doi: 10.1016/j.precamres.2005.05.010.

Geological Society of America Bulletin, September/October 2009

1321

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.
Bond, G.C., Nickeson, P.A., and Kominz, M.A., 1984,
Breakup of a supercontinent between 625 Ma and
555 Ma: New evidence and implications for continental histories: Earth and Planetary Science Letters, v. 70, p. 325345, doi: 10.1016/0012-821X(84)
90017-7.
Bonin, B., 2007, A-type granites and related rocks: Evolution of a concept, problems and prospects: Lithos,
v. 97, p. 129, doi: 10.1016/j.lithos.2006.12.007.
Cardona, A., Cordani, U.G., Ruiz, J., Valencia, V., Nutman,
A.P., and Snchez, A.W., 2006, U/Pb detrital zircon
geochronology and Nd isotopes from Paleozoic metasedimentary rocks of the Maranon complex: Insights
on the proto-Andean tectonic evolution of the eastern
Peruvian Andes: V South American Symposium on
Isotope Geology (Punta del Este, Uruguay), Extended
abstracts, p. 208211.
Carlotto, V., 1998, volution andine et raccourcissement
au niveau de Cusco (1316S, Prou) [Ph.D. thesis]:
France, Universit de Grenoble, 159 p.
Carlotto, V., Tintaya, D., Crdenas, J., Carlier, G., and
Rodrguez, R., 2006, Fallas transformates PermoTrisicas: La falla Patacancha-Tamburco (sur del Per):
Resmenes XIII Congreso Peruano de Geologa.
Casquet, C., Pankhurst, R.J., Fanning, C.M., Baldo, E.,
Galindo, C., Rapela, C.W., Gonzlez-Casado, J.M.,
and Dahlquist, J.A., 2006, U-Pb SHRIMP zircon
dating of Grenvillian metamorphism in Western
Sierras Pampeanas (Argentina): Correlation with the
Arequipa-Antofalla craton and constraints on the extent
of the Precordillera Terrane: Gondwana Research, v. 9,
p. 524529, doi: 10.1016/j.gr.2005.12.004.
Casquet, C., Pankhurst, R.K., Galindo, C., Rapela, C., Fanning, C.M., Baldo, E., Dahlquist, J., Gonzlez Casado,
J.M., and Colombo, F., 2008, A deformed alkaline
igneous rockcarbonatite complex from the western Sierras Pampeanas, Argentina: Evidence for late
Neoproterozoic opening of the Clymene Ocean?: Precambrian Research, v. 165, p. 205220, doi: 10.1016/
j.precamres.2008.06.011.
Cawood, P.A., 2005, Terra Australis Orogen: Rodinia
breakup and development of the Pacific and Iapetus
margins of Gondwana during the Neoproterozoic and
Paleozoic: Earth-Science Reviews, v. 69, p. 249279,
doi: 10.1016/j.earscirev.2004.09.001.
Cawood, P.A., and Buchan, C., 2007, Linking accretionary orogenesis with supercontinent assembly: EarthScience Reviews, v. 82, p. 217256, doi: 10.1016/
j.earscirev.2007.03.003.
Cawood, P.A., McCausland, P.J.A., and Dunning, G.R.,
2001, Opening Iapetus: Constraints from the Laurentian margin in Newfoundland: Geological Society of
America Bulletin, v. 113, p. 443453, doi: 10.1130/
0016-7606(2001)113<0443:OICFTL>2.0.CO;2.
Chavez, R., Gil, W., Carlotto, V., Cardenas, J., and Jaillard,
E., 1996, 3rd International Symposium on Andean
Geodynamics (ISAG, Saint Malo, France): Paris,
Rsum tendu, IRD Editions (Institut de Recherche
pour le Dveloppement), p. 319322.
Chew, D.M., Schaltegger, U., Mikovi, A., Fontignie, D.,
and Frank, M., 2005, Deciphering the tectonic evolution of the Peruvian segment of the Gondwanan
margin: Paris, 6th International Symposium on
Andean Geodynamics (ISAG Barcelona), IRD (Institut de Recherche pour le Dveloppement) Editions,
p. 166169.
Chew, D.M., Schaltegger, U., Koler, J., Whitehouse, M.,
Gutjahr, M., Spikings, R.A., and Mikovi, A., 2007,
U-Pb geochronologic evidence for the evolution of the
Gondwanan margin of the north-central Andes: Geological Society of America Bulletin, v. 119, p. 697
711, doi: 10.1130/B26080.1.
Clark, A.H., Kontak, D.J., and Farrar, E., 1984, A comparative study of the metallogenetic and geochronological
relationships in the northern part of the Central Andean
tin belt, SE Peru and NW Bolivia, in Janelidze, T.V.,
and Tvalchralidze, A.G., eds.: Stuttgart, Germany,
VI International Association on the Genesis of Ore
Deposits (IAGOD) Symposium, p. 269279.
Clark, A.H., Farrar, E., Kontak, D.J., Langridge, R.J.,
Arenas, M.J.F., France, L.J., McBride, S.L., Woodman, P.L., Wasteneys, H.A., Sandeman, H.A., and
Archibald, D.A., 1990, Geologic and geochronologic

1322

constraints on the metallogenic evolution of the Andes


of southeastern Peru: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 85,
p. 15201583.
Cordani, U.G., and Sato, K., 2000, Crustal evolution of the
South American platform based on Nd isotopic systematics on granitoid rocks: Episodes, v. 22, p. 167173.
Cordani, U.G., Brito-Neves, B.B., and DAgrella-Filho,
M.S., 2003, From Rodinia to Gondwana: A review
of the available evidence from South America: Gondwana Research, v. 6, p. 275283, doi: 10.1016/S1342937X(05)70976-X.
Corfu, F., Hanchar, J.M., Hoskin, P.W.O., and Kinny, P.,
2003, Atlas of zircon textures, in Hanchar, J.M., and
Hoskin, P.W.O., eds., Reviews in Mineralogy and Geochemistry, v. 53, p. 469500.
Couch, R., Whitsett, R., Huehn, B., and Briceno-Guarupe,
L., 1981, Structures of the continental margin in Peru
and Chile, in Kulm, L.D., Dymond, D., Dasch, E., and
Hussong, D.M., eds., Nazca Plate: Crustal formation
and Andean convergence: Geological Society of America Memoir, v. 154, p. 703726.
Craddock, J.P., Jackson, M., van der Pluijm, B.A., and
Versical, R.T., 1993, Regional shortening fabrics in
eastern North America: Far-field stress transmission
from the Appalachian-Ouachita orogenic belt: Tectonics, v. 12, p. 257264, doi: 10.1029/92TC01106.
Dalmayrac, B., Laubacher, G., and Marocco, R., 1980,
Gologie des Andes pruviennes. Caractres gnraux
de lvolution gologique des Andes pruviennes:
Travaux et Documents de lORSTOM, v. 122, 501 p.
Dalmayrac, B., Laubacher, G., and Marocco, R., 1988,
Caracteres generales de la evolucin geolgica de los
Andes Peruanos: Instituto Geolgico Minero y Metalrgico del Per, Boletn 12, 313 p.
Dalziel, I.W.D., Dalla Salda, L., and Gahagan, L.M., 1994,
Paleozoic Laurentia-Gondwana interaction and the origin of the Appalachian-Andean mountain system: Geological Society of America Bulletin, v. 106, p. 243252,
doi: 10.1130/0016-7606(1994)106<0243:PLGIAT>
2.3.CO;2.
de Haller, A., Corfu, F., Fontbot, L., Schaltegger, U., Barra,
F., Chiaradia, M., Frank, M., and Alvarado, J.Z., 2006,
Geology, geochronology, and Hf and Pb isotope data of
the Ral-Condestable iron oxide-copper-gold deposit:
Central coast of Peru: Economic Geology, v. 101,
p. 281310.
de la Roche, H., Leterrier, J., Grandclaude, P., and Marchal,
M., 1980, A classification of volcanic and plutonic
rocks using R1R2 diagram and major-element
analyses and its relationships with current nomenclature: Chemical Geology, v. 29, p. 183210, doi:
10.1016/0009-2541(80)90020-0.
Dodson, M.H., 1973, Closure temperature in cooling geochronological and petrological systems: Contributions
to Mineralogy and Petrology, v. 40, p. 259274.
Dumitru, T.A., Gans, P.B., Foster, D.A., and Miller, E.L.,
1991, Refrigeration of the western Cordilleran
lithosphere during Laramide shallow-angle subduction: Geology, v. 19, p. 11451148, doi: 10.1130/
0091-7613(1991)019<1145:ROTWCL>2.3.CO;2.
Dunstan, L.P., Gramch, J.W., Barnes, I.L., and Purdy, W.C.,
1980, The absolute abundance and the atomic weight
of a reference sample of thallium: Journal of Research
of the National Bureau of Standards, v. 85, p. 110.
Ellis, D.J., and Thompson, A.B., 1986, Subsolidus
and partial melting reactions in the quartz-excess
CaO+MgO+Al2 O3 +SiO2 +H2 O system under waterexcess and water-deficient conditions to 10 kb:
Some implications for the origin of peraluminous
melts from mafic rocks: Journal of Petrology, v. 27,
p. 91121.
Forsythe, R.D., Davidson, J., Mpodozis, C., and Jesinkey, C.,
1993, Lower Paleozoic relative motion of the Arequipa
block and Gondwana: Paleomagnetic evidence from
Sierra de Almeida of Northern Chile: Tectonics, v. 12,
p. 219236, doi: 10.1029/92TC00619.
Franzese, J.R., and Spalletti, L.A., 2001, Late Triassic
Early Jurassic continental extension in southwestern
Gondwana: Tectonic segmentation and pre-break-up
rifting: Journal of South American Earth Sciences,
v. 14, p. 257270, doi: 10.1016/S0895-9811(01)
00029-3.

Frost, B.R., Barnes, C.G., Collins, W.J., Arculus, R.J., Ellis,


D.J., and Frost, C.D., 2001, A geochemical classification of granitic rocks: Journal of Petrology, v. 42,
p. 20332048, doi: 10.1093/petrology/42.11.2033.
Fuck, R.A., Brito Neves, B.B., and Schobbenhaus, C., 2008,
Rodinia descendants in South America: Precambrian
Research, v. 160, p. 108126, doi: 10.1016/j.precamres.
2007.04.018.
Gerya, T.V., Yuen, D.A., and Maresch, W.V., 2004, Thermomechanical modelling of slab detachment: Earth and
Planetary Science Letters, v. 226, p. 101116, doi:
10.1016/j.epsl.2004.07.022.
Gohrbandt, K.H.A., 1992, Paleozoic paleogeographic and
depositional developments on the central proto-Pacific
margin of Gondwana: Their importance to hydrocarbon accumulation: Journal of South American Earth
Sciences, v. 6, p. 267287, doi: 10.1016/0895-9811(92)
90046-2.
Gutscher, M.-A., 2002, Andean subduction styles and their
effect on thermal structure and interplate coupling:
Journal of South American Earth Sciences, v. 15,
p. 310, doi: 10.1016/S0895-9811(02)00002-0.
Haeberlin, Y., 2002, Geological and structural setting, age,
and geochemistry of the orogenic gold deposits at the
Pataz province, Eastern Andean Cordillera, Peru: Terre
et Environnement, v. 36, 182 p.
Haeberlin, Y., Moritz, R., Fontbot, L., and Cosca, M., 2004,
Carboniferous orogenic gold deposits at Pataz, Eastern
Andean Cordillera, Peru: Geological and structural
framework, paragenesis, alteration, and 40Ar/39Ar geochronology: Economic Geology and the Bulletin of the
Society of Economic Geologists, v. 99, p. 73112.
Hansen, J., Skjerlie, K.P., Pedersen, R.B., and de la Rosa, J.,
2002, Crustal melting in the lower parts of island arcs:
An example from the Bremanger Granitoid Complex,
west Norwegian Caledonides: Contributions to Mineralogy and Petrology, v. 143, p. 316335.
Harris, N.B.W., and Inger, S., 1992, Trace element modelling of pelite-derived granites: Contributions to Mineralogy and Petrology, v. 110, p. 4656, doi: 10.1007/
BF00310881.
Harris, N.B.W., Pearce, J.A., and Tindle, A.G., 1986, Geochemical characteristics of collision-zone magmatism,
in Coward, M.P., and Rie, A.C., eds., Collision tectonics: The Geological Society of London, Special
Publication, v. 19, p. 6781.
Hartz, E.H., and Torsvik, T.H., 2002, Baltica upside
down: A new plate tectonic model for Rodinia and
the Iapetus Ocean: Geology, v. 30, p. 255258, doi:
10.1130/0091-7613(2002)030<0255:BUDANP>
2.0.CO;2.
Hoffman, P.F., 1991, Did the breakout of Laurentia turn
Gondwanaland inside-out?: Science, v. 252, p. 1409
1412, doi: 10.1126/science.252.5011.1409.
Horn, I., Rudnick, R.L., and McDonough, W.F., 2000,
Precise elemental and isotope ratio determination
by simultaneous solution nebulization and laser
ablation-ICP-MS: Application to U-Pb geochronology:
Chemical Geology, v. 164, p. 281301, doi: 10.1016/
S0009-2541(99)00168-0.
Jacay, J., Sempere, T., Carlier, G., and Carlotto, V., 1999,
Late Paleozoic-Early Mesozoic plutonism and related
rifting in the Eastern Cordillera of Peru: Gottingen, 4th
International Symposium on Andean Geodynamics,
ORSTROM Collection Colloques et Sminaires,
p. 358362.
Jackson, S.E., Pearson, N.J., Griffin, W.L., and Belousova,
E.A., 2004, The application of laser ablationinductively coupled plasma-mass spectrometry to
in situ U-Pb zircon geochronology: Chemical Geology,
v. 211, p. 4769, doi: 10.1016/j.chemgeo.2004.06.017.
Jaillard, E., Hrail, G., Monfret, T., Daz-Martnez, E., Baby,
P., Lavenu, A., and Dumont, J.F., 2000, Tectonic evolution of the Andes of Ecuador, Peru, Bolivia and northernmost Chile, in Cordani, U., Milani, E.J., Thomaz
Filho, A., and Campos Neto, M.C., eds., Tectonic
evolution of South America: Rio de Janeiro, 31st International Geological Congress, p. 481559.
James, D.E., and Sacks, I.S., 1999, Cenozoic formation
of the central Andes: A geophysical perspective, in
Skinner, B.J., ed., Geology and ore deposits of the
Central Andes: Society of Economic Geologists Special Publication, v. 7, p. 125.

Geological Society of America Bulletin, September/October 2009

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


U-Pb geochronology and geochemistry of the proto-Andean granitoids of Peru
Jones, K.A., 1991, Palaeozoic continental margin tectonics
in southern Armorica: Journal of the Geological Society, v. 148, p. 5564, doi: 10.1144/gsjgs.148.1.0055.
Kay, S.M., and Abbruzzi, J.M., 1996, Magmatic evidence for
Neogene lithospheric evolution of the central Andean
flat-slab between 30S and 32S: Tectonophysics,
v. 259, p. 1528, doi: 10.1016/0040-1951(96)00032-7.
Kay, S.M., and Mpodozis, C., 2002, Magmatism as a probe
to the Neogene shallowing of the Nazca plate beneath
the modern Chilean flat-slab: Journal of South American Earth Sciences, v. 15, p. 3957, doi: 10.1016/
S0895-9811(02)00005-6.
Kay, S.M., Ramos, V.A., Mpodozis, C., and Sruoga, P.,
1989, Late Paleozoic to Jurassic silicic magmatism
at the Gondwanaland margin: Analogy to the Middle
Proterozoic in North America?: Geology, v. 17, p. 324
328, doi: 10.1130/0091-7613(1989)017<0324:LPTJSM>
2.3.CO;2.
Kempe, U., Gruner, T., Nasdala, L., and Wolf, D., 2000,
Relevance of cathodoluminescence for interpretation
of U-Pb zircon ages (with an example of application
to a study of zircons from the Saxonian Granulite
complex, Norway), in Pagel, M., Barbin, V., Blanc, P.,
Ohnenstetter, D., eds., Cathodoluminescence in geosciences: Berlin, Springer, p. 415455.
Keppie, J.D., 2008, The Guayape-Papalutla fault system: A
continuous Cretaceous structure from southern Mexico
to the Chorts block? Tectonic implications: Comment:
Geology, v. 36, doi: 10.1130/G24949C.1.
Keppie, J.D., and Bahlburg, H., 1999, Puncoviscana Formation of northwestern and central Argentina: Passive
margin or foreland basin deposits?, in Ramos, V.A.,
and Keppie, J.D., eds., Laurentia-Gondwana connections before Pangea: Geological Society of America,
Special Paper, v. 336, p. 139143.
Keppie, J.D., and Ramos, V.A., 1999, Odyssey of terranes
in the Iapetus and Rheic Oceans during the Paleozoic,
in Ramos, V.A., and Keppie, J.D., eds., LaurentiaGondwana connections before Pangea: Geological
Society of America, Special Paper, v. 336, p. 267276.
Kontak, D.J., Clark, A.H., Farrar, E., and Strong, D.F., 1985,
The rift associated Permo-Triassic magmatism of the
Eastern Cordillera: A precursor to the Andean orogeny,
in Pitcher, W.S., Atherton, M.P., Cobbing, E.J., and
Beckinsale, R.D., eds., Magmatism at a plate edge: The
Peruvian Andes: London, Blackie and Son, p. 3644.
Kontak, D.J., Clark, A.H., Farrar, E., Archibald, D.A., and
Baadsgaard, H., 1990, Late Paleozoic-early Mesozoic magmatism in the Cordillera de Carabaya, Puno,
southeastern Peru: Geochronology and petrochemistry: Journal of South American Earth Sciences, v. 3,
p. 213230, doi: 10.1016/0895-9811(90)90004-K.
Koler, J., and Sylvester, P.J., 2003, Present trends and
the future of zircon in geochronology: laser ablation
ICPMS, in Hanchar, J.M., and Hoskin, P.W.O., eds.,
Reviews in Mineralogy and Geochemistry, v. 53,
p. 327341.
Lacombe, O., 2007, Comparison of paleostress magnitudes from calcite twins with contemporary stress
magnitudes and frictional sliding criteria in the continental crust: Mechanical implications: Journal of
Structural Geology, v. 29, p. 8699, doi: 10.1016/
j.jsg.2006.08.009.
Lancelot, J., Laubacher, G., Marocco, R., and Renaud, U.,
1978, U/Pb radiochronology of two granitic plutons
from the eastern Cordillera (Peru)Extent of Permian
magmatic activity and consequences: International
Journal of Earth Sciences, v. 67, p. 236243.
Landenberger, B., and Collins, W.J., 1996, Derivation of
A-type granites from a dehydrated charnockitic lower
crust: Evidence from the Chaelundi Complex, Eastern
Australia: Journal of Petrology, v. 37, p. 145170, doi:
10.1093/petrology/37.1.145.
Laubacher, G., Sebrier, M., Fornari, M., and Carlier, G.,
1988, Oligocene and Miocene continental sedimentation, tectonics, and S-type magmatism in the southeastern Andes of Peru (Crucero Basin): Geodynamic
implications: Journal of South American Earth Sciences, v. 1, p. 225238, doi: 10.1016/0895-9811(88)
90001-6.
Li, Z.X., Bogdanova, S.V., Collins, A.S., Davidson, A., De
Waele, B., Ernst, E.E., Fitzsimons, I.C.W., Fuck, R.A.,
Gladkochub, D.P., Jacobs, J., Karlstrom, K.E., Lu, S.,

Natapov, L.M., Pease, V., Pisarevsky, S.A., Thrane, K.,


and Vernikovsky, V., 2008, Assembly, configuration
and break-up history of Rodinia: A synthesis: Precambrian Research, v. 160, p. 179210, doi: 10.1016/
j.precamres.2007.04.021.
Litherland, M., Aspden, J.A., and Jemielita, R.A., 1994, The
metamorphic belts of Ecuador: British Geological Survey: Overseas Memoir, v. 11, p. 147.
Loewy, S.L., Connelly, J.N., Dalziel, I.W.D., and Gower,
C.F., 2003, Eastern Laurentia in Rodinia: Constraints
from whole-rock Pb and U/Pb geochronology: Tectonophysics, v. 375, p. 169197, doi: 10.1016/S0040-1951
(03)00338-X.
Loewy, S.L., Connelly, J.N., and Dalziel, I.W.D., 2004, An
orphaned basement block: The Arequipa-Antofalla
basement of the central Andean margin of South America: Geological Society of America Bulletin, v. 116,
p. 171187, doi: 10.1130/B25226.1.
Ludwig, K.R., 2003, Isoplot 3.0: A geochronological toolkit
for Microsoft Excel: Berkeley Geochronology Center
Special Publication 4, 71 p.
Macfarlane, A.W., Marcet, P., LeHuray, A.P., and Petersen,
U., 1990, Lead isotopic provinces of the central Andes
inferred from ores and crustal rocks: Economic Geology and the Bulletin of the Society of Economic
Geologists, v. 85, p. 18571880.
Macfarlane, A.W., Tosdal, R.M., Vidal, C., and Paredes, J.,
1999, Geologic and isotopic constraints on the age
and origin of auriferous-quartz veins in the Parcoy
mining district, Pataz, Per, in Skinner, B.J., ed.,
Geology and ore deposits of the Central Andes: Society of Economic Geologists, Special Publication, v. 7,
p. 267279.
Mamani, M., Wrner, G., Hartmann, G., and Cassard, D.,
2005, A GIS-based isotope map of the Central Andes
(13S28S) and implication for ore formation: Barcelona, 6th International Symposium on Andean Geodynamics, IRD Editions (Institut de Recherche pour le
Dveloppement, Paris), p. 464467.
Maniar, P.D., and Piccoli, P.M., 1989, Tectonic discrimination of granitoids: Geological Society of
America Bulletin, v. 101, p. 635643, doi: 10.1130/
0016-7606(1989)101<0635:TDOG>2.3.CO;2.
Martignole, J., and Martelat, J.-E., 2003, Regional-scale
Grenvillian-age UHT metamorphism in the MollendoCamana Block (basement of the Peruvian Andes):
Journal of Metamorphic Geology, v. 21, p. 99120,
doi: 10.1046/j.1525-1314.2003.00417.x.
Meert, J.G., and Torsvik, T.H., 2003, The making and
unmaking of a supercontinent: Rodinia revisited: Tectonophysics, v. 375, p. 261288, doi: 10.1016/S00401951(03)00342-1.
Mgard, F., 1967, Commentaire dune coupe schmatique
a travers les Andes centrales du Prou: Revue de
Gographie Physique et de Gologie Dynamique, v. 9,
p. 335345.
Mgard, F., 1973, tude gologique dune transversale des
Andes au niveau du Prou Central [Ph.D. thesis]: Universit de Montpellier, 264 p.
Mgard, F., 1978, tude gologique des Andes du Prou
central: Contribution ltude gologique des Andes:
Mmoires de lORSTOM, v. 86, 310 p.
Mgard, F., Dalmayrac, B., Laubacher, G., Marocco, R.,
Martinez, C., Paredes, J., and Tomasi, P., 1971, La
chane hercynienne au Prou et en Bolivie: Premiers
rsultats: Cahiers de lORSTOM, srie gologie 3, 5 p.
Mikovi, A., Schaltegger, U., and Chew, D., 2005, Carboniferous plutonism along the Eastern Peruvian Cordillera:
Implications for the Late Paleozoic to early Mesozoic
Gondwanan tectonics: Barcelona, 6th International
Symposium on Andean Geodynamics, IRD Editions
(Institut de Recherche pour le Dveloppement, Paris),
p. 508511.
Mikovi, A., Schaltegger, U., and Koler, J., 2007,
Interrogating a paleo-cratonic marginThe Peruvian
Cordillera Oriental batholiths: Cologne, Germany,
Geochimica et Cosmochimica Acta, v. 71, Supplement,
Abstract 672, 17th Annual Goldschmidt Conference.
Miyashiro, A., Shido, F., and Ewing, M., 1970, Crystallization and differentiation in abyssal tholeiites and
gabbros from mid-oceanic ridges: Earth and Planetary Science Letters, v. 7, p. 361365, doi: 10.1016/
0012-821X(69)90050-8.

Mourier, T., Bengtson, P., Bonhomme, M., Buge, E.,


Cappetta, H., Crochet, J.-Y., Feist, M., Hirsch, K.,
Jaillard, E., Laubacher, G., Lefranc, J.-P., Moullade,
M., Noblet, C., Pons, D., Rey, J., Sigt, B., Tambareau,
Y., and Taquet, P., 1988, The Upper CretaceousLower
Tertiary marine to continental transition in the Bagua
basin, Northern Peru: Newsletters on Stratigraphy,
v. 19, p. 143177.
Mpodozis, C., and Kay, S.M., 1992, Late Paleozoic to Triassic evolution of the Gondwana margin; evidence
from Chilean Frontal Cordilleran batholiths (28S
to 31S): Geological Society of America Bulletin,
v. 104, p. 9991014, doi: 10.1130/0016-7606(1992)104
<0999:LPTTEO>2.3.CO;2.
Mukasa, S.B., 1986, Zircon U-Pb ages of super-units in
the Coastal Batholith, Peru: Implications for magmatic and tectonic processes: Geological Society of
America Bulletin, v. 97, p. 241254, doi: 10.1130/
0016-7606(1986)97<241:ZUAOSI>2.0.CO;2.
Mukasa, S.B., and Henry, D.J., 1990, The San Nicols
batholith of coastal Peru: Early Palaeozoic continental arc or continental rift magmatism?: The Geological Society of London, v. 147, p. 2739, doi: 10.1144/
gsjgs.147.1.0027.
Mukasa, S.B., and Tilton, G.R., 1985, Pb isotope systematics
as a guide to crustal involvement in the generation of
the Coastal Batholith, Peru, in Pitcher, W.S., Atherton,
M.P., Cobbing, E.J., and Beckinsale, R.D., eds., Magmatism at a plate edge: The Peruvian Andes: London,
Blackie and Son, p. 235238.
Nasdala, L., Lengauer, C.L., Hanchar, J.M., Kronz, A.,
Wirth, R., Blanc, P., Kennedy, A.K., and SeydouxGuillaume, A.-M., 2002, Annealing radiation damage
and the recovery of cathodoluminescence: Chemical
Geology, v. 191, p. 121140, doi: 10.1016/S0009-2541
(02)00152-3.
Noble and McKee, 1999, The Miocene metallogenic belt of
central and northern Peru, in Skinner, B.J., ed., Geology and ore deposits of the Central Andes: Society
of Economic Geologists, Special Publication, v. 7,
p. 153193.
Norrish, K., and Chappel, B.W., 1967, X-ray fluorescence
spectrography, in Zussman, J., ed., Physical methods in
determinative mineralogy: New York, Academic Press,
p. 161214.
Omarini, R., Sureda, R.J., Gotze, H.J., Seilacher, A., and
Pfluger, F., 1999, Puncoviscana folded belt in northwestern Argentina: Testimony of Late Proterozoic
Rodinia fragmentation and pre-Gondwana collisional
episodes: International Journal of Earth Sciences,
v. 88, p. 7697, doi: 10.1007/s005310050247.
Pankhurst, R.J., and Rapela, C.W., 1998, The proto-Andean
margin of Gondwana: The Geological Society of London, Special Publication, v. 142, 383 p.
Pankhurst, R.J., Hole, M.J., and Brook, M., 1988, Isotope
evidence for the origin of Andean granites: Transactions of the Royal Society of Edinburgh, Earth Sciences, v. 79, p. 123133.
Pankhurst, R.J., Riley, T.R., Fanning, C.M., and Kelley,
S.R., 2000, Episodic volcanism in Patagonia and the
Antarctic Peninsula: Chronology of magmatism associated with the break-up of Gondwana: Journal of
Petrology, v. 41, p. 605625, doi: 10.1093/petrology/
41.5.605.
Patio Douce, A.E., and Beard, J.S., 1995, Dehydrationmelting of biotite gneiss and quartz amphibolite from
3 to 15 kbar: Journal of Petrology, v. 36, p. 707738.
Peacock, M.A., 1931, Classification of igneous rock series:
Journal of Geology, v. 39, p. 5467.
Pearce, J.A., Harris, N.B.W., and Tindle, A.G., 1984, Trace
element discrimination diagrams for the tectonic interpretation of granitic rocks: Journal of Petrology, v. 25,
p. 956983.
Petersen, U., 1999, Magmatic and metallogenic evolution of
the Central Andes, in Skinner, B.J., ed., Geology and
ore deposits of the Central Andes: Society of Economic
Geologists Special Publication, v. 7, p. 109153.
Petford, N., Atherton, M.P., and Halliday, A.N., 1996, Rapid
magma production rates, underplating and remelting in the Andes: Isotopic evidence from northerncentral Peru (911 S): Journal of South American
Earth Sciences, v. 9, p. 6978, doi: 10.1016/0895-9811
(96)00028-4.

Geological Society of America Bulletin, September/October 2009

1323

Downloaded from gsabulletin.gsapubs.org on January 20, 2011


Mikovi et al.
Pitcher, W.S., 1997, The nature and origin of granite, 2nd
ed.: London, Weinheim, New York, Tokyo, Melbourne,
Madras, Chapman and Hall, 387 p.
Polliand, M., Schaltegger, U., Frank, M., and Fontbot,
L., 2005, Formation of intra-arc volcanosedimentary basins in the western flank of the central
Peruvian Andes during Late Cretaceous oblique
subduction: Field evidence and constraints from
U-Pb ages and Hf isotopes: International Journal
of Earth Sciences, v. 94, p. 231242, doi: 10.1007/
s00531-005-0464-5.
Powell, C.McA., Li, Z.X., McElhinny, M.W., Meert, J.G., and
Park, J.K., 1993, Paleomagnetic constraints on timing of
the Neoproterozoic breakup of Rodinia and the Cambrian formation of Gondwana: Geology, v. 21, p. 889
892, doi: 10.1130/0091-7613(1993)021<0889:PCOTOT>
2.3.CO;2.
Rakotosolofo, N.A., Tait, J.A., Carlotto, V., and Crdenas, J.,
2006, Palaeomagnetic results from the Early Permian
Copacabana Group, southern Peru: Implication for
Pangaea palaeogeography: Tectonophysics, v. 413,
p. 287299, doi: 10.1016/j.tecto.2005.10.043.
Ramos, V.A., 1988, Late Proterozoic-early Paleozoic of
South America: A collisional history: Episodes, v. 11,
p. 168174
Ramos, V.A., 2004, Cuyania, an exotic block to Gondwana:
Review of a historical success and the present problems: Gondwana Research, v. 7, p. 10091026, doi:
10.1016/S1342-937X(05)71081-9.
Ramos, V.A., 2008, The basement of the Central Andes:
The Arequipa and related terranes: Annual Review of
Earth and Planetary Sciences, v. 36, doi: 10.1146/
annurev.earth.36.031207.124304.
Ramos, V.A., and Aleman, A., 2000, Tectonic evolution of
the Andes, in Cordani, U., Milani, E.J., Thomaz Filho,
A., and Campos Neto, M.C., eds., Tectonic evolution
of South America: Rio de Janeiro, 31st International
Geological Congress, p. 635685.
Ramos, V.A., and Kay, S.M., 1991, Triassic rifting and associated basalts in the Cuyo basin, central Argentina, in
Harmon, R.S., and Rapela, C.W., eds., Andean magmatism and its tectonic setting: Geological Society of
America, Special Paper, v. 265, p. 113137.
Ramos, V.A., Jordan, T.E., Allmendinger, R.W., Mpodozis,
C., Kay, S.M., Corts, J.M., and Palma, M., 1986,
Paleozoic terranes of the Central Argentine-Chilean
Andes: Tectonics, v. 5, p. 855880, doi: 10.1029/
TC005i006p00855.
Rapalini, A.E., 1998, Syntectonic magnetization of the
mid-Palaeozoic Sierra Grande Formation: Further constraints on the tectonic evolution of Patagonia: Journal of the Geological Society, v. 155, p. 105114, doi:
10.1144/gsjgs.155.1.0105.
Rapela, C.W., Pankhurst, R.J., Casquet, C., Baldo, E.,
Saavedra, J., and Galindo, C., 1998, Early evolution of the Proto-Andean margin of South America: Geology, v. 26, p. 707710, doi: 10.1130/
0091-7613(1998)026<0707:EEOTPA>2.3.CO;2.
Rapp, R.P., and Watson, E.B., 1995, Dehydration melting of
metabasalt at 832 kbar: Implications for continental
growth and crust-mantle recycling: Journal of Petrology, v. 36, p. 891931.
Rast, N., 1984, The Alleghenian orogeny in eastern North
America: The Geological Society of London, Special
Publication, v. 14, p. 197217.
Restrepo-Pace, P.A., Ruiz, J., Gehrels, G., and Cosca,
M., 1997, Geochronology and Nd isotopic data of
Grenville-age rocks in the Colombian Andes: New
constraints for Late Proterozoic-Early Paleozoic paleocontinental reconstructions of the Americas: Earth and
Planetary Science Letters, v. 150, p. 427441, doi:
10.1016/S0012-821X(97)00091-5.
Reynolds, R.C., 1967, Estimation of mass absorption coefficient: The American Mineralogist, v. 58, p. 10691072.
Rizzotto, G.J., 1999, Petrologia e Ambiente Tectnico do
Grupo Nova Brasilndia-Rondnia [Masters thesis]:
Porto Alegre, Universidade Federal do Rio Grande do
Sul, 136 p.
Romeuf, N., Aguirre, L., Soler, P., Fraud, G., Jaillard, E.,
and Ruffet, G., 1995, Middle Jurassic volcanism in the
Northern and Central Andes: Revisita Geolgica de
Chile, v. 22, p. 245259.

1324

Roperch, P., and Carlier, G., 1992, Paleomagnetism of Mesozoic rocks from the central Andes of southern Peru
Importance of rotations in the development of the
Bolivian orocline: Journal of Geophysical Research,
v. 97, p. 17,23317,249, doi: 10.1029/92JB01291.
Rosas, S., Fontbot, L., and Tankard, A., 2007, Tectonic evolution and paleogeography of the Mesozoic Pucar Basin,
central Peru: Journal of South American Earth Sciences,
v. 24, p. 124, doi: 10.1016/j.jsames.2007.03.002.
Sadowski, G.R., 2002, The fit between Amazonia, Baltica and
Laurentia during the Mesoproterozoic assemblage of
the supercontinent Rodinia: Gondwana Research, v. 5,
p. 101107, doi: 10.1016/S1342-937X(05)70894-7.
Sadowski, G.R., and Bettencourt, J.S., 1996, Mesoproterozoic tectonic correlations between eastern Laurentia
and the western border of the Amazon Craton: Precambrian Research, v. 76, p. 213227, doi: 10.1016/
0301-9268(95)00026-7.
Snchez, A., 1983, Nuevo datos K-Ar en algunas rocas del
Per: Boletn de la Sociedad Geolgical del Per,
v. 71, p. 193202.
Snchez, A., 1995, Geologa de los cuadrngulos de Bagua
Grande, Jumbilla, Lonya Grande, Chachapoyas, Rioja,
Leimebamba y Bolvar (hojas 12-g, 12-h, 13-g, 13-h,
13-i, 14-h, 15-h): Lima, Instituto Geolgico Minero
y Metalrgico, Boletn Serie A, Carta Geolgica
Nacional, v. 56, 287 p.
Snchez Zavala, J.L., Centeno Garca, E., and Ortega
Gutirrez, F., 1999, Review of Paleozoic stratigraphy
of Mxico and its role in the Gondwana-Laurentia connections: Geological Society of America Special Paper,
v. 336, p. 211226.
Schreiber, D.W., 1989, Zur Genese von Goldquarzgngen der Pataz-Region im Rahmen der geologischen
Entwicklung der Ostkordillere Nordperus (unter
besonderer Bercksichtigung der Distrikte Parcoy,
La Lima und Buldibuyo): Heidelberger Geowissenschaftliche Abhandlungen, v. 29, 235 p.
Schreiber, D.W., Fontbot, L., and Lochmann, D., 1990,
Geologic setting, paragenesis, and physicochemistry of
gold quartz veins hosted by plutonic rocks in the Pataz
region: Economic Geology and the Bulletin of the
Society of Economic Geologists, v. 85, p. 13281347.
Sempere, T., 1995, Phanerozoic evolution of Bolivia and
adjacent regions, in Tankard, A.J., Suarez, R., and
Welsink, H.J., eds., Petroleum basins of South America: American Association of Petroleum Geologists
Memoir, v. 62, p. 207230.
Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto,
V., Jacay, J., Arispe, O., Nraudeau, D., Crdenas,
J., Rosas, S., and Jimnez, N., 2002, Late Permian
Middle Jurassic lithospheric thinning in Peru and
Bolivia, and its bearing on Andean tectonics: Tectonophysics, v. 345, p. 153181, doi: 10.1016/S0040-1951
(01)00211-6.
Shackleton, R.M., Ries, A.C., Coward, M.P., and Cobbold,
P.R., 1979, Structure, metamorphism and geochronology of the Arequipa Massif of coastal Peru: Journal
of the Geological Society, v. 136, p. 195214, doi:
10.1144/gsjgs.136.2.0195.
Sisson, T.W., Ratajeski, K., Hankins, W.B., and Glazner,
A.F., 2005, Voluminous granitic magma from common basaltic sources: Contributions to Mineralogy
and Petrology, v. 148, p. 635661, doi: 10.1007/
s00410-004-0632-9.
Skjerlie, K.P., Patio Douce, A.E., and Johnston, A.D., 1993,
Fluid absent melting of a layered crustal protolith:
Implications for the generation of anatectic granites:
Contributions to Mineralogy and Petrology, v. 114,
p. 365378, doi: 10.1007/BF01046539.
Soler, P., 1991, Contribution ltude du magmatisme
associe aux zones de subduction. Ptrographie, gochimie et gochimie isotopique des roches intrusives
sur un transect des Andes du Prou Central. Implications godynamiques et metallogeniques [Ph.D. thesis]:
Universit de Paris VI, 950 p.
Spikings, R.A., Winkler, W., Hughes, R.A., and Handler,
R., 2005, Thermochronology of allochthonous terranes in Ecuador: Unravelling the accretionary and
post-accretionary history of the Northern Andes: Tectonophysics, v. 399, p. 195220, doi: 10.1016/j.tecto.
2004.12.023.

Streckeisen, A., 1974, Classification and nomenclature of


plutonic rocks recommendations of the International
Union of Geological Sciences subcommission on the
systematics of igneous rocks: International Journal of
Earth Sciences, v. 63, p. 773786.
Tassinari, C.C.G., and Macambira, M.J.B., 1999, Geochronological provinces of the Amazonian craton: Episodes, v. 22, p. 174182.
Tassinari, C.C.G., Bettencourt, J.S., Geraldes, M.C.,
Macambira, M.J.B., and Lafon, J.M., 2000, The
Amazon craton, in Cordani, U., Milani, E.J., Thomaz
Filho, A., and Campos Neto, M.C., eds., Tectonic evolution of South America: 31st International Geological
Congress, Rio de Janeiro, p. 4195.
Tollo, R.P., Aleinikoff, J.N., Bartholomew, M.J., and Rankin,
D.W., 2004, Neoproterozoic A-type granitoids of the
central and southern Appalachians: Intraplate magmatism associated with episodic rifting of the Rodinian
supercontinent: Precambrian Research, v. 128, p. 338,
doi: 10.1016/j.precamres.2003.08.007.
Tosdal, R.M., 1996, The Amazon-Laurentian connection as
viewed from the Middle Proterozoic rocks in the central Andes, western Bolivia and northern Chile: Tectonics, v. 15, p. 827842, doi: 10.1029/95TC03248.
Turner, S.P., Platt, J.P., George, R.M.M., Kelly, S.P., Pearson,
D.G., and Nowell, G.M., 1999, Magmatism associated
with orogenic collapse of the Betic-Alboran Domain,
SE Spain: Journal of Petrology, v. 40, p. 10111036,
doi: 10.1093/petrology/40.6.1011.
Veevers, J.J., Conaghan, P.J., Powell, C.M., Cowan, E.J.,
McDonnell, K.L., and Shaw, S.E., 1994, in Veevers,
J.J., and Powel, C.M., eds., Permian-Triassic Pangean
basins and foldbelts along the Panthalassan margin of
Gondwana: Geological Society of America Memoir
184, p. 11171.
Vos, I.M.A., Bierlein, F.P., and Heithersay, P.S., 2007, A crucial role for slab break-off in the generation of major
mineral deposits: Insights from central and eastern
Australia: Mineralium Deposita, v. 42, p. 515522,
doi: 10.1007/s00126-007-0137-3.
Walawender, M.J., Gastel, R.G., Clinkenbeard, J.P.,
McCormick, W.V., Eastman, B.G., Wernicke, R.S.,
Wardlaw, M.S., Gunn, S.H., and Smith, B.M., 1990,
Origin and evolution of the zoned La Posta-type plutons, eastern Peninsular Ranges Batholith, southern
and Baja California, in Anderson, J.L., ed., The nature
and origin of Cordilleran magmatism: Geological
Society of America Memoir 174, p. 118.
Watson, E.B., 1996, Dissolution, growth and survival of zircons during crustal fusion: Kinetic principles, geological models and implications for isotopic inheritance:
Geological Society of America Special Paper 315,
p. 4356.
Wiedenbeck, M., All, P., Corfu, F., Griffin, W.L., Meier,
M., Oberli, F., Von Quadt, A., Roddick, J.C., and
Spiegel, W., 1995, Three nature zircon standards for
U-Th-Pb, Lu-Hf, trace element and REE analyses:
Geostandards Newsletter, v. 19, p. 123, doi: 10.1111/
j.1751-908X.1995.tb00147.x.
Wong, S.Y.M., Ton, A., and Wortel, M.J.R., 1997, Slab
detachment in continental collision zones: An analysis
of controlling parameters: Geophysical Research Letters, v. 24, p. 20952098, doi: 10.1029/97GL01939.
Xu, H., Ma, C., and Ye, K., 2007, Early cretaceous granitoids
and their implications for the collapse of the Dabie orogen, eastern China: SHRIMP zircon UPb dating and
geochemistry: Chemical Geology, v. 240, p. 238259,
doi: 10.1016/j.chemgeo.2007.02.018.
Zeil, W., 1983, Das prkambrische Basement der Anden:
Ein berblick, in Miller, H., and Rosenfeld, U., eds.,
Geowissenschaftliches
Lateinamerika-Kolloquium:
Teil I, Zentralblatt fr Geologie und Palontologie,
p. 246254.
Zen, E-an, 1988, Phase relations of peraluminous granitic
rocks and their petrogenetic implications: Annual
Review of Earth and Planetary Sciences, v. 16,
p. 2151.
MANUSCRIPT RECEIVED 17 JUNE 2008
REVISED MANUSCRIPT RECEIVED 10 NOVEMBER 2008
MANUSCRIPT ACCEPTED 12 NOVEMBER 2008
Printed in the USA

Geological Society of America Bulletin, September/October 2009

You might also like