Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Gondwana Research 35 (2016) 5978

Contents lists available at ScienceDirect

Gondwana Research

journal homepage: www.elsevier.com/locate/gr

Proto-Andean evolution of the Eastern Cordillera of Peru


David M. Chew a,, Giovanni Pedemonte b, Eoghan Corbett a
a
Department of Geology, School of Natural Sciences, Trinity College Dublin, Dublin 2, Ireland
b
Escuela de Geologa, Universidad Nacional Mayor de San Marcos, Lima, Peru

a r t i c l e i n f o a b s t r a c t

Article history: The Eastern Cordillera of Peru represents one of the longest (N1200 km) Paleozoic metamorphic and magmatic
Received 20 October 2015 belts exposed along the western Andean margin of South America. In this study, we examine the tectonothermal
Received in revised form 16 March 2016 evolution of a key segment of the metasedimentary basement of the Eastern Cordillera of Peru (the
Accepted 18 March 2016
Huaytapallana Complex) and demonstrate that it has experienced a hitherto undocumented high-grade orogenic
Available online 6 May 2016
event at 260 Ma (latest Middle Permian) based on UPb and ThPb monazite age data from paragneisses and U
Handling Editor: R.D. Nance Pb dating of zircon rims from leucosomes. These ages are interpreted as recording crystallization and are consis-
tent with 255 Ma rutile growth in lower-grade units. UPb apatite data (c. 260230 Ma) in all units are consistent
Keywords: with slow cooling from this 260 Ma metamorphic peak. UPb zircon geochronology of pre-tectonic plutons yield
Proto-Andes ages ranging from c. 302 Ma to c. 260 Ma. These geochronological data are augmented by new UPb apatite age
Peru data from other segments along the Eastern Cordillera of Peru. A regional synthesis of existing geochronological
Eastern cordillera constraints from the Eastern Cordillera of Peru demonstrates that the margin has experienced a polycyclic oro-
UPb geochronology genic history. Deformation and magmatism occurred at c. 480 Ma and c. 435 Ma during the Famatinian orogenic
Orogeny
cycle, was followed by a Late Silurian to Early Carboniferous (c. 420350 Ma) magmatic and metamorphic gap,
and terminated with Gondwanide magmatism and metamorphism at c. 315 Ma and c. 260 Ma. These Famatinian
and Gondwanide orogenic phases can be correlated into the Proto-Andean margin of Argentina and Chile and are
thus of regional extent. The evolution of the Proto-Andean margin is thus best explained by changes in tectonic
plate reorganization in a long-lived Paleozoic accretionary orogen which was undergoing phases of advance and
retreat, resulting in magmatic pulses and orogenic phases which can be correlated along the length of the plate
boundary.
2016 International Association for Gondwana Research. Published by Elsevier B.V. All rights reserved.

1. Introduction be difcult to identify in metasedimentary basement rocks where


lithostratigraphic correlation is difcult.
The Eastern Cordillera of Peru represents one of the longest The Eastern Cordillera of Peru consists of a series of Paleozoic
(N 1200 km) Paleozoic metamorphic and magmatic belts exposed on metasedimentary rocks, termed the Maran Complex north of 11S,
the western Andean margin of South America (Fig. 1). Whereas Andean and the Huaytapallana Complex south of 11S (Mgard, 1978; Pffner
deformation and magmatism have been extensively studied (e.g. and Gonzalez, 2013; Fig.1). The metasedimentary basement is intruded
Jaillard and Soler, 1996; Ramos, 2009; Pffner and Gonzalez, 2013), by several intrusive suites, including a series of OrdovicianEarly
the Paleozoic evolution of much of the Proto-Andean margin is more Silurian granitoids, CarboniferousEarly Permian I-type granitoids and
complex (e.g. Pankhurst et al., 1998; Thomas et al., 2004; Chew et al., diorites, and Late PermianTriassic granodiorites and S-type granites
2007; Miskovic et al., 2009) and more difcult to unravel. This is (Fig. 1, Miskovic et al., 2009). Ordovician plutonism is also observed
because even in regions such as the Eastern Cordillera of Peru where on the coast in the Arequipa basement (Fig. 1), a Proterozoic crustal
pre-Andean basement rocks are not buried by recent volcanism or block that experienced 0.91.2 Ga Grenville metamorphism (Loewy
foreland basin sediments, it is difcult to correlate orogenic phases in et al., 2004; Wasteneys et al., 1995), and in the offshore continuation
the basement rocks without comprehensive geochronological studies. of this crustal block on the outer shelf rise (Isla de las Hormigas Afuera,
This problem is exacerbated by the polycyclic orogenic history of the Fig. 1; Romero et al., 2013).
Paleozoic Proto-Andean margin (the Andes represent the locus of The timing of deformation in the metasedimentary basement of the
continued plate convergence through much of the Phanerozoic), Eastern Cordillera of Peru was originally regarded as Neoproterozoic in
and even the effects of Andean (EoceneOligocene) tectonism can age based on bulk, unabraded UPb zircon data (ca. 630610 Ma lower
intercepts) from granulitic gneisses in central Peru (Dalmayrac et al.,
1980). Recent studies demonstrate that it is Paleozoic in age and
Corresponding author. comprises at least two orogenic cycles, an Ordovician magmatic and

http://dx.doi.org/10.1016/j.gr.2016.03.016
1342-937X/ 2016 International Association for Gondwana Research. Published by Elsevier B.V. All rights reserved.
60 D.M. Chew et al. / Gondwana Research 35 (2016) 5978

80 W 76 W 72 W
0
Quito Cordillera Real

ECUADOR

ote
G uam
Late Cretaceous Salado
N accreted oceanic
terranes Amazonic

o
craton

Ala
Golfo de Loja
Guayquil

A
4 S
PT P
Amotape
PT
Precordillera
P
ba
m
ba
n ca ion C PERU
a ct
Hu efle
D

8 S C
Maran Complex
OS J
BRAZIL

P
Ou

Paracas
ter

Terrane
Sh
PA

elf
OC

PT
CI

Hi
gh
FI
EA

Huaytapallana Complex
OS
C

PT
N

Isla de OS
12 S las Fig. 2 E
A
Lima ST
Hormiga Huancayo ERN
Legend COR
J Early Jurassic quartz syentites DIL
LE
? RA
T Late Triassic Carabaya granitoids T C
PT PT
Permo-Triassic granodiorites/ OS T
PT S-type granites - monzogranites
C Carboniferous - Permian rhyodacites/
I-type diorites - granitoids J
OS Ordovician - Early Silurian granitoids OS Proterozoic
gneisses of the
Arequipa -
P Undifferentiated Palaeozoic metasediments Antofalla Basement
Precambrian gneisses /
Arequipa - Antofalla Basement
OS
16 S
Arequipa
0 100 200km

Fig. 1. (A) Map of South America illustrating the major tectonic provinces and the ages of their most recent metamorphic events from Chew et al. (2007, 2008), adapted from Cordani et al.
(2000). Precambrian and Paleozoic inliers in the Andean belt are shown in black and light gray, respectively. (B) Geological map of Peru and Ecuador illustrating the major Paleozoic meta-
morphic and magmatic belts along with the Proterozoic gneisses of the ArequipaAntofalla block. Adapted after Chew et al. (2007) using the geology of Litherland et al. (1994) and Leon
et al. (2000), the ages of plutonism from Miskovic et al. (2009) and the location of the outer shelf high from Romero et al. (2013). The black boxes highlighted on the Eastern Cordillera
illustrate the sampling localities in this study.

metamorphic event (Cardona, 2006; Chew et al., 2007; Cardona et al., correlated (Chew et al., 2007; Cardona et al., 2009; Ramos, 2009) with
2009; Willner et al., 2014), and a Carboniferous orogenic phase which the Early to Middle Ordovicianage Famatinian metamorphism and
is spatially associated with regional arc magmatism (Cardona, 2006; subduction-related magmatism and a Carboniferous (Early Gondwanide)
Chew et al., 2007; Cardona et al., 2009). These events have been event that affected much of the Proto-Andean margin of South America
D.M. Chew et al. / Gondwana Research 35 (2016) 5978 61

(Pankhurst et al., 1998; Thomas et al., 2004). In this study, we examine Wiley table, heavy liquids, and magnetic separation) at the Depart-
the tectonothermal evolution of a key segment of the metasedimentary ment of Geology, Trinity College Dublin (TCD). The fraction between
basement of the Eastern Cordillera of Peru (the Huaytapallana Complex 125 and 250 m was used in the majority of cases. U-bearing accessory
east of Huancayo, Figs. 1,2) and demonstrate that it has experienced mineral phases were picked under a binocular microscope in ethanol,
a hitherto undocumented high-grade orogenic event at c. 260 Ma mounted in 2.5 cm epoxy mounts, and polished to approximately half
(Middle Permian). These geochronological data are augmented by thickness prior to analysis. Cathodoluminescence (CL) images of zircon
new UPb accessory mineral age data from other segments along the grains were acquired at the Centre for Microscopy and Analysis at TCD
Eastern Cordillera of Peru (Fig. 1). The geological evolution of key using a Tescan MiraXMU Scanning Electron Microscope with a KE
sectors of the Eastern Cordillera is then synthesized by incorporating Developments Centaurus Cathodoluminescence Detector.
data from this work and other recent studies, which are used to make
a tectonic model for the Paleozoic evolution of this segment of the 3.1. UPb LA-ICPMS analyses
Proto-Andean margin.
All UPb data were acquired using a Photon Machines Analyte Exite
2. Geological setting and samples 193 nm ArF Excimer laser-ablation system with a Helex 2-volume abla-
tion cell coupled to a Thermo Scientic iCAP Qc at TCD. 0.65 l/min He
Samples for UPb accessory mineral (apatite, rutile, zircon, and mon- carrier gas was split evenly between the large outer sample chamber
azite) geochronology were collected mainly from metasedimentary and the small inner volume (the cup) where ablation occurs. A small
and intrusive rocks in the Huaytapallana Complex east of Huancayo volume of N2 (ca. 6 ml/min) to enhance signal sensitivity and reduce
(Figs. 1, 2). Additionally, apatite UPb geochronology was undertaken oxide formation and 0.7 l/min Ar nebulizer gas was then introduced to
on several samples from the study of Chew et al. (2007) to characterize the sample-gas mixture via an in-house smoothing device. 202Hg,
204,206,207,208
the late orogenic cooling history of key segments of the Eastern Pb, 232Th, and 238U were acquired for all UPb analyses.
91
Cordillera to the north (sampling areas denoted by black boxes on Zr, 43Ca, 48Ti, and 31P were employed as the internal standards for
Fig. 1). South of 11S, the metamorphic basement inliers of the East- zircon, apatite, rutile, and monazite, respectively. Zircon analyses
ern Cordillera have been divided into a series of massifs by Mgard employed either a 30 m laser spot, a 4 Hz laser repetition rate, and a
(1978) based primarily on their metamorphic grade. These inliers uence of 3.9 J/cm2 for detrital analyses, or a 14 m laser spot, a 5 Hz
(the ChupnHuasahuasi, MarayniocMarairazo, Huaytapallana, laser repetition rate, and a uence of 3.9 J/cm2 for dating thin zircon
and Huaytapallanakaru massifs, Fig. 10) were considered to be Precam- rim overgrowths in high-grade paragneiss samples (e.g. TCH-002,
brian in age by Mgard (1978). Although we view these massifs as prob- -003, and -007). Apatite and rutile analyses were acquired using either
able along-strike equivalents of the Maran Complex to the north, they a 30, 47, or 60 m laser spot (kept constant for a given session), a 5 Hz
are grouped here into one unit termed the Huaytapallana Complex. laser repetition rate, and a uence of 3.9 J/cm2. Monazite analyses
Three sampling transects (Figs. 2BD) were undertaken in the main were acquired using a 14 m laser spot, a 4 Hz laser repetition rate,
study area (Fig. 2A) and sample descriptions are listed in Appendix A. and a uence of 3.9 J/cm2.
The highest-grade rocks occur in the Huaytapallana Massif (Fig. 2C) of The raw isotope data were reduced using the Vizual Age data re-
Mgard (1978) and comprise high-grade garnetsillimanitebiotite duction scheme of Petrus & Kamber (2012) within the freeware
bearing paragneisses with abundant development of leucosomes IOLITE package of Paton et al. (2011). User-dened time intervals are
(Figs. 3A-C). A small foliated metadiorite body was observed cutting established for the baseline correction procedure to calculate session-
the paragneisses (Fig. 3D, sample TCH-004). Lower-grade rocks occur wide baseline-corrected values for each isotope. The time-resolved frac-
to the north in the southernmost portion of the MarayniocMarairazo tionation response of individual standard analyses is then characterized
massif (Fig. 2B) of Mgard (1978). This unit comprises albite-bearing using a user-specied down-hole correction model (such as an expo-
and chlorite-bearing mica schists with minor psammites and amphibo- nential curve, a linear t, or a smoothed cubic spline). The data reduc-
lites. Its contact with the Muchac Granodiorite (Fig. 3E) to the west is a tion scheme then ts this appropriate session-wide model UThPb
thrust fault (Fig. 2B). The southern sampling transect (Fig. 2D) was fractionation curve to the time-resolved standard data and the un-
undertaken in the northernmost portion of the Huaytapallanakaru knowns. Sample-standard bracketing is applied after the correction of
Massif of Mgard (1978). It comprises biotite-grade mica schists and down-hole fractionation to account for long-term drift in isotopic or el-
psammites cut by both the Villa Azul Batholith (Fig. 2D, Fig. 3F) and emental ratios by normalizing all ratios to those of the UPb reference
minor metadolerite intrusions (Fig. 3G, 3H). standards. Common Pb in the apatite and rutile standards was corrected
To the west of the Eastern Cordillera in this region, a succession of using the 207Pb-based correction method using a modied version of
Paleozoic sedimentary sequences crop out in semi-grabens in the the VizualAge DRS that accounts for the presence of variable common
Mantaro Valley, which are overlain by remnants of the TriassicJurassic Pb in the primary standard materials (Chew et al. 2014).
carbonate platform, the Pucar Group. To the east, Paleozoic units crop
out and comprise continental clastic sequences of the Devonian 3.2. UPb accessory mineral standards
Cabanillas Group, the continental and shelf sequences of the Mississip-
pian Ambo Group, the marine Upper Carboniferous Tarma Group, a con- For zircon, the analytical procedure utilized repeated blocks of four
densed PennsylvanianPermian carbonate platform of the Copacabana zircon standards followed by 20 unknown samples. The four zircon
Group, and continental red beds and volcaniclastics of the Triassic standard analyses comprised two analyses of the primary standard
Mitu Group (Fig. 4). The Villa Azul Batholith strikes NW-SE with a strike 91500 zircon (206Pb/238U TIMS age of 1065.4 0.6 Ma; Wiedenbeck
length of more than 600 km and an average width of 15 km (Cerrn and et al., 1995), followed by one analysis of Pleovice zircon (weighted
Ticona, 2003) and is mainly emplaced into metamorphic rocks of the mean 206Pb/238U TIMS age of 337.13 0.37 Ma; Slma et al. 2008)
Huaytapallana Complex. The sole existing geochronological constraint and Temora 2 zircon (206Pb/238U TIMS age of 416.8 1.3 Ma; Black
from the region is a KAr biotite age of 251 Ma from the Villa Azul et al. 2003). Pleovice and Temora 2 yielded LA-ICPMS UPb concordia
Batholith near Villa Azul Ranch (Stewart et al. 1974). ages of 337.8 1.3 Ma (MSWD = 5.1; n = 26) and 417.2 1.4 Ma
(MSWD = 0.107; n = 33) in this study.
3. Analytical technique For apatite and rutile, blocks of six standards and two NIST612 stan-
dard glass analyses were followed by 20 unknown samples. Madagascar
Zircon, apatite, rutile, and monazite grains were separated using apatite (Thomson et al. 2012) (weighted average ID-TIMS concordia age
standard separation and concentration procedures (jaw crushing, of 473.5 0.7 Ma, Cochrane et al. 2014) was used as the primary apatite
62
B

A C

D.M. Chew et al. / Gondwana Research 35 (2016) 5978


D

Fig. 2. (A) Regional geology of the Huaytapallana Complex northeast of the town of Huancayo, modied after Paredes (1994) and Cerrn & Ticona (2003). (BD) Large-scale geology maps with the sampling localities in this study depicted. The names
of the metasedimentary basement massifs of the Eastern Cordillera are denoted in blue text and follow Mgard (1978).
D.M. Chew et al. / Gondwana Research 35 (2016) 5978 63

A B

C D

E F
Enclaves

G H

Porphyritic Volcanic (FW2-006)

Huaytapallana Complex
Schists

Fig. 3. (AD) Field photographs of the high-grade metamorphic rocks in the study area. (A) Migmatitic paragneiss from Lake Lasuntay (TCH-001). (B, C) Migmatitic paragneisses from
Cerro Jacallate (BTCH-002; CTCH-003). (D) Foliated metadiorite which cuts the migmatitic paragneisses at Cerro Lasuntay (TCH-004). (EH) Field photographs of the low-grade
metamorphic rocks in the study area. (E) granite outcrop in the village of Muchac (FW2-011). (F) granite outcrop from the Villa Azul batholoth with abundant mac enclaves close to
the town of Pahual (FW2-010). (G) outcrop near the town of Ocoro where the contact relationships between a metadolerite (a porphyritic subvolcanic body, FW2-006) and schists of
the Huaytapallana Complex are evident. (H) Metadolerite outcrop near the town of Pariahuanca (FW2-003).

reference material. McClure Mountain syenite apatite (weighted mean (McDowell et al., 2005) yielded an unanchored UPb TeraWasserburg
207
Pb/235U age of 523.51 2.09 Ma, Schoene & Bowring, 2006) and concordia lower intercept age of 31.18 0.89 Ma (MSWD = 0.99;
Durango apatite (31.44 0.18 Ma, McDowell et al., 2005) were used n = 75).
as a secondary standards. McClure Mountain apatite yielded a UPb R10 rutile (UPb TIMS age of c. 1090 Ma (Luvizotto et al., 2009) was
TeraWasserburg concordia lower intercept age of 521.2 2.6 Ma used as the primary rutile standard while R19b rutile (weighted mean
(MSWD = 1.6; n = 56), anchored using a 207Pb/206Pb value of TIMS 206Pb/238U age of 489.5 0.9 Ma, Zack et al., 2011) and RZ-3 rutile
value of 0.88198 derived from an apatite ID-TIMS total UPb iso- (SIMS UPb a mean 207Pb/206Pb age of 1780.2 9.6 Ma, Shi et al., 2012)
chron (Schoene and Bowring, 2006). The Durango apatite standard were used as secondary standards. R19 rutile and RZ3 rutile yielded
64 D.M. Chew et al. / Gondwana Research 35 (2016) 5978

LITHOSTRATIGRAPHIC THICKNESS
ERA PERIOD EPOCH (m) LITHOLOGY
UNITS

PLEISTOCENE HOLOCENE
Colluvium 10 Gravels in sandy matrix, and layers of silt

QUATERNARY
CENOZOIC Alluvium 40 Sub-rounded gravels in sandy-clay matrix

Sandy claystones intercalated with


Lacustrine deposits micro-conglomerates
60
Poorly-consolidated conglomerates, intercalated
Alluvium
with thick and massive claystone beds
Disc.

Limestones, composed mainly of thickly bedded


Pucar Group 250 micrites with chert nodules, passing upwards into
UPPER (Lower units)
MESOZOIC

mudstones intercalated with red claystones and


TRIASSIC

fossilliferous limestone layers

Mitu Group 1500 Red sandstones and claystones intercalated with


felsic lava flows and volcanic breccia
MIDDLE
Disc. (= JURUA OROGENY)

Yellowish white limestones intercalated with


N

Copacabana
IA

LOWER 150 sandstones and red claystones, passing upwards


M

Group
R

into gypsum layers


PE
PALEOZOIC

Disc.
CARBONFEROUS

Dark gray siltstones intercalated with thick beds


UPPER Tarma Group 500 of medium-grained sandstones and calcareous
sandstones

Dark, fine-grained, thickly-bedded sandstones


LOWER Ambo Group 500 intercalated with gray claystones

Disc.

Cabanillas Dark, thinnly-bedded carbonaceous siltstones


DEVONIAN 1200
Group intercalated with thin beds of coarse-grained
sandstones

Disc.
Huaytapallana

Phyllites Albite-bearing and chlorite-bearing mica schists


Complex

with minor psammites and amphibolites


Schists
Pre-Devonian
sequences High-grade gamet-sillimanite-biotite bearing
Gneisses,
amphibolites paragneisses with abundant development of
leucosomes

Fig. 4. Stratigraphic column depicting the Phanerozoic cover sequences overlying the Huaytapallana Complex, adapted from Cerrn and Ticona (2003) and Rosas et al. (2007).

concordia ages of 490 12 Ma (MSWD = 0.45; n = 12) and 1783 data for all analyses are listed in Appendices BE. Representative SEM-
36 Ma (MSWD = 5.0; n = 12), respectively. 44069 monazite (TIMS CL images are presented in Fig. 6. The MSWD of the UPb concordia
UPb concordia age of 424.9 0.4 Ma, Aleinikoff et al., 2006) was ages employs both the MSWD for XY equivalence and the MSWD for
used as the primary monazite standard and A49 and A276 monazite concordance. The UPb systematics of individual samples are discussed
(TIMS UPb monazite ages of 1874.8 5.4 Ma and 1915.6 5.6 Ma, in the following section, whereas the geological signicance of the UPb
Yann Lahaye, unpublished data) were used as secondary standards. data from the Huaytapallana Complex is discussed in Section 5.1.
A49 monazite yielded a LA-ICPMS UPb concordia age of 1877
10 Ma (MSWD = 0.023) and a weighted average 208Pb232Th age of 4.1. Apatite UPb data from the Huaytapallana complex
1877 26 Ma (MSWD = 1.5). A276 monazite yielded a LA-ICPMS U
Pb concordia age of 1906.8 7.8 Ma (MSWD = 2.7) and a weighted av- Apatite UPb data were acquired from eight samples (metasedimentary
erage 208Pb232Th age of 1906 21 (MSWD = 1.9). In both cases, there is rocks and pre-tectonic intrusives) from the low-grade portions of
very good agreement between the LA-ICPMS UPb concordia ages and the the Huaytapallana Complex (Fig. 2B, 2D, Fig. 5) and for four high-
weighted average 208Pb232Th ages of the secondary monazite standards. grade paragneiss samples and one intrusive rock from the high-grade
portion of the Huaytapallana Complex (Fig. 2C, Fig. 7). In the low-
4. Results grade portions of the Huaytapallana Complex, the apatite UPb data
all lie on TeraWasserburg discordias, yielding lower intercept ages
UPb apatite, rutile, zircon, and monazite concordia diagrams are ranging from 281.9 7.8 Ma (FW2-010, Fig. 5R) to 228.1 5.1 Ma
presented in Figs. 5, 79, and summarized in Table 1. The raw UPb (FW2-007, Fig. 5Q). All eight of these TeraWasserburg discordias are
D.M. Chew et al. / Gondwana Research 35 (2016) 5978 65

unanchored, i.e. the upper intercept (the initial Pb composition) and the -006) yielding large uncertainties because individual analyses cluster
lower intercept age are determined exclusively from the spread in the close to common Pb. Three samples from the high-grade portion of
UPb data on TeraWasserburg concordia. Six of the eight samples the Huaytapallana Complex lie on TeraWasserburg discordias, yielding
(Fig. 5; FW2-002, -004, -007, -010, -014, -015) have a sufciently unanchored lower intercept ages ranging from 229.9 20 Ma (TCH-
large spread in UPb ratio that they yield lower intercept age uncer- 001, Fig. 7C) to 250 16 Ma (TCH-003, Fig. 7F). Two samples (TCH-
tainties less than 3% (Table 1), with two samples (Fig. 5; FW2-003, 002, -004) do not lie on a TeraWasserburg intercept, implying

0.188
1160
310 0.195 Pariahuanca Intrusive
0.049
1120 common Pb 0.184 FW2-002 (zircon)
1023.5 2.9 Ma
300 0.185 1080 1080
0.180 MSWD = 1.4
Pb/238U

Pb/238U
0.047

Pb/238U
0.175 1040
290 0.176
1000
206

206

206
0.045 0.165
960 0.172
280
Lampa Granodiorite 0.155 920 Pariahuanca Intrusive
0.168
FW2-001 (zircon) FW2-002 (zircon)
0.043
290.3 4.0 Ma 880 Pb loss
0.145 0.164
MSWD = 2.2 980
207 235
Pb/ U A 207
Pb/235U B 207
Pb/235U C
0.041 0.135 0.160
0.30 0.32 0.34 0.36 0.38 1.2 1.6 2.0 2.4 2.8 3.2 1.55 1.65 1.75 1.85 1.95

360
0.056 Pariahuanca Intrusive Pariahuanca Intrusive
Pariahuanca mafic intrusive 1.0 0.5
FW2-002 (apatite) FW2-002 (rutile)
FW2-003 (zircon)
257.8 2.3 Ma
0.052 274.1 1.2 Ma
320 0.8 Tera Wasserburg intercept age: 0.4 MSWD = 2.0
MSWD = 1.3
Pb/238U

257.3 5.5 Ma, MSWD = 2.1

Pb/206Pb
0.048 inherited 207
Pb/206Pb initial: 0.8793 anchored at 207Pb/206Pb
grains 0.6 0.3 initial = 0.852
206

0.044

207
0.4 4000 0.2
0.040
Pb/206Pb

240
00

750
0.2 0.1

550
E F

350
0.036 24
800

Pb loss
D
207

207 235
0.032 Pb/ U 0.0
238 206
U/ Pb 0.0
238
U/ Pb206

0.24 0.28 0.32 0.36 0.40 0.44 0 4 8 12 16 20 24 8 12 16 20 24 28 32

0.049 1.0
Pariahuanca mafic intrusive 5200
Ocoro Porphrytic intrusive 300 Ocoro Porphryitic intrusive
0.8 FW2-003 (apatite)
0.047 FW2-006 (zircon) FW2-006 (apatite)
260.1 3.9 Ma 290 0.8
Tera Wasserburg intercept age: MSWD = 2.3 Tera Wasserburg intercept age:
0.6 27818 Ma, MSWD = 2.0 0.045 229 61 Ma, MSWD = 1.3
280
Pb/238U

4400 207
Pb/206Pb initial: 0.8447 0.6 207
Pb/206Pb initial: 0.8546
inherited grains 4400
0.043 270
0.4
206

0.4
4000 0.041
260
4000
Pb/206Pb

0
Pb/206Pb

0.2 360 0 250 0


0 0.039
0.2 360 0
00

20 0
400

00

20 I
400
12

G
12

240
H
207

238
U/206Pb
207

207
0.0 0.037 Pb/235U 238
U/206Pb
0.0
0 4 8 12 16 20 24 28 0.25 0.27 0.29 0.31 0.33 0.35 0.37 0 4 8 12 16 20 24 28

1.0
30
0.23 Yerba Buena Psammite 1300 Yerba Buena Psammite Yerba Buena Psammite
FW2-004 (detrital zircon) FW2-004 (detrital zircon) FW2-004 (apatite)
0.21 25 0.8
n = 104/117 n = 104/117
Tera Wasserburg intercept age:
1100
Relative probability

0.19 20 236.11.9 MSWD = 2.7


Number
Pb/238U

0.6 207
Pb/206Pb initial: 0.8692
0.17 15
206

0.15 900 0.4 4000


10
0.13
Pb/206Pb

700
5 0.2 00
24
800

0.11
J 0
K L
207

207
0.09 Pb/235U 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 0.0
238
U/206Pb
0.6 1.0 1.4 1.8 2.2 2.6 206
Pb/238U age (Ga) 0 4 8 12 16 20 24 28

Fig. 5. UPb zircon Wetherill concordias (A, B, C, D, H, J, M, O, S), UPb detrital zircon probability density diagrams (K, N), UPb apatite TeraWasserburg concordias (E, G, I, L, Q, R, T, U) and
UPb rutile TeraWasserburg concordias (F, P) for the FW2 sample suite.
66 D.M. Chew et al. / Gondwana Research 35 (2016) 5978

0.4 0.053
30 Chuyas Psammite
1800 FW2-007 (detrital zircon) 0.051 320
25 n = 87/100

Relative probability
0.3
Pb/238U

Pb/238U
Number
1400 20 0.049
206

206
0.2 15
1000 0.047
Chuyas Psammite Pahual Granite
10
FW2-007 (detrital zircon) FW2-010 (zircon)
0.1 600 n = 87/100 0.045 302.0 3.7 Ma
5
MSWD = 2.0
0
N 280
207
Pb/235U O
0.0
207
Pb/ U235 M 0 0.5 1.0 1.5 2.0 2.5 0.043
0.30 0.32 0.34 0.36 0.38 0.40
206
0 1 2 3 4 5 6 Pb/238U age (Ga)

0.16
Chuyas Psammite Pahual Granite
0.8 FW2-007 (apatite) FW2-010 (apatite)
0.14 Chuyas Psammite 0.8
FW2-007 (rutile)
Tera Wasserburg intercept age: Tera Wasserburg intercept age:
250.9 7.4 Ma
0.12 0.6 228.1 5.5 Ma, MSWD = 2.1 281.9 7.8 Ma, MSWD = 1.06
MSWD = 1.1 0.6
Pb/206Pb

207
4400 Pb/206Pb initial: 0.8409 207
Pb/206Pb initial: 0.8112
4400
0.10 anchored at
207
Pb/206Pb 0.4 0.4
207

initial = 0.852 4000


0.08 4000

Pb/206Pb
0
Pb/206Pb

360 0 0
0.2 0.2 360 0
500

400

0 0
300

0.06 20
P
00

00

400
400
2
Q R
12

12
207
238
U/206Pb
207

238
0.04 0.0
238
U/206Pb U/206Pb
0.0
12 16 20 24 28 0 4 8 12 16 20 24 28 0 4 8 12 16 20 24

1.0 1.0
0.058 Muchac Granite Puquian Pegmatite Puquian Psammite
360 FW2-014 (apatite) FW2-015 (apatite)
FW2-011(zircon)
278.9 2.9 Ma inherited 0.8 0.8
0.054 MSWD = 1.9 340 grains Tera Wasserburg intercept age: Tera Wasserburg intercept age:
236.7 4.1 Ma, MSWD = 5.1 246.0 4.5 Ma, MSWD = 3.3
Pb/238U

207
320 0.6 Pb/206Pb initial: 0.8465 0.6
207
Pb/206Pb initial: 0.8718
0.050
300
206

common Pb 0.4 4000 0.4 4000


0.046
280
Pb/206Pb

0.042 0.2 00 0.2 00


260 24 T 24
800

800

207 S U
Pb/235U
207

238
0.038 0.0 0.0 U/206Pb
0.24 0.28 0.32 0.36 0.40 0.44 0.48 0.52 0 4 8 12 16 20 24 28 0 4 8 12 16 20 24 28

Fig. 5 (continued).

open system behavior most likely associated with the samples residing using an initial 207Pb/206Pb composition of 0.852 (the Stacey and
for a duration in the temperature window for Pb retention (closure Kramers model calculated at 255 Ma) with a conservative uncertainty
temperature estimates are c. 425 C to 500 C; Chamberlain and of 0.02 (2.3%). The samples were anchored due to the relatively small
Bowring, 2001). In these samples, individual 207Pb-corrected ages spread in UPb ratios, but as they plot so close to radiogenic composi-
(essentially a TeraWasserburg lower intercept age for a specied tions, they are very insensitive to the choice of initial Pb composition.
initial 207Pb/206Pb composition; Williams, 1998) are presented. The
207
Pb-corrected ages are calculated using a starting estimate for the 4.3. Zircon UPb data
age of the apatite and adopting an iterative approach using the Stacey
and Kramers terrestrial Pb evolution model (Chew et al. 2011a). Zircon UPb data were acquired from six intrusive rocks in the low-
207
Pb-corrected ages in these samples range from 200.3 4.4 Ma to grade portions of the Huaytapallana Complex (Fig. 2B, 2D, Fig. 5), along
310.6 16 Ma (Fig. 7E) and from 157.1 9.7 Ma to 220.3 3.1 Ma with detrital zircon spectra from two psammites in this same unit. UPb
(Fig. 7H). zircon data were also acquired from zircon rims associated with
leucosome development from three paragneisses in the high-grade por-
4.2. Rutile UPb data tions of the Huaytapallana Complex (Fig. 2C, Figs. 67). The six intrusive
rocks yield concordia ages ranging from 260.1 3.9 Ma to 302.0
Rutile UPb data were acquired for two samples from the low-grade 3.7 Ma (Fig. 5), with one Late Mesoproterozoic age of 1023.5 2.9 Ma
portions of the Huaytapallana Complex (Fig. 2D, Fig. 5). Both samples (FW2-002, Fig. 5C). Two intrusive samples (FW2-001, -010) yield en-
yield rutile UPb data that lie on TeraWasserburg discordias, yielding tirely concordant data, while the remainder yield concordant analyses
lower intercept ages of 257.8 2.3 Ma (FW2-002, Fig. 5F) and with discordant analyses interpreted as older inherited grains (FW2-
250.9 7.4 Ma (FW2-007, Fig. 5P). The UPb data lie close to radiogenic 006, Fig. 5h), inherited grains and Pb loss (FW2-003, Fig. 5d), common
compositions on the TeraWasserburg concordia and were anchored Pb and Pb loss (FW2-002, Fig. 5b) and inherited grains and common
Table 1
Summary of the geochronology data.

Sample Locality UTM UTM UTM Elevation Description UPb zircon UPb apatite UPb rutile age2 UPb monazite age1 Th-Pb monazite
name east north square (masl) age1 age2 age3

FW2-001 Lampa 511,410 8,672,822 18 L 2788 Granodiorite 290.3 4.0 Ma


FW2-002 Pariahuanca 516,401 8,670,733 18 L 2164 Microgranite 1023.5 2.9 Ma 257.3 5.5 Ma 257.8 2.3 Ma
FW2-003 Pariahuanca 516,402 8,670,663 18 L 2161 Mac intrusive 274.1 1.2 Ma 278 18 Ma
FW2-004 Yerba Buena 520,991 8,671,229 18 L 2520 Psammite detrital (720 Ma)4 236.1 1.9 Ma

D.M. Chew et al. / Gondwana Research 35 (2016) 5978


FW2-006 Ocoro 518,778 8,669,841 18 L 2351 Porphyritic volcanic 260.1 3.9 Ma 229 61 Ma
FW2-007 Chuyas 512,833 8,673,097 18 L 2446 Psammite detrital (445 Ma)4 228.1 5.5 Ma 250.9 7.4 Ma
FW2-010 Pahual 510,256 8,673,770 18 L 3144 Granitic Intrusive 302.0 3.7 Ma 281.9 7.8 Ma
FW2-011 Muchac 483,587 8,698,990 18 L 3897 Granitic Intrusive 278.9 2.9 Ma
FW2-014 Puquin 487,985 8,704,142 18 L 3364 Pegmatite 236.7 4.1 Ma
FW2-015 Puquin 488,068 8,704,074 18 L 3420 Psammite 246.0 4.5 Ma
TCH-001 Lasuntay Lake 492,919 8,681,706 18 L 4665 Paragneiss 229 20 Ma 261.6 1.8 Ma 264.0 2.9 Ma
TCH-002 Mt. Jacallate 493,037 8,682,128 18 L 4759 Leucosome in gneiss 260.4 2.5 Ma5 200310 Ma6 262.9 2.0 Ma 260.9 2.5 Ma
TCH-003 Mt. Jacallate 493,327 8,682,523 18 L 4860 Paragneiss 260.4 2.5 Ma5 250 16 Ma 267.5 1.4 Ma 262.8 2.1 Ma
TCH-004 Mt. Lasuntay 494,034 8,681,915 18 L 4840 Metadiorite 157220 Ma6
TCH-007 Mt. Yanauksha 496,436 8,678,999 18 L 4400 Paragneiss 230.3 6.5 Ma 264.1 1.9 Ma 260.9 2.7 Ma
DC 5/5-7 Maraon Complex 168,481 9,243,396 18 M 1422 Leucosome in gneiss 477.9 4.3 Ma7 349.3 5.8 Ma
DC 5/5-10 Balsas 171,840 9,228,140 18 M 1067 Microgranite dyke 343.6 2.6 Ma7 344.3 4.5 Ma
DC 5/6-5 Pomabamba 239,567 9,058,652 18 L 1857 Orthogneiss 445.9 2.4 Ma7 393.8 9.5 Ma
SU 03-20 Hualluniyocc 438,679 8,752,628 18 L 2753 Adamellite 325.4 0.6 Ma7 315 12 Ma
SU 03-21 Huacapistana 446,986 8,761,090 18 L 1662 Granite 310.1 2.3 Ma7 363 44 Ma
DC 4/5-2 Sitabamba 199,248 9,111,706 18 L 3063 Orthogneiss 442.4 1.4 Ma7 365 14 Ma8 titanite age: 437.1 5.3 Ma8
1
Wetherill Concordia age.
2
TeraWasserburg lower intercept age.
3
Weighted average ThPb age.
4
Detrital sample, youngest population or grain in parentheses.
5
Composite concordia age from samples TCH-002 and TCH-003.
6
Does not dene a clear intercept on a TeraWasserburg Concordia, range in 207Pb-corrected ages presented.
7
UPb zircon ages from Chew et al. (2007).
8
UPb apatite and titanite data from Chew et al. (2014).

67
68 D.M. Chew et al. / Gondwana Research 35 (2016) 5978

Pb (FW2-011, Fig. 5s). The detrital zircon spectrum from the psammite
TCH-002 092
1 sample FW2-004 (Fig. 5K) yields a relatively restricted detrital
1: 264.08.2 Ma 3: 1081.212.4 Ma
2: 257.29.5 Ma 4: 1066.417.8 Ma zircon population with peaks at c. 750 Ma and 10001200 Ma. Sample
2
FW2-007 (Fig. 5N) yields a detrital zircon population with peaks at
3
470550 Ma, 750 Ma, and a broad range of detrital zircons from 900
4
to 1800 Ma with the most signicant peaks at 1050 and 1200 Ma.
A very small population of approximately thirty zircon grains was
5 obtained from three high-grade paragneisses (TCH-002, -003, and
TCH-003 081 -007) in the high-grade portions of the Huaytapallana Complex. SEM-
4
1: 259.13.9 Ma CL imaging of these zircons (representative images in Fig. 6) shows
TCH-002 076
2: 1033.023.8 Ma 3 that the zircons range in size from c. 50 to 100 m. The cores of the
1: 254.710.7 Ma
3: 293.84.5 Ma
3: 1001.626.8 Ma
zircons have textures consistent with magmatic zoning with variable
1 CL response from grain to grain and are interpreted as detrital zircons
4: 980.623.9 Ma
3 5: 976.818.5 Ma incorporated into the metasedimentary protolith. These cores are man-
2 tled by thin (c. 5200 m) zircon rims with crudely developed zoning,
1 interpreted as recording high-grade metamorphism and leucosome de-
50 microns velopment. The cores yield ages between 550 and 1100 Ma with Th/U
ratios between 0.1 and 0.5 (Fig. 7a), consistent with a magmatic precur-
Fig. 6. Representative SEM cathodoluminescence (CL) images of zircon from samples TCH- sor. The rims with the lowest Th/U ratios (b0.01; Fig. 7b) yield a
002 and TCH-003 with UPb LA-ICPMS analysis spots and ages illustrated. concordia age of 260.4 2.5 Ma, with the low Th/U ratios consistent

0.24
TCH 002,003,007 TCH 002,003,007 zircon Lasuntay Lake paragneiss
0.049
zircon cores and rims Concordia age: 0.8 TCH-001 (apatite)
0.20 300
0.45 - 0.46 260.4 2.5 Ma
0.047 0.04 Tera Wasserburg intercept age:
1000 0.34 - 0.35 MSWD = 1.08
0.03 22920 Ma, MSWD = 4.0
Pb/238U

0.16 0.11 - 0.54 0.6 207


0.34 - 0.5 0.045 4400 Pb/206Pb initial: 0.8513
0.10 - 0.43
0.12
206

0.18 - 0.25 0.043 0.006 0.007 0.4


600 0.006
0.08 0.38 - 0.39 TCH-002: blue 4000
0.041 0.014 0.009
0.11 TCH-003: red
Pb/238U

0
360

Pb/206Pb
TCH-007: green 250 0.006 0.2
0.04 200
0.07 00
20

00
Th/U ratio: italics 0.039

400
0.005
B C
12
inset figure
206

filled ellipses
to right A
207

207
207
Pb/ U 235
0.037 Pb/235U used in age calc. 0.0
238
U/206Pb
0.00
0.0 0.4 0.8 1.2 1.6 2.0 2.4 0.24 0.28 0.32 0.36 0.40 0 4 8 12 16 20 24 28

1.0 400 1.0


5200 Mt. Jacallete paragneiss 5200 Mt. Jacallete paragneiss
Pb-corr. age

310.6 16 Ma
TCH-002 (apatite) 360 TCH-003 (apatite)
MSWD = 1.5
0.8 320 0.8
Tera Wasserburg intercept age:
280 25016 Ma, MSWD = 3.2
207

207
0.6 240 0.6 Pb/206Pb initial: 0.8597
4400 4400
200
0.4 160 200.3 4.4 Ma 0.4
4000 120 MSWD = 0.19 4000
Pb/206Pb

Pb/206Pb

0 0
0.2 360 0 80 0.2 360 0
0 0
00

20 Mt. Jacallete paragneiss 20


00

400
400

D F
12

40
12

TCH-002 (apatite) E
207

207

238
0.0
238
U/206Pb 0 0.0 U/206Pb
0 4 8 12 16 20 24 28 0 4 8 12 16 20 24 28

1.0 250 1.0


5200 Mt. Lasuntay metadiorite 5200 Mt. Yanauksha paragneiss
Pb-corr. age

TCH-004 (apatite) TCH-007 (apatite)


Pb/206Pb

230
0.8 0.8
210 Tera Wasserburg intercept age:
220.3 3.1 Ma 230.36.5 Ma, MSWD = 3.6
207

190 MSWD = 1.02


207

207
0.6 0.6 Pb/206Pb initial: 0.8402
4400 4400
170
0.4 150 157.1 9.7 Ma 0.4
4000 MSWD = 4.2 4000
Pb/206Pb

0 130
0.2 360 0 0.2 0
360 0
0 Mt. Lasuntay metadiorite 0
00

0
0

400

20
400

2 G 110
I
0

12
12

TCH-004 (apatite)
H
207

238
0.0
238
U/206Pb 90 0.0 U/206Pb
0 10 20 30 40 0 4 8 12 16 20 24 28

Fig. 7. UPb zircon Wetherill concordias (A, B), UPb apatite TeraWasserburg concordias (C, D, F, G, I) and UPb apatite 207Pb-corrected ages (E, H) for the TCH sample suite.
D.M. Chew et al. / Gondwana Research 35 (2016) 5978 69

with a metamorphic origin (Hoskin and Schaltegger, 2003). Analyses 4.4. Monazite UPb and ThPb data
marginally older than the concordia age on Fig. 7b are interpreted as ei-
ther core-rim mixtures with relatively low Th/U ratios indicating that Monazite UPb and ThPb data were acquired for four high-grade
only a small inherited component is present (two oldest red ellipses paragneiss samples from the Huaytapallana Complex (Fig. 8, samples
on Fig. 7b) or a young c. 285 Ma magmatic grain with igneous Th/U TCH-001, -002, -003, -007). UPb Wetherill Concordia ages range
ratios of 0.10.43 (green ellipses on Fig. 7b). from 261.6 1.8 Ma (Fig. 8a, TCH-001) to 267.5 1.4 Ma (Fig. 8e,

232
Lasuntay Lake paragneiss Th-208Pb age (Ma)
0.045 TCH-001 (monazite) 290
280
Concordia age:
261.6 1.8 Ma 280
0.043 MSWD = 1.4

270
Pb/238U

0.041
260
206

0.039
250
240 Weighted average age:

0.037
207
Pb/235U A 240
264.0 2.9 Ma, MSWD = 1.8 B
0.24 0.26 0.28 0.30 0.32 0.34

285
232
Mt. Jacallete paragneiss Th-208Pb age (Ma)
0.047
TCH-002 (monazite)
Concordia age: 290
275
0.045 262.9 2.0 Ma
MSWD = 0.9
265
Pb/238U

0.043 270

255
206

0.041

250
0.039 245
Weighted average age:
0.037
207
Pb/235U C 260.9 2.5 Ma, MSWD = 2.5 D
235
0.26 0.28 0.30 0.32 0.34

Mt. Jacallete paragneiss 300 295


232
Th-208Pb age (Ma)
0.047 TCH-003 (monazite)
Concordia age: 285
0.045 267.5 1.4 Ma
MSWD = 0.9 275

0.043 270 265


Pb/238U

255
0.041
206

245
0.039
235
240 Weighted average age:
207
Pb/ U235 E 262.8 2.1 Ma, MSWD = 2.9 F
0.037 225
0.25 0.27 0.29 0.31 0.33 0.35

232
Mt. Yanauksha paragneiss Th-208Pb age (Ma)
0.045 TCH-007 (monazite) 285
280
Concordia age:
264.1 1.9 Ma 275
0.043 MSWD = 1.4

265
Pb/238U

0.041
255
206

0.039
245 Weighted average age:
240 260.9 2.7 Ma, MSWD = 2.1

0.037
207
Pb/235U G 235
H
0.25 0.27 0.29 0.31 0.33 0.35

Fig. 8. UPb monazite Wetherill concordias (A, C, E, G) and UPb monazite 232Th208Pb ages (E, H) for the TCH sample suite.
70 D.M. Chew et al. / Gondwana Research 35 (2016) 5978

TCH-003) whereas ThPb ages range from 260.9 2.5 Ma (Fig. 8d, TCH- age from Chew et al. 2014 (Figs. 9f,g). The ve apatite UPb ages
003) to 264.0 2.9 Ma (Fig. 8b, TCH-001). With the exception of sample (Figs. 9ae) all lie on TeraWasserburg discordias, yielding lower inter-
TCH-003, all UPb ages and ThPb ages from the same sample are cept ages ranging from 315 12 Ma (SU 03-20, Fig. 9d) to 393.8
within analytical uncertainty, clustering at c. 262 Ma. 9.5 Ma (DC 5/6-5, Fig. 9c). The initial Pb composition of four of the sam-
ples (DC 5/5-7, 5/5-10, 5/6-5, and SU 03-20) was anchored using an ini-
4.5. Apatite UPb data from other localities in the Eastern Cordillera tial 207Pb/206Pb composition derived from the Stacey and Kramers
model with a conservative uncertainty of 0.02 (2.3%), as otherwise
Apatite UPb and ThPb data were acquired for ve samples (one they exhibit a relatively poorly constrained intercept age due to the
leucosome and four granitoids) from the sample suite of Chew et al. small spread in UPb ratios. The samples have moderately radiogenic
(2007) in the central and northern portions of the Eastern Cordillera compositions so they are not particularly sensitive to the choice of initial
of Peru (black boxes highlighted on Fig. 1). These data are presented Pb composition. One apatite sample (SU 03-21, Fig. 9e) was not
on Fig. 9ae, along with one apatite UPb age and one titanite UPb anchored as individual analyses cluster close to common Pb.

Maraon leucosome Balsas Microgranite dyke Sitabamba orthogneiss


0.8 0.8 0.8 DC 5/6-5 (apatite)
DC 5/5-7 (apatite) DC 5/5-10 (apatite)

Tera Wasserburg intercept age: Tera Wasserburg intercept age: Tera Wasserburg intercept age:
0.6 349.3 5.8 Ma, MSWD = 0.91 0.6 344.3 4.5 Ma, MSWD = 0.85 0.6 393.8 9.5 Ma, MSWD = 1.5
4400 anchored at 207Pb/206Pb 4400 anchored at 207Pb/206Pb 4400 anchored at 207Pb/206Pb
initial: 0.8589 0.02 initial: 0.8585 0.02 initial: 0.8621 0.02
0.4 0.4 0.4
0 0 0
360 360 360
Pb/206Pb

00 00 00
Pb/206Pb

Pb/206Pb
0.2 28 0.2 28 0.2 28
00 00 00
0

0
20 A 20 B 20 C

400
120

120

120
400

400
207
207

207
238
238
U/206Pb U/206Pb 238
U/206Pb
0.0 0.0 0.0
0 4 8 12 16 20 0 4 8 12 16 20 0 4 8 12 16 20

Hualluniyocc Adamellite Huacapistana Granite


0.8 SU03-20 (apatite) 0.8 SU03-21 (apatite)

Tera Wasserburg intercept age: Tera Wasserburg intercept age:


0.6 315 12 Ma, MSWD = 1.9 0.6 363 44 Ma, MSWD = 1.3
4400 anchored at 207Pb/206Pb 4400 207
Pb/206Pb initial: 0.8714
initial: 0.8566 0.02 (no anchor)
0.4 0.4
0 0
360 360
00 00
Pb/206Pb

Pb/206Pb

0.2 28 0.2 28
00 00
0

20 D 20
120

120
400

E
400
207

207

238
U/206Pb 0.0
238
U/206Pb
0.0
0 4 8 12 16 20 0 4 8 12 16 20

500 250 0
U-poor apatite Age (Ma)
FT ap 1.0 DC 4/5/1 and DC 4/5/2 titanite:
0.8 and K-fsp
437.1 5.3 Ma, MSWD = 4.4
U-Pb ap
600
U-Pb ttn
0.8
0.6
Pb

Pb

U-Pb zr
G T (C) 1200 DC
206

206

0.6 tita 4/5


nit /1
Pb/

Pb/

0.4 e
207

207

DC 4/5/2 apatite: 0.4


365 14 Ma DC
0.2 MSWD = 2.1 tita 4/5
0.2 nit /2
2800 e
2000 F 2000 H
1200 1200
400 400
0.0 0.0
0 4 8 12 16 20 0 4 8 12 16
238
U/206Pb 238
U/206Pb

Fig. 9. UPb apatite TeraWasserburg concordias from samples from the Eastern Cordillera of Northern (AC) and Central (DE) Peru. UPb apatite (F) and titanite (H) TeraWasserburg
concordias from the Sitabamba orthogneiss taken from Chew et al. (2014).
D.M. Chew et al. / Gondwana Research 35 (2016) 5978 71

5. Synthesis of the timing of metamorphism and magmatism in the sample suite, Fig. 2c) are interpreted as recording crystallization during
Peruvian Eastern Cordillera peak metamorphism, and these high-temperature chronometers
constrain the timing of high-grade metamorphism and leucosome
This section synthesizes the geological evolution of the Huaytapallana development in this sector of the Huaytapallana Complex at 260 Ma
Complex (5.1) before summarizing recent studies on other sectors of (latest Middle Permian). The UPb apatite system records post-
the Eastern Cordillera of Peru (Fig. 10). These regional syntheses are orogenic cooling in the high-grade paragneisses (closure temperature
then incorporated into a model which summarizes the development estimates are c. 425 C to 500 C; Chamberlain and Bowring, 2001).
with the Eastern Cordillera of Peru with time (Section 6). Several samples (TCH-001, TCH-003, TCH-007) yield apatite UPb
TeraWasserburg lower intercept ages consistent with post-orogenic
5.1. Evolution of the Huaytapallana Complex (Fig. 10a) cooling from the 260 Ma metamorphic peak, but two samples
(paragneiss TCH-002 and metadiorite TCH-004) do not yield a Tera
The UPb monazite and zircon and ThPb monazite data from the Wasserburg intercept array, with 207Pb corrected ages ranging
garnetsillimanite paragneisses south of Glaciar Huaytapallana (TCH from 310.6to 200.3 Ma (TCH-002) and 220.3 to 157.1 Ma (TCH-004).

Fig. 10. Geological map of the Eastern Cordillera of Peru based on Fig. 1B. The age of the major metamorphic events are based on the geochronology data of Chew et al. (2007)
(boxes B, C, D, G, H), Cardona (2006, 2007) and Cardona et al. (2009) (box E), Castroviejo et al. (2009), Tassinari et al. (2011) and Willner et al. (2014) (box G) and this study (box A). The
locations of ultramac ophiolitic rocks are taken from Castroviejo et al. (2010). The interpolation of the ages of peak metamorphism in between these study areas is speculative. The names
of the metasedimentary basement inliers (massifs) of the Eastern Cordillera south of 11S are denoted in red text and follow Mgard (1978).
72 D.M. Chew et al. / Gondwana Research 35 (2016) 5978

Although the metadiorite (TCH-004) cuts the Huaytapallana Complex, it and it is likely now tectonically interleaved within the Huaytapallana
does share the same regional tectonic foliation and the cause of this late Complex. The timing of high-grade metamorphism and leucosome
open system behavior in the apatite UPb data from this sample is development in the high-grade paragneisses of the Huaytapallana
uncertain. Complex is constrained to 260 Ma by UPb and ThPb monazite age
The lower-grade, southern sample suite (samples FW2-001FW2- data and UPb dating of zircon rims from leucosomes, consistent with
010, Fig. 2D) did not yield monazite, but rutile was recovered in sample c. 255 Ma rutile growth in the lower-grade schists and psammites to
FW2-002 (Parihuanca microgranite intrusive) and sample FW2-007 the south. The majority of UPb apatite data (c. 260230 Ma) in all
(Chuyas psammite). The UPb rutile system has a closure temperature units are compatible with cooling from this 260 Ma metamorphic peak.
of c. 620 C (Vry and Baker, 2006; Cherniak, 2000). UPb rutile Tera
Wasserburg lower intercept ages (FW2-002, 257.3 5.5 Ma, Fig.4F; 5.2. BalsasCallangate (Fig. 10b)
FW2-007, 250.9 7.4 Ma, Fig. 5P) are thus interpreted as recording
crystallization and are consistent with the 260 Ma metamorphic peak The geology of the BalsasCallangate region consists of high-grade
documented in the high-grade paragneisses. The detrital zircon spec- biotitegarnet paragneisses of the Maran Complex intruded by the
trum in sample FW2-004 (Yerba Buena Psammite, Fig. 5K) yield a detri- post-tectonic BalsasCallangate pluton. Leucosome development in
tal zircon population with two peaks at c. 750 Ma and 10001200 Ma. the Maran Complex paragneisses is constrained by a UPb SIMS zir-
Sample FW2-007 (Chuyas Psammite, Fig. 5N) yields a more dispersed con concordia age of 477.9 4.3 Ma (sample DC 5/5-7, Chew et al.,
detrital zircon population more characteristic of the Proto-Andean mar- 2007). A post-tectonic microgranite dyke has yielded a UPb SIMS zir-
gin of the Eastern Cordillera of Peru (cf. Chew et al., 2008), with peaks at con age of 343.6 2.6 Ma (sample DC 05/5-10, Chew et al., 2007),
470550 Ma, 750 Ma, and a broad range of detrital zircons from 900 to with a prominent inherited population of xenocrystic zircon which de-
1800 Ma with the most signicant peaks at 1050 and 1200 Ma. The ne a UPb SIMS zircon concordia age of 483.8 3.6 Ma. Detrital zircons
Pariahuanca microgranite sample has yielded a UPb zircon crystalliza- from the paragneisses yield no zircons younger than ca. 750 Ma (sample
tion age of 1023.5 2.9 Ma. Given that the youngest detrital zircons DC 5/5-4, Chew et al., 2007). Snchez (1983, 1995) reported KAr dates
peaks which place maximum deposition ages on the psammites in of 346.7 7.3 Ma and 329 10 Ma for the Balsas and Callangate
this sector are c. 750 Ma (FW2-004, Yerba Buena Psammite, Fig. 5K) monzogranitic plutons, which have subsequently yielded 40Ar39Ar
and c. 470 Ma (FW2-007, Chuyas Psammite, Fig. 5N), the c. 1023 Ma ages of c. 350330 Ma (Snchez et al., 2006) and UPb zircon LA-ICPMS
age for the Pariahuanca microgranite implies that it must be a locally re- ages of between 320 and 313 Ma (Miskovic et al., 2009). The c. 478 Ma
stricted basement inlier, but unfortunately, its contact relationships Maran Complex leucosome (DC 5/5-7) and c. 344 Ma post-tectonic
with the metasedimentary rocks of the Huaytapallana Complex are microgranite dyke (DC 5/5-10) yield apatite UPb TeraWasserburg
not exposed. Isolated occurrence of foliated Mesoproterozoic granitoids lower intercept ages of 349.3 5.8 Ma (Fig. 9A) and 344.3 4.5 Ma
have been documented elsewhere in the Eastern Cordillera of Peru (Fig. 9B), respectively. This northernmost portion of the Paleozoic
south of 11S by Miskovoic et al. (2009), including the 985 14 Ma metamorphic belt of the Peruvian Eastern Cordillera is interpreted as a
Satipo tonalite, the 1071 23 Ma Mariposa alkali feldspar granite, high-grade paragneiss unit which underwent Famatinian orogenesis at
and the 1123 13 Ma Querobamba granite). c. 478 Ma prior to intrusion of the BalsasGollnCallangate batholith
All younger igneous intrusive samples in this southern sector of the at c. 320313 Ma. Any subsequent tectonothermal events were not
studied area have been affected by tectonism (either weakly to moder- sufciently hot to reset the UPb apatite and KAr biotite systems.
ately developed foliations, or chloritization and kinking of FeMg
phases such as biotite and strained extinction in quartz). UPb zircon 5.3. SitabambaPomabamba (Fig. 10c)
crystallization ages in the Villa Azul Batholith (Fig. 2D) range from
302.0 3.7 Ma (FW2-010, Pahual Granite, Fig. 5 O) to 290.3 4.0 Ma The protolith of the Sitabamba orthogneiss intruded the Maran
(FW2-001, Lampa Granodiorite, Fig. 5A), whereas intrusions cutting Complex in the southwest portion of the Maran Complex outcrop
the Huaytapallana Complex immediately to the east yield UPb (Wilson et al., 1995; Chew et al., 2007). Thermobarometric estimates
zircon crystallization ages ranging from 274.1 1.2 Ma (FW2-003, for the peak garnet-bearing metamorphic assemblage is c. 700 C and
Pariahuanca metadolerite, Fig. 5D) to 260.1 3.9 Ma (FW2-006, 12 kbar (Chew et al., 2005). Crystallization of the granodioritic protolith
Ocoro metadolerite, a metamorphosed porphyritic intrusive, Fig. 5H). is constrained by UPb zircon concordia ages of 442.4 1.4 Ma (TIMS)
With the exception of sample FW2-010 from the Villa Azul Batholith and 444.2 6.4 Ma (LA-ICPMS) (sample DC 4/5-2 from Sitabamba,
(Pahual Granite, Fig. 5R), all samples from the southern sector of the Chew et al., 2007). Approximately 65 km along strike to the SE at
studied area (FW2-002, FW2-003, FW2-004, FW2-006, FW2-007) Pomabamba, a similar orthogneiss body yields a UPb zircon SIMS
yield apatite UPb TeraWasserburg lower intercept ages consistent concordia age of 445.9 2.4 Ma (sample DC 5/6-5, Chew et al., 2007).
with cooling from the 260 Ma metamorphic peak. Sample FW2-010 in- LA-ICPMS dating of xenocrystic cores in the Sitabamba orthogneiss
stead yields an older apatite UPb TeraWasserburg lower intercept age yield a broad spectrum of detrital zircon ages (cf. Bahlburg and
of 281.9 7.8 Ma and may have been only partially reset during peak Berndt, 2016), with a youngest inherited zircon core of 473 18 Ma
metamorphism at 260 Ma. Only three ages were obtained from the (Chew et al., 2007). Titanite from the peak metamorphic assemblage
northern sampling transect in the Huaytapallana Complex. These in- of the Sitabamba orthogneiss has yielded a LA-ICPMS TeraWasserburg
clude a UPb zircon age of 278.9 2.9 Ma for the Muchac Granite lower intercept age of 437.1 5.3 Ma (Fig. 9H, Chew et al., 2014). Post-
(FW2-011; Fig. 5S), which is within analytical uncertainty of the UPb metamorphic peak cooling is constrained by UPb apatite Tera
zircon age of 292 20 Ma reported by Miskovic et al. (2009) for the Wasserburg lower intercept ages on the cooling path of 365 14 Ma
same pluton. The two apatite UPb TeraWasserburg lower intercept (DC 4/5-2; Fig. 9F, Chew et al., 2014) and 393.8 9.5 Ma (DC 5/6-5,
ages (samples FW2-014 and FW2-015, Fig. 5T and 5U) from this tran- Fig. 9C). The SitabambaPomabamba unit is interpreted as a high-
sect are consistent with cooling from a c. 260 Ma metamorphic peak. grade orthogneiss terrane which underwent orogenesis at c. 437 Ma,
In summary, the evolution of this sector of the Eastern Cordillera east demonstrably younger than the age of peak metamorphism in the Bal-
of Huancayo (Fig. 2A, Fig. 10a) can be considered to represent a sas sector to the north.
metasedimentary sequence (maximum deposition ages of c. 750 Ma
and c. 470 Ma from psammite samples FW2-004 and FW2-007), which 5.4. PatazParcoy (Fig. 10d)b
were possibly deposited on Late Mesoproterozic basement. The contact
relationships with the Late Mesoproterozic basement rocks (UPb zircon In the PatazParcoy region, the Eastern Cordillera is characterized by
age of 1023.5 2.9 Ma from sample FW2-002, Fig. 5C) are not exposed a pre-Silurian metamorphic basement and cover sequences intruded by
D.M. Chew et al. / Gondwana Research 35 (2016) 5978 73

Mississippian calc-alkaline plutons. The pre-Silurian rocks comprise rims (Th/U ratios b1) from one of these migmatites has yielded an age
the type region of the metasedimentary rocks of the Maran of 325 8 Ma (Cardona et al., 2009), while a garnetiferous gneiss asso-
Complex, structurally overlain by volcaniclastics of the Vijus Formation, ciated with a banded migmatite body from the same region has yielded
turbiditic rocks of the Contaya Formation, and felsic volcaniclastics and a SmNd garnetwhole rock isochron of c. 295 13 Ma (Cardona,
turbitiditic rocks of the Atahualpa Formation. The greenschist to lower 2006). Both these ages were obtained from the same area where
amphibolite-facies Maran Complex is believed to have undergone Dalmayrac et al. (1988) reported a Neoproterozoic UPb zircon lower
four phases of deformation (Schreiber, 1989), in contrast to the Lower intercept age from a garnetiferous paragneiss.
Paleozoic cover sequences which have experienced two moderate fold- The greenschist-facies Western Schist Belt (4) is characterized by
ing phases and lower greenschist-facies metamorphism. The Maran signicant intercalations of metabasite. Peak metamorphic conditions
Complex is in faulted contact with its cover sequences, and opinion is di- of 34 kbar and 350400 C were documented by Cardona et al.
vided on whether the contact is conformable or an orogenic unconfor- (2007). The youngest detrital zircons cluster at c. 320 Ma (Cardona
mity (Haeberlin, 2002). A c. 470 Ma maximum depositional age for et al., 2009), KAr muscovite ages cluster at c. 300 Ma (Cardona,
the Maran Complex in the Pataz has been obtained from UPb detri- 2006), and the belt is covered by Permian sedimentary rocks. The geo-
tal zircon dating (Sample AM076, Chew et al., 2007). The middle and chronology data of Cardona et al. (2009) for the HunucoLa Unin re-
upper portions of the Contaya Formation have yielded a graptolite gion demonstrate that the Eastern Schist Belt (youngest detrital zircons
fauna of Didymograptus sagitticaulis, Climacograptus ruedemanni, at c. 460 Ma, mica cooling ages as old as c. 420 Ma) most likely
Diplograptus sp., and Dictionema sp. (Wilson and Reyes, 1964) of middle underwent deformation during the Early Silurian, possibly contempora-
to Upper Darriwilian age (formerly Llanvirn, 465.7458.4 Ma; Cooper neous with the c. 437 Ma event in the Sitabamba and Pataz regions to
et al., 2012). This is consistent with a UPb zircon maximum deposition- the north. Famatinian events (c. 480 Ma) are also recorded by the
al age of 466.8 8.1 Ma from the overlying volcaniclastic Atahualpa small inlier of amphibolite-facies gneisses within the Eastern Schist
Formation (Witt et al., 2013). belt, which has also experienced isolated development of c. 315 Ma
The arc-related Pataz batholith (Schreiber et al., 1990) is a migmatites. This c. 315 Ma event is also recorded in the lower-grade
60-km-long, dioritic to monzogranitic, composite intrusion. The batho- Western Schist Belt (youngest detrital zircons at c. 320 Ma, mica cooling
lith and its aureole are well studied as they form the central part of a ages as old as c. 300 Ma). There is a large range in the 40Ar39Ar and K
160-km-long orogenic gold belt extending along the Eastern Cordillera Ar muscovite and biotite cooling age data of Cardona (2006), with some
in northern Peru. 40Ar39Ar dating of the Pataz Batholith granodiorites ages as young as c. 180 Ma, demonstrating that this sector of the Eastern
(Haeberlin et al., 2004) has yielded plateau ages between 329.2 Cordillera experienced either tectonothermal events in the Mesozoic, or
1.4 Ma (biotite) and 319 3.2 Ma (hornblende). UPb zircon ages partial resetting of the ArAr sytem during Cenozoic Andean orogenesis.
from the batholith are typically older, ranging from 338 to 336 Ma
(Schaltegger et al., 2006; Witt et al., 2013). UPb zircon dating of samples 5.6. Pacococha (Fig. 10f)
from intrusions spatially associated with the Pataz batholith, including
quartz syenite and quartz monzonite, has yielded ages ranging from c. The Pacococha adamellite intrudes the metasedimentary basement
293 to 332 Ma (Miskovic et al., 2009). The youngest detrital zircon ages of the ChupnHuasahuasi massif of Mgard (1978). It is post-tectonic
from the Maran Complex rocks and the biostratigraphical ages from with respect to the ductile deformation fabrics in the metasedimentary
its Early Paleozoic cover in the Pataz region imply the Maran Complex basement and is overlain unconformably by Mississippian sediments
must have undergone post-470 Ma orogenesis, and we tentatively corre- (Mgard, 1978). It is undeformed on a hand specimen scale, but mag-
late this event with the c. 437 Ma event experienced by the Sitabamba matic biotite is pervasively chloritized and the pluton is prominently
orthogneiss, consistent with the Late Ordovician (c. 435 Ma) low-grade jointed with a series of vertical basic dykes exploiting the ssures
metamorphic event inferred in the Maran Complex by Macfarlane (Mgard, 1978). UPb SIMS zircon dating yields a concordia age of
(1999) based on RbSr whole-rock data. Any post-Carboniferous 474.2 3.4 Ma (Sample SU 03-19; Chew et al., 2007), implying that
tectonothermal events in the PatazParcoy region were not sufciently the ductile deformation within the metasedimentary basement oc-
hot to reset the 40Ar39Ar biotite system. curred prior to c. 474.2 Ma and is most likely Famatinian (c. 480 Ma)
in age (Chew et al., 2007). Mgard (1978) reports a KAr biotite age of
5.5. HunucoLa Unin (Fig. 10e) 346 10 Ma from the Pacococha adamellite demonstrating that any
post-Carboniferous tectonothermal events in the ChupnHuasahuasi
Cardona et al. (2009) have subdivided the Maran Complex in the massif were not sufciently hot to reset the 40Ar39Ar biotite system.
HunucoLa Unin region into four main units. These comprise (1) a
small inlier of amphibolite-facies gneisses with peak metamorphic tem- 5.7. Tapo (Fig. 10g)
peratures of 590615 C (Cardona et al. 2007) and intruded by a
mylonitized granitoid. UPb SIMS dating of metamorphic zircon rims The Tapo ultramac massif crops out east of the town of Tarma,
from this gneiss yield an age of 484 12 Ma while the protolith of 40 km to the southeast of the Pacococha adamellite in the Chupn
the mylonitized granitoids is constrained by a UPb TIMS zircon age of Huasahuasi massif. It is one of a series of isolated occurrences of ultra-
468 5 Ma (Cardona et al., 2009). mac rocks that occur scattered along a 300 km long NNWSSE
This inlier is enclosed within unit (2), the Eastern Schist Belt com- trending belt from the Hunuco region in the north to the Tarma region
prising micaceous schists with sporadic intercalations of metabasite (Castroviejo et al., 2010). These ultramac rocks are interpreted as a se-
and calc-silicate rocks. Calculated PT conditions range from 3 to 5 kbar ries of dismembered ophiolites with ocean ridge or ocean island chem-
and 350450 C to 710 kbar and 540660 C. Mac and ultramac ical characteristics (Castroviejo et al., 2009; Tassinari et al., 2011). The
rocks crop out as discontinuous lenses along the eastern margin of Tapo ultramac massif was thrust onto early Carboniferous sedimenta-
this belt (Grandin and Navarro, 1979). The youngest detrital zircons in ry rocks of the Ambo Group, and the basal thrust plane is affected by
the Eastern Schist Belt cluster at c. 465 Ma (Cardona et al., 2009) with Andean folds (Castroviejo et al., 2009). The age of the igneous protolith
40
Ar39Ar and RbSr muscovite metamorphic ages as old as c. 420 Ma of the Tapo ultramac massif is constrained by a SmNd whole-rock
(Cardona et al., 2007). Fossiliferous Carboniferous sedimentary rocks (gabbro and chromitite) isochron of 718 47 Ma (Tassinari et al.,
unconformably overlie the eastern schists (Dalmayrac et al., 1988). 2011) interpreted as the time of oceanic crust formation, while
Unit (3) of Cardona et al. (2009) comprises a series of isolated Nd(718 Ma) values of +8.0 for the chromites and +8.4 for the gabbro
migmatite bodies spatially associated with some undeformed granitoids imply both the chromite and host gabbro formed at the same time
within the Eastern Schist Belt. UPb SIMS dating of metamorphic zircon from the same depleted magma source. Willner et al. (2014) calculated
74 D.M. Chew et al. / Gondwana Research 35 (2016) 5978

high-pressure conditions (1113 kbar at 500540 C) for garnet am- similar scenario is envisaged for the Late Mesoproterozoic basement
phibolites within the Tapo ultramac massif using a PT pseudosection in the Peruvian Eastern Cordillera.
approach, which contrasts with pressuretemperature conditions of There is little evidence on the western Gondwanan margin for
2.5 0.5 kbar, 320 20 C in the adjacent phyllites and greenschists magmatism during the late Neoproterozoic. A-type granitic plutonism
of the metasedimentary basement massif. An SmNd mineral (garnet, has been documented (775690 Ma) in the Eastern Cordillera of Peru
chlorite)whole-rock isochron of 465 24 Ma from a rodingite within (Miskovic et al., 2009) while juvenile extensional magmatism (dacite
the Tapo ultramac massif is interpreted as recording a pulse of uid in- dikes) has been dated at 635 4 Ma in the Antofalla Basement of north-
ux into the disrupted ultramac body along major shear zones during ern Chile (Loewy et al. 2004). Tassinari et al. (2011) attribute the forma-
peak metamorphism (Willner et al., 2014), and this age is within error tion of the igneous protolith of the Tapo ultramac massif to the early
of the K/Ar amphibole age of 448 26 Ma of Tassinari et al. (2011). stages of Rodinian breakup and formation of oceanic lithosphere on
The high-pressure metamorphism indicates that the ophiolitic relic the Proto-Andean margin at 720 Ma. Late Neoproterozoic extension in
body was subducted at a convergent margin or submerged in a collision the Antofalla Basement probably resulted in the partial detachment of
zone during the Famatinian orogenic cycle (Willner et al., 2014). the Antofalla Basement, which was subsequently re-accreted during
the Early Paleozoic Famatinian Orogeny (Loewy et al., 2004; Ramos,
2008), although the inferred collisional suture in the Argentinian Puna
5.8. TarmaLa Merced transect (Fig. 10h)
has been demonstrated to be pre-Famatinian in age (Zimmermann
et al., 2014).
The MarayniocMarairazo massif is composed of polyphase-
The Neoproterozoic age previously assigned to metamorphism of
deformed schists and paragneisses cut by a suite of pre- to post-
the Maran Complex (Dalmayrac et al., 1980) is now regarded to be
tectonic Carboniferous intrusive rocks (Mgard, 1978; Chew et al.,
a result of unresolved multiple inheritance in bulk, unabraded zircon
2007) with abundant development of syn-tectonic migmatites and
fractions (Chew et al., 2007). However, UPb detrital zircon dating of
anatectic melts. The syn-D2 migmatitic Huacapistana granite (sam-
autochthonous Neoproterozoic to early Phanerozoic sedimentary se-
ple SU 03-21) and syn-D2 migmatic paragneisses in the
quences from the northern and central segments of the western
metasedimentary basement (sample SU 03-24) yield UPb SIMS zircon
Gondwanan margin yield a major peak at 550650 Ma (Chew et al.,
ages of 310.1 2.3 Ma and 312.9 3.0 Ma (Chew et al., 2007). The
2008). These authors suggested the source of this detritus could be a
timing of this Late Carboniferous (c. 312 Ma) tectonothermal event is
late Neoproterozoic magmatic arc buried beneath the modern Andean
corroborated by UPb TIMS zircon dating of deformed plutons at
foreland. By at least 530 Ma, segments of the western Gondwanan mar-
325.43 0.57 Ma (the strongly foliated Hualluniyocc adamellite, sam-
gin were clearly a destructive margin (the onset of subduction-related
ple SU 03-20, Chew et al., 2007) and entirely undeformed plutons at
plutonism in the Sierra Pampeanas, Rapela et al. 1998).
307.05 0.65 Ma (Utcuyacu granite, sample SU 03-22, Chew et al.,
In discussing the Paleozoic evolution of the Eastern Cordillera of Peru
2007). The Utcuyacu Granite is a non-deformed monzonitic granite
below, we employ the terminology of the well-established Paleozoic
that cuts all fabrics in the metasedimentary basement and was original-
orogenic and magmatic phases documented in Argentina and Chile.
ly regarded as Andean (Cretaceous to Neogene) in age (Mgard, 1978).
LA-ICPMS UPb zircon dating of inherited zircon cores in the anatectic
6.2. c. 480 Ma (~Early Famatinian) metamorphism
Huacapistana Granite (SU 03-21) and the migmatitic paragneiss (SU
03-24) yield ages which range between 500 and 340 Ma, demonstrating
Famatinian magmatism and metamorphism in Northern Argentina
that the sedimentary protolith was Carboniferous in age and cannot
(Pankhurst et al., 1998) is broadly contemporaneous with the timing
have experienced Famtinian (c. 480 Ma) or c. 435 Ma orogenesis. A
of metamorphism and magmatism documented in the Balsas
garnetbiotite schist from the metasedimentary basement (sample SU
Callangate (Fig. 10b), La Unin-Huanuco (Fig. 10e) and Pacococha
03-25) has yielded pressuretemperature conditions of 600 C and
Tapo (Fig. 10f, Fig. 10g) sectors of the Eastern Cordillera of Peru
11 kbar (Chew et al., 2005) and an unpublished 40Ar39Ar biotite age
(Chew et al., 2007; Miskovic et al., 2009). In the BalsasCallangate
of 204.2 0.8 Ma. The c. 325 Ma Hualluniyocc adamellite (SU 03-20)
and La UninHuanuco sectors, these units are characterized by high-
and c. 310 Ma anatectic Huacapistana granite (SU 03-21) yield apatite
grade rocks (paragneisses with abundant leucosome development in
UPb TeraWasserburg lower intercept ages of 315 12 Ma (Fig. 9D)
Balsas-Callangate; orthogneisses cut by mylonitized granitoids in La
and 363 44 Ma (Fig. 9E), respectively. This implies that post-
UninHuanuco). In the PacocochaTapo sector, peak metamorphic
Carboniferous tectonothermal events in the MarayniocMarairazo mas-
temperatures did not result in anatexis, but the PT and geological
sif were sufciently hot to reset the apatite 40Ar39Ar biotite system, but
constraints from the Tapo ultramac massif indicate that this relic
not the apatite UPb system.
ophiolitic body was subducted at a convergent margin or submerged
in a collision zone during the Famatinian orogenic cycle (Willner
6. Proto-Andean evolution of the Eastern Cordillera of Peru et al., 2014), contemporaneous with the docking of the Paracas terrane
(Ramos, 2008; Miskovic et al., 2009) which is discussed below. A key di-
6.1. Pre-Phanerozoic evolution agnostic test for identifying metasedimentary sequences which have
experienced Famatinian tectono-magmatic events in the Proto-Andes
No basement exposures older than Late Mesoproterozoic (1200 is the presence or absence of Famatinian peaks in their detrital zircon
1000 Ma) have been identied in the Peruvian Eastern Cordillera. spectra, as detritus of this age is ubiquitous in post-Ordovician
The isolated Late Mesoproterozoic basement exposures identied (meta)sedimentary units in the northern and central Andes (Chew
by Miskovic et al. (2009) and this study comprise foliated granitoids et al., 2008).
and crop out between 12 and 11S. The tectonic history of the The cause of Famatinian metamorphism in the Eastern Cordillera has
Mesoproterozoic basement inliers elsewhere in the northern Andes proven controversial in the past. At the same latitude as the Eastern
(e.g. the Garzn Massif, Santa Marta Massif, the Guajira Peninsula and Cordillera, a lack of sialic basement beneath the Peruvian Western
the Santander Massif in Colombia) clearly demonstrates the presence Cordillera has been inferred since at least the Middle Cretaceous based
of a dismembered orogenic belt of Grenvillian age in the Northern on the isotopically extremely primitive intrusives of the Coastal Batho-
Andes. Although Phanerozoic tectonics may have redistributed some lith and the Cordillera Blanca of Peru, a dominantly mac underlying
of the basement terranes, particularly in Colombia, they are still viewed crust with average density of up to 3.0 g/cm3 and a complete absence
as para-autochthonous domains that have remained in proximity to the of zircon inheritance (e.g. Polliand et al., 2005; de Haller et al., 2006).
margin of Amazonia (Cardona et al. 2010; Chew et al., 2011b), and a Chew et al. (2007) inferred that the disposition of the Early to Middle
D.M. Chew et al. / Gondwana Research 35 (2016) 5978 75

Ordovician metamorphism and the associated magmatic arc which they considered equivalent (if slightly younger) to the MiddleLate
identied in the north and central portions of the Eastern Cordillera of Ordovician Oclyic orogenic belt (Ramos et al., 1986; Bahlburg and
Peru arc was related to the presence of an original embayment in the Herv, 1997). In the Puna of northwestern Argentina, folding during
western Gondwanan margin during the early Paleozoic. This was sug- the Oclyic orogeny pre-dates the Faja Eruptiva magmatism which is
gested to explain the present 400 km distance from the present trench constrained to 444 Ma (Bahlburg et al., 2016). The Faja Eruptiva plutons
to the Ordovician magmatic front, which must be a minimum gure intruded in a sinistral strike-slip regime after the main folding phase of
as at least 100 km of crust has been eliminated by subsequent subduc- the Oclyic orogeny had deformed the Ordovician basins of the Puna
tion erosion on the margin (von Huene and Scholl, 1991) and a and the Cordillera Oriental. Bahlburg et al. (2016) link Oclyic orogene-
minimum Andean shortening of 175 km must be also be taken into ac- sis to compressional movements within the accretionary early Paleozoic
count (Ramos, 2008). Ramos (2008) favored an alternative hypothesis, Terra Australis orogen of Cawood (2005).
with the Western Cordillera being underlain by a continental (i.e. sialic)
basement block, the Paracas para-autochthonous terrane, which collid- 6.4. The Late SilurianEarly Carboniferous (c. 420350 Ma) magmatic and
ed during the late Early Ordovician against the Gondwana margin caus- metamorphic gap
ing Famatinian metamorphism and the development of its associated
magmatic arc. The Paracas terrane is inferred to have separated from The central and southern Andes between 4S and 30S, and possibly
the ArequipaAntofalla micro-continent along the Paracas fault which as far south as 42S, has no record of subduction and arc magmatism
denes the northern limit of the ArequipaAntofalla block (Miskovic during the Devonian and early Carboniferous (Bahlburg and Herv,
et al., 2009). Ramos (2008) considered the sialic basement in the 1997; Chew et al., 2007; Cardona et al., 2009; Bahlburg et al., 2009;
offshore platform of central Peru, termed the Paracas High on the conti- Pankhurst et al., 2016), and the margin may have been a passive one
nental shelf (Ramos and Alemn, 2000), to represent a portion of the (Bahlburg and Herv, 1997; Cawood, 2005).
Paracas terrane. The continental shelf high is exposed on the Las
Hormigas de Afuera Islands (Fig. 1, Fig. 10), and subsequent UPb zircon 6.5. c. 315 Ma (Early Gondwanide) metamorphism
SIMS geochronology demonstrates that it represents a continental base-
ment block (predominantly high-grade hypersthene-bearing gneisses) Following this phase of passive margin sedimentation on the West
of Famatinian age (467.9 4.5 Ma, Romero et al., 2013), contempora- Gondwana margin, a continental arc formed during the Middle Carbon-
neous with the arc-derived c. 475465 Ma San Nicols batholith sited iferous (Cawood and Buchan, 2007), as indicated by the intrusion of
on the Arequipa basement of southern Peru (Mukasa and Henry, batholiths in the Chilean Frontal Cordillera (Mpodozis and Kay, 1992)
1990; Loewy et al., 2004). A collisional origin for the Paracas terrane and the emplacement of composite calc-alkaline hornblende and
against the Western Gondwanan margin is also supported by the recog- biotite-bearing granodioritic and granitic batholiths over 1200 km of
nition of the ophiolitic suture within the western margin of the strike of the Eastern Cordillera of Peru (Miskovic et al., 2009). The
Maran Massif by Castroviejo et al. (2009, 2010) and Tassinari et al. depleted HFS elements, mature isotopic signatures, and the presence
(2011), and the recognition of high-pressure Famatinian metamorphic of upper crustal xenoliths suggest these Peruvian Eastern Cordillera
conditions in the Tapo ultramac massif (Willner et al., 2014). The granitoids represent arc-derived magmas that extensively assimilated
dense, isotopically juvenile crust underlying the Western Cordillera the basement country rocks (Macfarlane, 1999). This margin switch to
has been reconciled with a sialic Paracas terrane (Ramos, 2008, 2010; active subduction marks the onset of the pan-Pacic Gondwanide Orog-
Romero et al., 2013) based on the evidence that the intra-arc Cretaceous eny, a terminal phase in the Terra Australis Orogen of Cawood (2005).
Huarmney basin of the Western Cordillera was developed by attenua- Early Gondwanide (c. 315 Ma) metamorphism is recognized in
tion of pre-existing sialic crust and the corresponding eruption of the HunucoLa Unin (Fig. 10e) and TarmaLa Merced transect
large volumes of mac, isotopically juvenile igneous rocks (Mpodozis (Fig. 10h) sectors of the Eastern Cordillera. High-grade Early Gondwanide
and Ramos, 1989). metamorphism near Hunuco is restricted to the isolated development
of c. 315 Ma migmatites in the Eastern Schist Belt, which has predomi-
6.3. c. 435 Ma (Late Famatinian) metamorphism nantly experienced Late Famatinian metamorphism. The Western
Schist Belt has experienced low-grade Early Gondwanide metamor-
High-grade c. 435 Ma (Late Famatinian) metamorphism is most phism based on the presence of detrital zircons as young as c. 320 Ma
evident in the SitabambaPomabamba orthogneiss (Fig. 10c). In and mica cooling ages as old as c. 300 Ma. In the TarmaLa Merced tran-
metasedimentary rocks, this Late Famatinian orogenic event can be sect (Fig. 10h), high-grade Early Gonwandide (c. 315 Ma) metamor-
temporally distinguished from older, c. 480 Ma (Early Famatinian) phism is recognized in syn-D2 migmatites through the Maraynioc
metamorphism by the presence of Famatinian peaks in their detrital Marairazo massif of the Huaytapallana Complex and is conrmed by
zircon spectra, and discriminated from younger (c. 315 Ma) Early geochronology of pre- and post-tectonic Carboniferous plutons. This
Gondwanide orogenesis when they are overlain by Carboniferous Early Gondwanide (c. 315 Ma) event broadly overlaps with a short
sedimentary rocks such as the Mississippian Ambo Group. Based on (c. 5 Ma) gap in arc magmatism in the Eastern Cordillera of Peru attrib-
their detrital zircon spectra, the Maran Complex rocks and its uted to a temporary at-slab geometry (Miskovic et al., 2009) and with
Early Paleozoic cover in the PatazParcoy region (Fig. 10d) have the timing of the high-pressure, low-temperature Toco tectonic event
undergone post-470 Ma orogenesis which is tentatively correlated that folded the Sierra del Tigre turbidites in northern Chile (Bahlburg
with the c. 437 Ma event experienced by the SitabambaPomabamba and Breitkreuz, 1991), with an upper age limit provided by the
orthogneiss. Similarly, in the HunucoLa Unin sector (Fig. 10e), the emplacement of c. 310290 Ma plutons into the folded turbidites
Eastern Schist Belt has undergone post c. 465 Ma metamorphism, with (Bahlburg and Herv, 1997).
40
Ar39Ar and RbSr muscovite metamorphic ages as old as c. 420 Ma.
Magmatism of Late Famatinian age is somewhat restricted and includes 6.6. c. 260 Ma (Late Gondwanide) metamorphism
the c. 442 Ma granodiorite protolith of the SitabambaPomabamba
orthogeniss, an alkali feldspar granite (446.5 9.7 Ma) from the The Late Permian to Late Triassic stage of the Gondwanide orogeny
eastern margin of the Cuzco batholith (Miskovic et al., 2009) and the in Peru saw the emplacement of compositionally restricted, highSiO2
possibly co-magmatic, calc-alkaline Ollantaytambo pyroclastic volca- granites slightly inboard of the Carboniferous arc and peaked at
nics (Bahlburg et al., 2006). We consider this event to mark the last 235240 Ma (Miskovic et al., 2009). The magmatic peak is broadly
orogenic episode of the Early Phanerozoic Famatinian cycle along this contemporaneous with a phase of extension and the formation of the
segment of the Proto-Andean margin. In this sense, it could be continental epiclastics of the Mitu Group (Mgard, 1978). UPb dating
76 D.M. Chew et al. / Gondwana Research 35 (2016) 5978

of detrital zircons in sedimentary horizons constrains the age of the the syn-orogenic Carboniferous, highLILE/HFSE, calc-alkaline, arc-
Mitu Group to 240217 Ma (Spikings et al., 2016). Late Gondwanide derived granitoids to late Permian to late Triassic highThRb, alkali
(c. 260 Ma) metamorphism in the Eastern Cordillera of Peru is rst doc- feldspar granites without signicant HFSE depletion (Miskovic et al.,
umented in this study and has only been identied in the Huaytapallana 2009). This 60 Ma magmatic pulse formed by decompression melting
Complex (in the southern portions of the MarayniocMarairazo massif, associated with lithospheric thinning and peaked at 235240 Ma, con-
the Huaytapallana massif, and the north part of the Huaytapallanakaru temporaneous with the formation of the syn-rift continental epiclastics
massif). This high-grade metamorphic event is contemporaneous with of the Mitu Group (Miskovic et al., 2009) which is constrained to c. 240
a Late Permian deformation phase (Jurua Orogeny) that truncates the 217 Ma (Spikings et al., 2016). This study demonstrates that early in the
Copacabana Group below the overlying 240217 Ma Mitu Group late Permian to late Triassic magmatic pulse, there was a c. 260 Ma
(Fig. 4; Rosas et al., 2007). All magmatism in this sector is pre- to syn- phase of high-grade metamorphism and leucosome development in
tectonic with respect to the regional c. 260 Ma (Late Gondwanide) the southern Peruvian Eastern Cordillera. The origin of the Gondwanide
metamorphic peak, ranging from c. 300260 Ma. The timing of high- orogenic cycle along the Proto-Andean margin is thus best explained by
grade metamorphism and leucosome development in the high-grade changes in tectonic plate reorganization (Ramos, 1988) along an
paragneisses of the Huaytapallana Complex is constrained to accretionary orogen which was undergoing phases of advance and re-
c. 260 Ma by UPb and ThPb monazite age data and UPb dating treat, resulting in magmatic pulses and orogenic phases (e.g. the Late
of zircon rims from leucosomes, consistent with c. 255 Ma rutile growth Carboniferous Early Gondwanide/Toco event and the Late Gondwanide
in the lower-grade schists and psammites to the south. The majority of mid-Permian San Rafael event) that can be correlated along the plate
UPb apatite data (c. 260230 Ma) in all units are compatible with boundary.
cooling from this c. 260 Ma metamorphic peak. The Late Gondwanide
metamorphism in the Eastern Cordillera is broadly contemporaneous
7. Conclusions
with the mid-Permian San Rafael (Sanrafaelic) event (Bahlburg and
Herv, 1997; Ramos, 2000). The San Raphael event is marked by intense
The Huaytapallana Complex east of Huancayo can be considered to
folding and thrusting, resulting in a pronounced angular unconformity
represent a metasedimentary sequence (maximum deposition ages of
between Late Carboniferous to Early Permian turbidites and the exten-
c. 750 Ma and c. 470 Ma from two separate psammite samples) that
sive Permo-Triassic Choiyoi Volcanics in Argentina and Chile (Kay et al.,
was possibly deposited on Late Mesoproterozic basement. The timing
1989; Ramos, 2000; Ramos and Aleman, 2000).
of high-grade metamorphism and leucosome development in the
high-grade paragneisses of the Huaytapallana Complex is constrained
6.7. Origin of Gondwanide metamorphism in the Eastern Cordillera of Peru
to c. 260 Ma by UPb and ThPb monazite age data and UPb dating
of zircon rims from leucosomes, consistent with c. 255 Ma rutile growth
During the nal stages of assembly of the Pangean supercontinent
in the lower-grade schists and psammites to the south. The majority
(c.320250 Ma; Cawood and Buchan, 2007), major plate boundary reor-
of UPb apatite data (c. 260230 Ma) in all units are compatible
ganization was accompanied by regional orogenesis (the 305230 Ma
with cooling from this c. 260 Ma metamorphic peak. UPb zircon geo-
Gondwanide Orogeny) affecting the entire Pacic margin of Pangea. It
chronology of pre-tectonic plutons yield ages ranging from c. 302 Ma
marks the nal stage in the Neoproterozoic to late Paleozoic Terra
to c. 260 Ma.
Australis Orogen of Cawood (2005). It represents an accretionary orog-
This is the rst documented occurrence of c. 260 Ma metamorphism
eny on a plate margin, which in the absence of colliding tectonic blocks
in this sector of the Proto-Andean margin, and a regional synthesis of
during subduction and plate convergence must be driven by an increase
existing geochronological constraints from the Eastern Cordillera of
in coupling across the plate boundary (Cawood and Buchan, 2007).
Peru demonstrates that the margin has experienced a polycyclic oro-
Possible mechanisms for increased coupling within accretionary
genic history. The Famatinian orogenic history can be subdivided
orogens include subduction of buoyant oceanic lithosphere (resulting
into two phases of deformation and magmatism at c. 480 Ma (~ Early
in at-slab subduction), terrane accretion or tectonic plate reorganiza-
Famatinian) and c. 435 Ma (Late Famatinian). Following the Late Siluri-
tion. Flat-slab subduction is thought an unlikely mechanism for the
anEarly Carboniferous (c. 420350 Ma) magmatic and metamorphic
Gondwanide deformation on the Proto-Andean margin as it should re-
gap, two phases of Late Paleozoic metamorphism and magmatism at
sult in igneous rocks with adakitic compositions, a cessation of igneous
c. 315 Ma (Early Gondwanide) and c. 260 Ma, and are best developed
activity in the established arc and oceanic plateau accretion (Cawood
in the central and southern portions of the Eastern Cordillera of Peru.
and Buchan, 2007). These phenomena are not observed on the West
These events, corresponding to the Famatinian and Gondwanide oro-
Gondwanan margin; neither is there evidence for terrane accretion dur-
genic cycles, can be correlated into the Proto-Andean margin of
ing the Carboniferous and Permian outboard of the Peruvian Eastern
Argentina and Chile and are thus of regional extent. The evolution of
Cordillera. In contrast, tectonic plate reorganization can result in rapid
the Proto-Andean margin is thus best explained by changes in tectonic
increases in the absolute motion of the overriding plate, resulting in oro-
plate reorganization in a long-lived Paleozoic accretionary orogen
genesis which affects the length of the orogen/plate boundary synchro-
which was undergoing phases of advance and retreat, resulting in
nously. It can therefore be distinguished from the effects of at-slab
magmatic pulses and orogenic phases that can be correlated along the
subduction or terrane accretion by correlating the timing of tectonic
plate boundary.
events along the margin and also coeval tectonic events around the
world that may have led to plate reorganization (Cawood and Buchan,
2007). Acknowledgments
Accretionary orogens can be broken into two end-member types:
retreating orogens which undergo extension in response to lower This material is in part based upon works supported by Science
plate retreat resulting in forearc accretion and opening of back arc ba- Foundation Ireland under Grant No. 12/IP/1663. GP thanks Dr. Carlos
sins, and advancing orogens which develop when the overriding plate Cabrera and Msc. Ciro Bedia of the Universidad Nacional Mayor de San
is advancing toward the downgoing plate resulting in arc accretion Marcos for nancial support, Msc. Hugo Rivera Mantilla, Dr. Luis Reyes
and retro-arc fold and thrust belts (Cawood and Buchan, 2007). The Rivera, and Agapito Snchez Fernndez of the Instituto Geolgico
mode of plate convergence may switch between that of an advancing Minero y Metalrgico (INGEMMET) for comments and suggestions on
or retreating orogen, a switching process that appears to be intimately the manuscript and the petrology group (GEIPPE) at San Marcos (Luis
linked to accretionary orogenesis (Collins, 2002). This process is docu- ngel Daz, Edison Morales, and Jorge Espejo) for their support in the
mented in the Eastern Cordillera of Peru, where there is a shift from eldwork.
D.M. Chew et al. / Gondwana Research 35 (2016) 5978 77

Appendix A. Supplementary data Chew, D.M., Schaltegger, U., Mikovic, A., Fontignie, D., Frank, M., 2005. Deciphering the
tectonic evolution of the Peruvian segment of the Gondwanan margin. 6th Interna-
tional Symposium on Andean Geodynamics (ISAG 2005, Barcelona), Extended
Supplementary data to this article (Appendix Asample descrip- Abstracts, pp. 166169.
tions; Appendices BEapatite, rutile, zircon, and monazite U-Pb age Chew, D.M., Sylvester, P.J., Tubrett, M.N., 2011a. UPb and Th-Pb dating of apatite by LA-
ICPMS. Chemical Geology 280, 200216.
data) can be found online at Supplementary data associated with this Cochrane, R., Spikings, R.A., Chew, D., Wotzlaw, J.F., Chiaradia, M., Tyrrell, S., Schaltegger,
article can be found in the online version, at http://dx.doi.org/10.1016/ U., Van der Lelij, R., 2014. High temperature (N350 C) thermochronology and mech-
j.gr.2016.03.016. anisms of Pb loss in apatite. Geochimica et Cosmochimica Acta 127, 3956.
Collins, W.J., 2002. Hot orogens, tectonic switching, and creation of continental crust.
Geology 30, 535538.
Cooper, R.A., Sadler, P.M., Hammer, ., Gradstein, F.M., 2012. The Ordovician Period, the
References Geologic Time Scale 2012. pp. 489523.
Cordani, U.G., Sato, K., Teixeira, W., Tassinari, C.C.G., Basei, M.A.S., 2000. Crustal evolution
Aleinikoff, J.N., Schenck, W.S., Plank, M.O., Srogi, L.A., Fanning, C.M., Kamo, S.L., of the South American platform. In: Cordani, U.G., Milani, E.J., Thomaz Filho, A.,
Bosbyshell, H., 2006. Deciphering igneous and metamorphic events in high- Campos, D.A. (Eds.), Tectonic Evolution of South America. 31st International Geolog-
grade rocks of the Wilmington Complex, Delaware: morphology, cathodoluminescence ical Congress, Rio de Janeiro, pp. 1940.
and backscattered electron zoning, and SHRIMP UPb geochronology of zircon and Dalmayrac, B., Laubacher, G., Marocco, R., 1980. Gologie des Andes pruviennes:
monazite. Geological Society of America Bulletin 118, 3964. caractres gnraux de l'volution gologique des Andes pruviennes. Travaux et
Bahlburg, H., Berndt, J., 2016. Provenance from zircon UPb age distributions in crustally Document de l'ORSTROM, Paris.
contaminated granitoids. Sedimentary Geology 336, 161170. Dalmayrac, B., Laubacher, G., Marocco, R., 1988. Caracteres generales de la evolucin
Bahlburg, H., Berndt, J., Gerdes, A., 2016. The ages and tectonic setting of the Faja Eruptiva geolgica de los Andes Peruanos. Instituto Geolgico Minero y Metalrgico, Lima.
de la Puna Oriental, Ordovician, NW Argentina. Lithos 256-257, 4154. de Haller, A., Corfu, F., Fontbote, L., Schaltegger, U., Barra, F., Chiaradia, M., Frank, M.,
Bahlburg, H., Breitkreuz, C., 1991. Paleozoic evolution of active margin basins in the Alvarado, J.Z., 2006. Geology, geochronology, and Hf and Pb isotope data of the
southern Central Andes (northwestern Argentina and northern Chile). Journal of RaulCondestable iron oxidecoppergold deposit, central coast of Peru. Economic
South American Earth Sciences 4, 171188. Geology 101, 281310.
Bahlburg, H., Carlotto, V., Cardenas, J., 2006. Evidence of Early to Middle Ordovician arc Grandin, G., Navarro, J., 1979. Las rocas ultrabasicas en el Peru, las intrusiones lenticulares
volcanism in the Cordillera Oriental and Altiplano of southern Peru, Ollantaytambo y los sills de la regin de Huanuco-Monzon. Boletn Sociedad Geolgica del Per 63,
Formation and Umachiri beds. Journal of South American Earth Sciences 22, 5265. 99115.
Bahlburg, H., Herve, F., 1997. Geodynamic evolution and tectonostratigraphic terranes of Haeberlin, Y., 2002. Geological and Structural Setting, Age, and Geochemistry of the
northwestern Argentina and northern Chile. Geological Society of America Bulletin Orogenic Gold Deposits at the Pataz Province, Eastern Andean Cordillera. Universit
109, 869884. de Genve, Terre et Environnement, Peru, p. 182.
Bahlburg, H., Vervoort, J.D., Du Frane, S.A., Bock, B., Augustsson, C., Reimann, C., 2009. Haeberlin, Y., Moritz, R., Fontbote, L., Cosca, M., 2004. Carboniferous orogenic gold de-
Timing of crust formation and recycling in accretionary orogens: insights learned posits at Pataz, Eastern Andean Cordillera, Peru: geological and structural framework,
from the western margin of South America. Earth-Science Reviews 97, 215241. paragenesis, alteration, and Ar-40/Ar-39 geochronology. Economic Geology and the
Black, L.P., Kamo, S.L., Allen, C.M., Aleinikoff, J.N., Davis, D.W., Korsch, R.J., Foudoulis, C., Bulletin of the Society of Economic Geologists 99, 73112.
2003. TEMORA 1: a new zircon standard for Phanerozoic UPb geochronology. Hoskin, P.W.O., Schaltegger, U., 2003. The composition of zircon and igneous and meta-
Chemical Geology 200, 155170. morphic petrogenesis. In: Hanchar, J.M., Hoskin, P.W.O. (Eds.), Zircon. Reviews in
Cardona, A., 2006. Reconhecimento da evoluo tectnica da proto-margem andina do Mineralogy and Geochemistry Vol. 53. Mineralogical Society of America, pp. 2762.
centro-norte peruano, baseada em dados geoqumicos e isotpicos do embasamento Jaillard, E., Soler, P., 1996. Cretaceous to early Paleogene tectonic evolution of the northern
da Cordilheira Oriental na regio de Hunuco-La Unin. Instituto de Geocincias, Uni- Central Andes (018S) and its relations to geodynamics. Tectonophysics 259, 4153.
versity of So Paulo, p. 256. Kay, S.M., Ramos, V.A., Mpodozis, C., Sruoga, P., 1989. Late Paleozoic to Jurassic silicic
Cardona, A., Chew, D., Valencia, V.A., Bayona, G., Miskovic, A., Ibanez-Mejia, M., 2010. magmatism at the Gondwana margin: analogy to the middle Proterozoic in North
Grenvillian remnants in the Northern Andes: Rodinian and Phanerozoic paleogeo- America? Geology 17, 324328.
graphic perspectives. Journal of South American Earth Sciences 29, 92104. Leon, W., Palacios, O., Vargas, L., Sanchez, A., 2000. Memoria explicativa del Mapa
Cardona, A., Cordani, U.G., Ruiz, J., Valencia, V.A., Armstrong, R., Chew, D., Nutman, A., Geologico del Peru (1999), Carta Geolgica Nacional, Boletn No. 136, escala 1:
Sanchez, A.W., 2009. UPb zircon geochronology and Nd isotopic signatures of the 1,000,000. Lima. Instituto Geolgico Minero y Metalrgico.
pre-Mesozoic metamorphic basement of the eastern Peruvian Andes: growth and Litherland, M., Aspden, J.A., Jemielita, R.A., 1994. The Metamorphic Belts of Ecuador.
provenance of a Late Neoproterozoic to Carboniferous accretionary orogen on the HMSO, London, London.
northwest margin of Gondwana. Journal of Geology 117, 285305. Loewy, S.L., Connelly, J.N., Dalziel, I.W.D., 2004. An orphaned basement block: the Arequi-
Cardona, A., Cordani, U.G., Sanchez, A., 2007. Metamorphic, geochronological and geo- paAntofalla basement of the central Andean margin of South America. Geological
chemical constraints from the Pre-Permian basement of the eastern Peruvian Society of America Bulletin 116, 171187.
Andes (10 S): a Paleozoic extensionalaccretionary orogen? 20th Colloquium on Luvizotto, G.L., Zack, T., Meyer, H.P., Ludwig, T., Triebold, S., Kronz, A., Munker, C., Stockli,
Latin American Earth Sciences, Kiel, Germany, pp. 2930 D.F., Prowatke, S., Klemme, S., Jacob, D.E., von Eynatten, H., 2009. Rutile crystals as
Castroviejo, R., Machar, J., Castro, P., Pereira, E., Rodrigues, J.F., Tassinari, C.G., Willner, A., potential trace element and isotope mineral standards for microanalysis. Chemical
Acosta, J., 2010. Signicado de las oolitas Neoproterozoicas de la Cordillera Oriental Geology 261, 346369.
del Per (9301130), XV Congresso Peruano de Geologa, Resmenes Extendidos. Macfarlane, A., 1999. Isotopic studies of northern Andean crustal evolution and ore metal
Sociedad Geolgica del Per, Cusco, pp. 5153. sources. In: Skinner, B.J. (Ed.), Geology and Ore Deposits of the Central Andes. Society
Castroviejo, R., Rodrigues, J.F., Acosta, J., Pereira, E., Romero, D., Quispe, J., Esp, J.A., 2009. of Economic Geologists, Boulder, pp. 195217.
Geologa de las ultramatas pre-andinas de Tapo y Acobamba, Tarma, Cordillera McDowell, F.W., McIntosh, W.C., Farley, K.A., 2005. A precise 40Ar39Ar reference age for
Oriental del Peru. Geogaceta, Sociedad Geolgica de Espaa 46, 710. the Durango apatite (UTh)/He and ssion-track dating standard. Chemical Geology
Cawood, P.A., 2005. Terra Australis Orogen: Rodinia breakup and development of the 214, 249263.
Pacic and Iapetus margins of Gondwana during the Neoproterozoic and Paleozoic. Mgard, F., 1978. Etude gologique des Andes du Prou Central. Contribution a l'tude
Earth Science Reviews 69, 249279. gologique des Andes N 1. Mmoire ORSTROM 86, 310 (Paris).
Cawood, P.A., Buchan, C., 2007. Linking accretionary orogenesis with supercontinent Mikovi, A., Schaltegger, U., Spikings, R.A., Chew, D.M., Koler, J., 2009. Tectono-magmatic
assembly (vol 82, pg 217, 2007). Earth-Science Reviews 85, 82. evolution of western Amazonia: geochemical characterisation and zircon UPb geo-
Cerrn, F., Ticona, P., 2003. Memoria descriptiva de la revisin y actualizacin del chronologic constraints from the Peruvian Eastern Cordilleran granitoids. Geological
cuadrngulo de Pampas (25-n) Escala 1:100 000, Lima, Instituto Geolgico Minero Society of America Bulletin 121, 12981324.
y Metalrgico. Mpodozis, C., Kay, S.M., 1992. Late Paleozoic to Triassic evolution of the Gondwana mar-
Chamberlain, K.R., Bowring, S.A., 2001. Apatite-feldspar UPb thermochronometer: a gin: evidence from Chilean Frontal Cordilleran batholiths (28S to 31S). Geological
reliable, mid-range (450 C), diffusion-controlled system. Chemical Geology 172, Society of America Bulletin 104, 9991014.
173200. Mpodozis, C., Ramos, V.A., 1989. The Andes of Chile and Argentina. In: Ericksen, G.E.,
Cherniak, D.J., 2000. Pb diffusion in rutile. Contributions to Mineralogy and Petrology 139, Caas Pinochet, M.T., Reinemud, J.A. (Eds.), Geology of the Andes and Its Relation
198207. to Hydrocarbon and Mineral Resources. Circumpacic Council for Energy and Mineral
Chew, D.M., Cardona, A., Miskovic, A., 2011b. Tectonic evolution of western Amazonia Resources, Houston, pp. 5990.
from the assembly of Rodinia to its break-up. International Geology Review 53, Mukasa, S.B., Henry, D.J., 1990. The San-Nicols batholith of coastal Peru: early Paleozoic
12801296. continental arc or continental rift magmatism? Journal of the Geological Society 147,
Chew, D.M., Magna, T., Kirkland, C.L., Miskovic, A., Cardona, A., Spikings, R., Schaltegger, U., 2739.
2008. Detrital zircon ngerprint of the Proto-Andes: evidence for a Neoproterozoic Pankhurst, R.J., Herve, F., Fanning, C.M., Calderon, M., Niemeyer, H., Griem-Klee, S., Soto, F.,
active margin? Precambrian Research 167, 186200. 2016. The pre-Mesozoic rocks of northern Chile: UPb ages, and Hf and O isotopes.
Chew, D.M., Petrus, J.A., Kamber, B.S., 2014. UPb LA-ICPMS dating using accessory Earth-Science Reviews 152, 88105.
mineral standards with variable common Pb. Chemical Geology 363, 185199. Pankhurst, R.J., Rapela, C.W., Saavedra, J., Baldo, E., Dahlquist, J., Pascua, I., Fanning, C.M.,
Chew, D.M., Schaltegger, U., Koler, J., Whitehouse, M.J., Gutjahr, M., Spikings, R.A., 1998. The Famatinian magmatic arc in the central Sierras Pampeanas: an Early to
Mikovic, A., 2007. UPb geochronologic evidence for the evolution of the Mid-Ordovician continental arc on the Gondwana margin. In: Pankhurst, R.J.,
Gondwanan margin of the north-central Andes. Geological Society of America Rapela, C.W. (Eds.), The Proto-Andean Margin of Gondwana. London, Special Publica-
Bulletin 119, 697711. tions, Geological Society, pp. 343367.
78 D.M. Chew et al. / Gondwana Research 35 (2016) 5978

Paredes, J., 1994. Geologa del Cuadrngulo de Jauja, Hoja 24-M. Boletn. Serie a: Carta Schreiber, D.W., Fontbote, L., Lochmann, D., 1990. Geologic setting, paragenesis, and
Geolgica Nacional Vol. 48. Instituto Geolgico Minero y Metalrgico, Lima. physicochemistry of gold quartz veins hosted by plutonic rocks in the Pataz region.
Paton, C., Hellstrom, J., Paul, B., Woodhead, J., Hergt, J., 2011. Iolite: freeware for the visu- Economic Geology 85, 13281347.
alisation and processing of mass spectrometric data. Journal of Analytical Atomic Shi, G.H., Li, X.H., Li, Q.L., Chen, Z.Y., Deng, J., Liu, Y.X., Kang, Z.J., Pang, E.C., Xu, Y.J., Jia, X.M.,
Spectrometry 26, 25082518. 2012. Ion microprobe UPb age and Zr-in-rutile thermometry of rutiles from the
Petrus, J.A., Kamber, B.S., 2012. VizualAge: a novel approach to laser ablation ICP-MS UPb Daixian rutile deposit in the Hengshan Mountains, Shanxi Province, China. Economic
geochronology data reduction. Geostandards and Geoanalytical Research 36, Geology 107, 525535.
247270. Slama, J., Koler, J., Condon, D.J., Crowley, J.L., Gerdes, A., Hanchar, J.M., Horstwood, S.A., M.,
Pffner, O.A., Gonzalez, L., 2013. MesozoicCenozoic evolution of the western margin of Morris, G.A., Nasdala, L., Norberg, N., Schaltegger, U., Schoene, B., Tubrett, M.N.,
South America: case study of the Peruvian Andes. Geosciences 3, 262310. Whitehouse, M.J., 2008. Pleovice zircon a new natural reference material for
Polliand, M., Schaltegger, U., Frank, M., Fontbot, L., 2005. Formation of intra-arc UPb and Hf isotopic microanalysis. Chemical Geology 249, 135.
volcanosedimentary basins in the western ank of the central Peruvian Andes during Spikings, R., Reitsma, M.J., Boekhout, F., Mikovi, A., Ulianov, A., Chiaradia, M., Gerdes, A.,
late cretaceous oblique subduction: eld evidence and constraints from UPb ages Schaltegger, U., 2016. Characterization of Triassic Rifting in Peru and implications for
and Hf isotopes. International Journal of Earth Sciences 94, 231242. the early disassembly of western Pangaea. Gondwana Research 35, 124143.
Ramos, V.A., 1988. The tectonics of the Central Andes; 30 to 33S latitude. In: Clark, S., Stewart, J.W., Evernden, J.F., Snelling, N.J., 1974. Age determinations from Andean Peru: a
Burchel, D. (Eds.), Processes in Continental Lithospheric Deformation, Special reconnaissance survey. Geological Society of America Bulletin 85, 11071116.
Paper of the Geological Society of America, pp. 3154. Tassinari, C.C.G., Castroviejo, R., Rodrigues, J.F., Acosta, J., Pereira, E., 2011. A
Ramos, V.A., 2000. The Southern Central Andes. In: Cordani, U.G., Milani, E.J., Thomaz Neoproterozoic age for the chromitite and gabbro of the Tapo ultramac massif,
Filho, A., Campos, D.A. (Eds.), Tectonic Evolution of South America. 31st International Eastern Cordillera, Central Peru and its tectonic implications. Journal of South
Geological Congress, Rio de Janeiro, pp. 561604. American Earth Sciences 32, 429437.
Ramos, V.A., 2008. The basement of the Central Andes: the Arequipa and related terranes. Thomas, W.A., Astini, R.A., Mueller, P.A., Gehrels, G.E., Wooden, J.L., 2004. Transfer of the
Annual Review of Earth and Planetary Sciences 36, 289324. argentine Precordillera terrane from Laurentia: constraints from detrital-zircon
Ramos, V.A., 2009. Anatomy and global context of the Andes: Main geologic features and geochronology. Geology 32, 965968.
the Andean orogenic cycle. In: Kay, S.M. (Ed.), Backbone of the Americas: Shallow Thomson, S.N., Gehrels, G.E., Ruiz, J., Buchwaldt, R., 2012. Routine low-damage apatite U
Subduction, Plateau Uplift, and Ridge and Terrane Collision. Geological Society of Pb dating using laser ablation-multicollector-ICPMS. Geochemistry Geophysics
America Memoir Vol. 204, pp. 3165. Geosystems 13.
Ramos, V.A., 2010. The Grenville-age basement of the Andes. Journal of South American von Huene, R., Scholl, D.W., 1991. Observations at convergent margins concerning sedi-
Earth Sciences 29, 7791. ment subduction, subduction erosion, and the growth of continental crust. Reviews
Ramos, V.A., Aleman, A., 2000. Tectonic evolution of the Andes. In: Cordani, U.G., Milani, of Geophysics 29, 279316.
E.J., Thomaz Filho, A., Campos, D.A. (Eds.), Tectonic Evolution of South America. Vry, J.K., Baker, J.A., 2006. LA-MC-ICPMS PbPb dating of rutile from slowly cooled gran-
31st International Geological Congress, Rio de Janeiro, pp. 635685. ulites: conrmation of the high closure temperature for Pb diffusion in rutile.
Ramos, V.A., Jordan, T.E., Allmendinger, R.W., Mpodozis, C., Kay, S.M., Cortes, J.M., Palma, Geochimica et Cosmochimica Acta 70, 18071820.
M., 1986. Paleozoic terranes of the Central ArgentineChilean Andes. Tectonics 5, Wasteneys, H.A., Clark, A.H., Farrar, E., Langridge, R.J., 1995. Grenvillian granulite-facies
855880. metamorphism in the Arequipa massif, Peru a LaurentiaGondwana link. Earth
Rapela, C.W., Pankhurst, R.J., Casquet, C., Baldo, E., Saavedra, J., Galindo, C., 1998. Early and Planetary Science Letters 132, 6373.
evolution of the proto-Andean margin of South America. Geology 26, 707710. Wiedenbeck, M., Alle, P., Corfu, F., Grifn, W.L., Meier, M., Oberli, F., von Quadt, A.,
Romero, D., Valencia, K., Alarcon, P., Pena, D., Ramos, V.A., 2013. The offshore basement of Roddick, J.C., Spiegel, W., 1995. Three natural zircon standards for UThPb, LuHf,
Peru: evidence for different igneous and metamorphic domains in the forearc. Journal trace element and REE analyses. Geostandards Newsletter 19, 123.
of South American Earth Sciences 42, 4760. Williams, I.S., 1998. U-Th-Pb Geochronology by Ion Microprobe. In: McKibben, M.A.,
Rosas, S., Fontbote, L., Tankard, A., 2007. Tectonic evolution and paleogeography of the Shanks III, W.C., Ridley, W.I. (Eds.), Applications of Microanalytical Techniques to
Mesozoic Pucara Basin, Central Peru. Journal of South American Earth Sciences 24, Understanding Mineralizing Processes. Reviews in Economic Geology, pp. 135.
124. Willner, A.P., Tassinari, C.C.G., Rodrigues, J.F., Acosta, J., Castroviejo, R., Rivera, M., 2014.
Snchez, A., 1983. Nuevos datos KAr en algunas rocas del Per. Boletn Sociedad Contrasting Ordovician high- and low-pressure metamorphism related to a
Geolgica del Per 71, 193202. microcontinent-arc collision in the Eastern Cordillera of Peru (Tarma province). Jour-
Snchez, A., 1995. Geologa de los cuadrngulos de Bagua Grande, Jumbilla, Lonya Grande, nal of South American Earth Sciences 54, 7181.
Chachapoyas, Rioja, Leimebamba y Bolvar (hojas 12-g, 12-h, 13-g, 13-h, 13-i, 14-h, Wilson, J., Reyes, L., 1964. Geologa del Cuadrngulo de Pataz, Carta Geolgica Nacional,
15-), Lima (Instituto Geolgico Minero y Metalrgico). Boletn Serie a 9 (91 pp.) Instituto Geolgico Minero y Metalrgico, Lima.
Snchez, A., Carrasco, S., Galdos, J., Lipa, V., 2006. Geologia del Batolito Gollon - Buldibuyo Wilson, J., Reyes, L., Garayar, J., 1995. Geologa de los Cuadrngulos de Pallasca,
Informe Tecnico (Codigo: A6139), Lima. Instituto Geolgico Minero y Metalrgico. Tayabamba, Corongo, Pomabamba, Carhuaz y Huari Escala 1:100 000, Carta Geolgica
Schaltegger, U., Chew, D., Miskovic, A., 2006. Neoproterozoic to Early Mesozoic Evolution Nacional, Boletn Serie A 60, 63 p., Lima, Instituto Geolgico Minero y Metalrgico.
of the Western Gondwana Margin: Evidence from the Eastern Cordillera of Peru. XIII Witt, W.K., Hagemann, S.G., Villanes, C., Zeng, Q.T., 2013. New geochronological results
Congresso Peruano de Geologia, Lima, October, 2006 (Extended Abstract). and structural evolution of the Pataz gold mining district: implications for the timing
Schoene, B., Bowring, S.A., 2006. UPb systematics of the McClure Mountain syenite: and origin of the batholith-hosted veins. Ore Geology Reviews 50, 143170.
thermochronological constraints on the age of the Ar-40/Ar-39 standard MMhb. Con- Zack, T., Stockli, D.F., Luvizotto, G.L., Barth, M.G., Belousova, E., Wolfe, M.R., Hinton, R.W.,
tributions to Mineralogy and Petrology 151, 615630. 2011. In situ UPb rutile dating by LA-ICP-MS: Pb-208 correction and prospects for
Schreiber, D.W., 1989. Zur Genese von Goldquarzgngen der Pataz-Region im Rahmen geological applications. Contributions to Mineralogy and Petrology 162, 515530.
der geologischen Entwicklung der Ostkordillere Nordperus (unter besonderer Zimmermann, U., Bahlburg, H., Mezger, K., Berndt, J., Kay, S.M., 2014. Origin and age of ul-
Bercksichtigung der Distrikte Parcoy, La lima und Buldibuyo). Heidelberger tramac rocks and gabbros in the southern Puna of Argentina: an alleged Ordovician
Geowissenschaftliche Abhandlungen 29 (235 pp.). suture revisited. International Journal of Earth Sciences 103, 10231036.

You might also like