Sie sind auf Seite 1von 45

GR-01278; No of Pages 45

Gondwana Research xxx (2014) xxxxxx

Contents lists available at ScienceDirect

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

GR Focus Review

The geological history of northwestern South America: from Pangaea to the early
collision of the Caribbean Large Igneous Province (29075 Ma)
Richard Spikings a,, Ryan Cochrane b, Diego Villagomez c, Roelant Van der Lelij d, Cristian Vallejo e,
Wilfried Winkler f, Bernado Beate g
a

Department of Earth and Environmental Science, University of Geneva, Rue des Maraichers 13, Geneva 1205, Switzerland
Thomson Reuters, London, UK
Tectonic Analysis Ltd., Geneva, Switzerland
d
Norges Geologiske Underskelse, 7491 Trondheim, Norway
e
Geostrat S.A., Quito, Ecuador
f
ETH Zrich, Geological Institute, ETH Zentrum, NO E 59, Sonneggstrasse 5, CH-8092 Zrich, Switzerland
g
Facultad de Geologa, Minas y Petrleos, Escuela Politcnica Nacional, A.P. 17-01-2759, Quito, Ecuador
b
c

a r t i c l e

i n f o

Article history:
Received 24 March 2014
Received in revised form 4 June 2014
Accepted 25 June 2014
Available online xxxx
Handling Editor: M. Santosh
Keywords:
Pangaea
South America
Tectonic reconstruction
Geochronology
Geochemistry
Thermochronology

a b s t r a c t
Northwestern South America preserves a record of the assembly of western Pangaea, its disassembly and
initiation of the far western Tethys Wilson Cycle, subsequent Pacic margin magmatism and ocean plateau
continent interaction since the Late Cretaceous. Numerous models have been presented for various time
slices although they are based on either spatially restricted datasets, or dates that are inaccurate estimates
of the time of crystallisation. Here we review a very large quantity of geochronological, geochemical,
thermochronological, sedimentological and palaeomagnetic data that collectively provide tight constraints for
geological models. These data have been collected over a trench (Pacic)-parallel distance of N1500 km
(Colombia and Ecuador), and reveal important temporal trends in rifting and subduction. The temporal framework for our model constraints are obtained from robust, concordant zircon U-Pb ages of magmatic rocks during
29075 Ma. The Late Cretaceous thermal history of the margin (b350 C) is described by 40Ar/39Ar and ssion
track data, and the higher temperature and thus older (pre-75 Ma) history are constrained by apatite U-Pb
thermochronology. Variations in the isotopic compositions of Hf (zircon), Nd (whole) and O (quartz) with
time have been used to track the evolution of the source of magmatism, and are used as proxies for crustal
thickness. Atomic chemical compositions, combined with isotopes and dense mineral assemblages are used to
differentiate between continental and oceanic environments. These data show that rifting within western
Pangaea started at 240 Ma, leading to sea oor spreading between blocks of Central and South America by
216 Ma. Pacic active margin commenced at 209 Ma, and continued until 115 Ma above an east-dipping subduction zone that was rolling back, attenuating South America and forming new continental crust. The opening of the
South Atlantic drove South America westwards, compressed the Pacic margin of northwestern South America
at 115 Ma and obducted an exhumed subduction zone. Passive margin conditions prevailed until the Oceanic
Plateau and its overlying intra-oceanic arc (The Rio Cala Arc) collided and accreted to South America at 75 Ma.
2014 Published by Elsevier B.V. on behalf of International Association for Gondwana Research.

Contents
1.
2.
3.
4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Geological framework of northwestern South America (Colombia and Ecuador)
Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Triassic: the disassembly of Pangaea and the formation of a passive margin . .
4.1.
Historical perspective and occurrence . . . . . . . . . . . . . . . .
4.2.
Geochronology . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.1.
Cordillera Real of Ecuador and Cordillera Central of Colombia .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

Corresponding author. Tel.: +41(0)223793176.


E-mail addresses: richard.spikings@unige.ch (R. Spikings), ryan.cochrane@thomsonreuters.com (R. Cochrane), diego.villagomez@gmail.com (D. Villagomez), roelantvdl@gmail.com
(R. Van der Lelij), Cristian.vallejo@yahoo.com (C. Vallejo), wilfried.winkler@erdw.ethz.ch (W. Winkler), bbeate@uio.satnet.net (B. Beate).

http://dx.doi.org/10.1016/j.gr.2014.06.004
1342-937X/ 2014 Published by Elsevier B.V. on behalf of International Association for Gondwana Research.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

0
0
0
0
0
0
0

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

4.2.2.
Comparison with the ages of Permian and Triassic rocks in Venezuela and Peru . . . . . . . . .
Geochemistry of the granites and migmatites . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3.1.
Cordillera Real of Ecuador and Cordillera Central of Colombia . . . . . . . . . . . . . . . . .
4.3.2.
Comparison with Permian and Triassic rocks in Venezuela and Peru . . . . . . . . . . . . . .
4.4.
Geochemistry of the amphibolites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.5.
Zircon Hf isotope geochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.5.1.
Zircon Hf isotope geochemistry of the granites and migmatitic leucosomes . . . . . . . . . . .
4.5.2.
Zircon Hf isotope geochemistry of the amphibolites . . . . . . . . . . . . . . . . . . . . . .
4.5.3.
Comparison with Zircon Hf isotope compositions in Peru . . . . . . . . . . . . . . . . . . .
4.6.
Thermal histories during the Triassic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.7.
Interpretation: Permian and Triassic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.7.1.
Arc magmatism and metamorphism during 290240 Ma along western Pangaea . . . . . . . .
4.7.2.
Initiating the disassembly of western Pangaea during 240200 Ma . . . . . . . . . . . . . . .
4.8.
Conjugate margins to northwestern Gondwana . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.9.
Rifting between North and South America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.
Latest TriassicLower Cretaceous: arc magmatism and tectonic switching . . . . . . . . . . . . . . . . . . .
5.1.
Historical perspective and occurrence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.1.
Latest TriassicJurassic granitoid intrusions . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.2.
Late JurassicEarly Cretaceous rocks to the west of the Jurassic intrusions . . . . . . . . . . . .
5.2.
Geochronology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.1.
Latest Triassic and Jurassic intrusions: Cordillera Real, Cordillera Central and the Santander Massif
5.2.2.
Early Cretaceous magmatic and sedimentary rocks: Cordillera Real and Cordillera Central . . . .
5.2.3.
Comparison with Peru and the Merida Andes of Venezuela . . . . . . . . . . . . . . . . . .
5.3.
Geochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.1.
Latest Triassicearliest Cretaceous granitoids . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.2.
Early Cretaceous igneous rocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.3.
Comparison with magmatic rocks from Peru . . . . . . . . . . . . . . . . . . . . . . . . .
5.4.
The tectonic setting during the latest TriassicJurassic (210145 Ma) . . . . . . . . . . . . . . . . . .
5.4.1.
Why is there a gap in the Jurassic arc in Peru? . . . . . . . . . . . . . . . . . . . . . . . .
5.5.
The tectonic setting during the Early Cretaceous (145115 Ma) . . . . . . . . . . . . . . . . . . . .
5.6.
Compression during the Early Cretaceous . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.7.
The Chaucha Terrane and the Taham Terrane . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.8.
Comparison with Peru (145115 Ma) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.
The tectonic history of northwestern South America during 11575 Ma . . . . . . . . . . . . . . . . . . . .
6.1.
The formation of the Caribbean Large Igneous Province and its collision with South America. . . . . . . .
6.1.1.
Geochemistry and geochronology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1.2.
Time of initial accretion with South America . . . . . . . . . . . . . . . . . . . . . . . . .
6.1.3.
The nature of the CLIPSouth America suture . . . . . . . . . . . . . . . . . . . . . . . .
7.
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Appendix A.
Supplementary data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3.

1. Introduction
The northwestern South American plate hosts a Grenvillian basement, which was modied during the amalgamation and disassembly
of Pangaea, subsequent prolonged active margin magmatism and the
collision of the voluminous Caribbean Large Igneous Province, which
added new crust to South America. This manuscript is mainly a review
of a very large quantity of data, although some new U-Pb (apatite)
and 40Ar/39Ar dates are presented. These data are used to generate
robust constraints for any model that describes the disassembly
and fragmentation of western Pangaea, the subsequent evolution of
the Pacic margin offshore northwestern South America during the
JurassicEarly Cretaceous, and the early evolution of the Caribbean
region and its interaction with South America. The review is organised
into sections according to geological time, and compares the evolution
of northwestern South America (north of 5S) with the margin of Peru
during 29075 Ma.
Wide disagreements exist over the tectonic origin of voluminous
magmatic units, including Triassic anatectites, JurassicEarly Cretaceous
arc rocks, obducted M-HP/LT rocks and allochthonous units that comprise the western cordilleras and the forearc. These contrasting interpretations result in signicantly different interpretations for plate
reconstructions during the TriassicLate Cretaceous (e.g. Litherland
et al., 1994; Spikings et al., 2001; Pratt et al., 2005; Pindell and

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

Kennan, 2009; Villagmez and Spikings, 2013; Cochrane et al., 2014a).


Contrasting models partly exist because of the misinterpretation of
K/Ar and Rb/Sr dates that were published in the 1980's and 1990's as
accurate estimates of crystallisation age, ignoring the effects of daughter
isotope loss. We discard K/Ar and Rb/Sr dates in favour of recently published concordant zircon U-Pb dates, which are more accurate estimates
of crystallisation age. The U-Pb dates are combined with geochemical
and isotope data, sedimentological data and eld relationships to
constrain the magmatic source regions and tectonic environment
within which the rocks formed. The tectonic histories are subsequently
investigated using thermochronological and palaeomagnetic data.
We show that western Pangaea started to disassemble by rifting of
continental crust of Central America from South America at ~ 240 Ma,
and that these had completely separated by ~216 Ma. The northwestern
margin of South America remained passive until ~ 209 Ma within
Pangaea, and arc magmatism occurred during 209114 Ma, accompanying the separation of North and South America at ~ 180 Ma. The
Jurassic magmas formed in a continental arc, which questions previous
interpretations that place the Jurassic trench far from the location of the
Jurassic arcs, due to the presence of suspect continental terranes. We
draw a single east-dipping subduction zone during 209114 Ma,
which retreated oceanward and extended the South American margin,
culminating in compression that drove rock uplift and exhumation.
Finally, we present evidence for an east-facing intra-oceanic arc, which

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

formed on an intra-oceanic plateau prior to its collision and accretion


with South America at ~ 75 Ma, resulting in growth of the continent.
Our interpretation differs from other models that rely on large-scale
plate reconstructions to constrain the positions of continents and subduction zones.

2. Geological framework of northwestern South America (Colombia


and Ecuador)
The South American Plate forms a relict part of western Gondwana,
and formed during the opening of the Central Atlantic, South Atlantic
and the Inter-American Gap (Gulf of Mexico and the proto-Caribbean)
during 180120 Ma (E.g. Eagles, 2007). The Atlantic margin remains
passive whereas the western margin became active at ~ 500480 Ma
(e.g. Pankhurst et al., 2000; Van der Lelij, 2013), soon after the opening
of the Iapetus Ocean during 570535 Ma (Cawood et al., 2001). The
Northern plate margin formed during rifting from Yucatan in the Middle
Jurassic (Pindell et al., 2005), forming the region of the proto-Caribbean.
At the present time, the Northern Andes are separated from the Central
Andes by the Huancabamba Deection (at ~6S), which marks a distinct
change in the strike orientation of the Andean chain (see inset in Fig. 1).
The oldest rocks exposed in northwestern South America are
Grenvillian aged gneisses, which are dispersed in inliers throughout
the Eastern Cordillera of Colombia, Santander Massif and the Sierra
Nevada de Santa Marta (Fig. 1; e.g. Restrepo-Pace et al., 1997), where
they are considered to form part of the Chibcha Terrane (e.g. Toussaint
and Restrepo, 1994), although none crop out in Ecuador. The Phanerozoic
rocks, which form the focus of this review, can be separated into a
relatively undifferentiated oceanic Late Cretaceous sequence, which is
faulted against older, differentiated continental crust.
The Late Cretaceous oceanic rocks form the basement to the Western
Cordillera and the forearcs of Colombia (Calima Terrane; Fig. 1; e.g. Kerr
et al., 1997) and Ecuador (Pallatanga-Pion Terrane; e.g. Vallejo et al.,
2009). Geochemical, isotopic and geochronological data suggest that
these ultramac and mac rocks formed in an oceanic hot-spot setting
during 9987 Ma (e.g. Kerr et al., 1997; Vallejo et al., 2006; Villagmez
et al., 2011), and that they are equivalent to the oceanic plateau rocks
that form the Caribbean Plate (e.g. Sinton et al., 1998). Field relationships and U-Pb zircon dates show that the oceanic plateau was intruded
by an east-facing intra-oceanic arc prior to their collision with South
America in the Campanian (e.g. Feininger and Bristow, 1980; Vallejo
et al., 2006), although this is not consistent with the plate reconstructions of Lebrat et al. (1987) and Pindell and Kennan (2009), who suggest
that arcs at this time were west-facing and were also intruding the
continental margin of South America. The accretion of allochthons of
the Caribbean Large Igneous Province added at least 5x1061x107 km3
of new crust to the South American Plate (Cochrane, 2013).
The Early Cretaceous continental margin hosts N-S trending linear
belts that are exposed within the Cordillera Central of Colombia,
and the Cordillera Real of Ecuador (Fig. 1). Traversing eastwards from
the Campanian suture, these are M-HP/LT complexes of amphibolites,
blueschists and eclogites, Early Cretaceous arc rocks of the
Quebradagrande and Alao sequences, undifferentiated Palaeozoic
rocks, which underwent anatexis in the Triassic, foliated Early
Cretaceous arc plutons, and large unfoliated Jurassic batholiths along
the eastern ank of the Cordillera Central in Colombia, and Cordillera
Real in Ecuador. Further east, the Eastern Cordillera of Colombia has
no equivalent topographic feature in Ecuador, and it includes the high
plains of the Santander Massif in the north (Fig. 1). The Santander
Massif hosts the oldest segment of the latest TriassicEarly Cretaceous
continental arc sequence, which has no equivalent assemblage in
Ecuador. Litherland et al. (1994) suggest that these belts are all in
tectonic contact, and these allochthonous units were juxtaposed during
compression at 140120 Ma. Alternatively, Pratt et al. (2005) suggest
that the contacts are intrusive, and the rock units within Ecuador are

autochthonous, which is similar to the model proposed by Villagmez


and Spikings (2013) and Cochrane (2013) for Ecuador and Colombia.
The retro-arc foreland basins of the Middle and Lower Magdallena
Valley basins in Colombia, and the Oriente Basin of Ecuador preserve a
record of the evolution of the margin during the JurassicRecent, and
these are utilised throughout the review to support or negate various
hypotheses.
3. Methodology
This review presents a very large quantity of data that was mainly
previously peer reviewed and published, and the details of the methodologies used by each study are provided in the respective publications. A
summary of the geochronological data and pertinent geochemical and
isotopic data is presented in Tables 1 and 2, and the complete geochemical dataset used to plot all of the geochemical gures is provided as a
supplementary le.
A majority of the geochronological, isotopic and geochemical
data for the period spanning 290100 Ma (Mikovi et al., 2009;
Villagmez et al., 2011; Cochrane, 2013; Cochrane et al., 2014a,b) was
acquired at laboratories at the Universities of Geneva and Lausanne
(Switzerland), and at the Goethe Universitt Frankfurt. Accuracy and
external reproducibility of the methods were determined by analysing
i) Harvard 91500 (Wiedenbeck et al., 1995) and Pleovice zircon
(Slma et al., 2008) during geochronology, ii) GJ-1 and Pleovice zircon
(Slma et al., 2008) during Hf isotopic analyses, iii) standard JNdi-1 for
Nd isotopic analyses. Other geochronological analyses were performed
at laboratories at Curtin University, the University of Grenoble (Riel
et al., 2013), the Australian National University (Vinasco et al., 2006;
Restrepo et al., 2011), the University of Arizona (Bustamante et al.,
2010; Cardona et al., 2010; Weber et al., 2010) and Westflische
Wilhelms-Universitt Munster (Lu-Hf; John et al., 2010). All zircon
U-Pb dates that are used in the interpretations are concordant, and
dene a single age population with respect to their mean square
weighted deviate statistic. The methodology used for each measurement is highlighted in Tables 1 and 2.
Dense mineral assemblages and ssion-track data for the period
after 100 Ma (e.g. Spikings et al., 2000, 2010) were obtained at laboratories at ETH-Zrich. A majority of geochemical and geochronological data
from the Caribbean Large Igneous Province was obtained from laboratories at the University of Geneva (40Ar/39Ar; Luzieux et al., 2006), the
Australian National University (U-Pb; Vallejo et al., 2006), and the
University of Lausanne (Mamberti et al., 2003).
4. Triassic: the disassembly of Pangaea and the formation of a
passive margin
Triassic rocks within the Cordillera Real, Amotape Terrane (Ecuador)
and Cordillera Central (Colombia) are dominated by widely dispersed
outcrops of variably foliated granites, gneissic granites and migmatites,
and less abundant amphibolites, ultramac rocks and metasedimentary rocks (Fig. 2). Several studies have shown that the
magmatic and metamorphic rocks that formed during the Triassic are
geochemically distinct from younger magmatic rocks, and formed in a
different tectonic environment.
4.1. Historical perspective and occurrence
Triassic rocks within the Cordillera Real and Amotape Terrane of
Ecuador include granitoids of the Tres Lagunas and Moromoro units,
migmatites of the Sabanilla unit, geographically scattered amphibolitic
dykes and sills (Piedras and Monte Olivo units), and sedimentary
rocks of the Piuntza unit (Fig. 2). The granites were rst described by
Colony and Sinclair (1932). Mapping by the British Geological Survey
during 19861993 linked these occurrences into a semi-continuous
belt, and they were grouped into the Loja Terrane (Litherland et al.,

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

GP

Caribbean Plate

SNSM

PSB

F
SM

10N

SP

Venezuela
SM

CAF

ChocPanam
Terrane

OPF

MA

MMV

Tahami
Terrane
EC

Pa

GF

5N

IF

UMV

WC
LB

CC

0m

CPV
Huancabamba
Deflection

300

SJ

Calima
Terrane

Colombia
300

Pallatanga-Pion
Terrane

0m

Carnegie
Ridge

Coastal
Batholith
Arequipa
Terrane

CR

IAD
WC
SLB

South American Plate

Pu

N
Ri azc
dg a
e

Nazca Plate

SZ

PF

OB

Ecuador

IF
PB

Triassic Anatectite
Latest Triassic Jurassic arc

RC

Amotape
Terrane

Early Cretaceous arc


CL

Peru

80W

Early Cret. M-HP/LT

75W

Fig. 1. Digital elevation model for northwestern South America showing the cordilleras, suspect terranes, main faults and the exposure of TriassicEarly Cretaceous magmatic rocks in Ecuador
and Colombia. Inset shows the location of the Arequipa Terrane in southern Peru, and the Coastal Batholith. Faults: CAF: Cauca-Almaguer Fault, GF: Garrapatas Fault, IF: Ibagu Fault, OPF: OtuPericos Fault, PaF: Palestina Fault, PF: Peltetec Fault, PuF: Pujili Fault, SJF: San-Jeronimo Fault, SMF: Santa Marta Fault. Other abbreviations: CC: Cordillera Central, CL: Celica-Lancones Basin, CPV:
Cauca-Pata Valley, CR: Cordillera Real, GP: Guajira Peninsula, IAD: Interandean Depression, IF: Ingapirca Fault (western boundary of the Guamote Sequence), LB: Llanos Basin, MA: Merida
Andes, MMV: Middle Magdalena Valley Basin, OB: Oriente Basin, PB: Pion Block, PSB: Plato-San Jorge Basin, RC: Raspas Complex, SLB: San Lorenzo Block, SM: Santander Massif, SP: Sierra
de Perija, SZ: Sub-Andean Zone, UMV: Upper Magdalena Valley Basin, WC: Western Cordillera. Geology from Litherland et al. (1994) and Gmez et al. (2007).

1994) along with undifferentiated Palaeozoic metamorphic rocks of the


Chiguinda and Agoyan units. The Loja Terrane is bound to the east by the
Llanganates Fault, and to the west by the Baos fault. Triassic igneous and
metamorphic lithologies are dominated by cordierite and garnet bearing

monzogranites and diorites (Tres Lagunas unit), and medium- to highgrade, sillimanite and kyanite bearing orthogneisses and migmatites
(Sabanilla Unit). Litherland et al. (1994), Noble et al. (1997), Riel
et al. (2013) and Cochrane et al. (2014a) present a large quantity of

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

geochronological and geochemical data from the igneous and metamorphic rocks in Ecuador (Table 1). The Piuntza unit consists of metamorphosed and skarnied siliciclastic rocks, tuffs and limestones that
host Triassic bivalves (Litherland et al., 1994). The unit is exposed
beyond the structural limits of the Loja Terrane of Litherland et al.
(1994) along the eastern ank of the southern Cordillera Real (Fig. 2),
where it is surrounded by the Jurassic Zamora Batholith although the
nature of the contact is either unknown or unreported.
Widely dispersed and variably deformed Permian and Triassic
meta-granitoids, ultra-macmac rocks and metasedimentary rocks
occur within the northern Cordillera Central of Colombia (Fig. 2).
These rocks were initially described by Hall et al. (1972), Feininger
et al. (1972) and Gonzlez (1980), who considered them to be PermoTriassic on the basis of K/Ar dates. Restrepo and Toussaint (1988)
suggested that the Permo-Triassic rocks dened the basement of the
fault-bounded Taham Terrane, and placed them within the Cordillera
Central Polymetamorphic Complex (Restrepo and Toussaint, 1982).
The Taham Terrane of Restrepo and Toussaint (1988) is bound by the
Ot-Pericos fault to the east, which separates it from Grenvillian aged
metamorphic basement of the Chibcha Terrane (e.g. Ordoez-Carmona
et al., 1999), and the San Jernimo Fault to the west, which separates it
from the Quebradagrande Arc (Fig. 1). Maya and Gonzlez (1995) and
Villagmez et al. (2011) group the Triassic metamorphosed igneous
and sedimentary rocks into the Cajamarca Complex, which will be
adopted in this manuscript. The geological map of Colombia (Gmez
et al., 2007) reveals a paucity of Triassic lithologies within the Cordillera
Central south of the Ibagu Fault (Fig. 2). Vinasco et al. (2006), Martnez
(2007), Cardona et al. (2010), Montes et al. (2010), Weber et al. (2010),
Restrepo et al. (2011), Villagmez et al. (2011) and Cochrane et al.
(2014a) present a large quantity of geochemical data and concordant
zircon U-Pb dates (Table 1, Fig. 2) from the Cajamarca Unit, conrming
the Permo-Triassic crystallisation ages of the igneous rocks. Similar to
the Cordillera Real in Ecuador, the Permo-Triassic rocks of the Cordillera
Central intrude and are faulted against Palaeozoic metamorphic rocks
such as the La Miel Unit (e.g. Restrepo et al., 1991; Villagmez et al.,
2011), although these will not be considered further in this review.
Martnez (2007) report a series of metagabbros and amphibolites in
the northern Cordillera Central, which they attribute to a Triassic
ophiolitic sequence, referred to as the Aburr Ophiolite (Fig. 2).
Permo-Triassic igneous and metamorphic rocks have also been
recognised in the Guajira Peninsula, Sierra Nevada de Santa Marta and
at the base of boreholes drilled though the Plato-San Jorge Basin located
north of the Cordillera Central (Cardona et al., 2010; Montes et al., 2010;
Weber et al., 2010; Fig. 1). Granites from the basement of the Plato-San
Jorge Basin are mildly deformed, while the intrusions from the Sierra
Nevada de Santa Marta are mylonitised.
4.2. Geochronology
4.2.1. Cordillera Real of Ecuador and Cordillera Central of Colombia
Early attempts to date the Permo-Triassic crystalline rocks utilised
the K/Ar and Rb/Sr methods (e.g. Feininger et al., 1972; Hall et al.,
1972 McCourt et al., 1984; Restrepo et al., 1991; Litherland et al.,
1994; Ordoez and Pimentel, 2002), resulting in a large scatter of ages
spanning between the PermianTertiary due to variable degrees of
daughter isotope loss. This review of geochronological work is restricted
to more accurate and peer-reviewed measurements of the crystallisation
ages of granitoids and mac intrusions, which have been provided by
numerous concordant zircon and few monazite U-Pb dates (Table 1),
obtained using TIMS, SHRIMP and LA-ICPMS. Unless otherwise stated,
the LA-ICPMS and SHRIMP dates that are reported here were obtained
from the rims of zircons, and are considered to date either the most
recent phase of magmatic crystallisation or the most recent metamorphic event that crystallised zircon.
Fourteen metagranites and migmatites of the Tres Lagunas and
Sabanilla units in the Cordillera Real and Amotape Complex of

Ecuador yield concordant zircon and monazite U-Pb dates ranging


between 207.6 9.2 Ma and 247.2 4.3 Ma (Figs. 2 and 3a; Table 1;
Litherland et al., 1994; Aspden et al., 1995; Chew et al., 2008; Riel
et al., 2013; Cochrane et al., 2014a). These dates overlap with concordant zircon U-Pb dates obtained from twenty six gneissic granites and
pegmatites exposed in the Cordillera Central, Sierra Nevada de Santa
Marta and the Guajira Peninsula in Colombia, which range between
222 10 Ma and 288.1 4.5 Ma (Vinasco et al., 2006; Cardona et al.,
2010; Montes et al., 2010; Weber et al., 2010; Villagmez et al., 2011;
Restrepo et al., 2011; Cochrane et al., 2014a; Figs. 2 and 3a). A majority
of these crystalline rocks are Triassic, although six rocks from northern
Colombia yield Permian ages. Ordonez-Carmona et al. (2001) report
a Sm-Nd whole rockgarnet isochron date of 226 17 Ma from a
granulite in the Taham Terrane, which they interpret as a cooling age
following peak metamorphism. Multi-phase, plateau 40Ar/39Ar dates
(Table 1) from Triassic granites and migmatites in Colombia and
Ecuador (Spikings et al., 2001; Vinasco et al., 2006; Cochrane et al.,
2014a) are younger than the U-Pb dates obtained from the same rocks,
and Triassic dates span between 213.7 0.9 Ma and 243 4 Ma
(Table 1). The 40Ar/39Ar dates reect the time of cooling of each rock
through mineral-specic argon partial retention zones (i.e. 550300 C;
hornblende, muscovite and biotite) subsequent to crystallisation and
metamorphic retrogression.
Concordant zircon U-Pb dates of amphibolites and a plagiogranite
from the Cordillera Real and Cordillera Central range between
216.6 0.4 Ma and 243 4 Ma (Fig. 3d; Noble et al., 1997; Vinasco
et al., 2006; Martnez, 2007; Cochrane et al., 2014a). Within Ecuador,
these are exposed as dykes and sills (e.g. the Piedras and Monte Olivo
units; Fig. 2), whereas they are more massive in the Taham Terrane of
the Cordillera Central. The youngest of these ages was obtained from a
plagiogranite that formed by hydrothermal metamorphism of the
Aburr Ophiolite in northern Colombia, and thus is a minimum age for
the ophiolite (Martnez, 2007).
A majority of U-Pb dates of zircons extracted from meta-granites
and migmatites range between 240 and 230 Ma (Fig. 3a, d), and a comparison with latitude (Fig. 3d) does not reveal any trends for the Triassic
period, with the exception of a possible increase in age in far northern
Colombia within the Guajira Peninsula. Permian ages are restricted to
exposures in the Sierra Nevada de Santa Marta, and faulted blocks in
the region of the Ibagu Fault (Fig. 2). This may reect exposure, or perhaps approximate the primary distribution of Permian magmatic
intrusions.
Some concordant 206Pb/238U dates determined by in-situ methods
were also obtained from the cores of zircon grains that were identied
using cathodoluminescence. A frequency analysis of the distribution of
these older dates from the granitoids and migmatitic leucosomes yields
broad peaks at 420580 Ma and 9501200 Ma (Fig. 3b), which are the
ages of protolith rocks and inherited grains. These age peaks are typical
of the distribution of dates obtained from detrital zircons from most
Palaeozoic terranes along western South America (e.g. Chew et al.,
2007). The younger age group broadly corresponds with the age of the
Famatinian Arc, the Braziliano Orogeny and the timing of rifting during
the fragmentation of Rodinia. The Famatinian arc (~ 510415 Ma;
E.g. see zircon U-Pb ages presented in Pankhurst et al., 2000; Cardona
et al., 2007; Chew et al., 2007; Bahlburg et al., 2009; Mikovi et al.,
2009; Villagmez et al., 2011; Van der Lelij, 2013) formed during the
subduction of Pacic lithosphere beneath western South America subsequent to the fragmentation of Rodinia, and has been recorded in
Venezuela, Colombia, Peru and Argentina (see previous citations).
Inherited zircons with U-Pb dates spanning the 450650 Ma range
also occur in Cretaceous and Tertiary sedimentary rocks of the Amazon
Foreland Basin in Ecuador (Martin-Gombojav and Winkler;, 2008),
although intrusions of the Famatinian arc have not been recorded in
Ecuador, and within Colombia they are only recorded within the northern Cordillera Central (440470 Ma; La Miel orthogneiss; Villagmez
et al., 2011), Quetame, Floresta and Santander massifs (Horton et al.,

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

Table 1
Summary of data collected from Permo-Triassic rocks of Ecuador and Colombia.
Sample

Unit

Lithology

S-type granites and migmatitic leucosomes


Ecuador
Eastern Cordillera
09RC25
Tr. Lagunas
metagranite
09RC31
Tr. Lagunas
metagranite
09RC42
Sabanilla
metagranite
09RC53
Tr. Lagunas
metagranite
09RC44
Sabanilla
paragneiss
09RC45
Sabanilla
Paragneiss
09RC56
Tr. Lagunas
metagranite
11RC03
Agoyan fm.
metagranite
Tres Lagunas
Tr. Lagunas
granite
Amotape Complex
09RC40
Moromoro
migmatite
VI-08-12
La Bocana
migmatite
PU-08-10
La Bocana
migmatite
AV-08-31
La Bocana
migmatite
AV-08-28d
La Bocana
migmatite
Moromoro
Moromoro
granite
Colombia
Central Cordillera
10RC04
Cajamarca
metagranite
10RC40
Cajamarca
metagranite
10RC41
Cajamarca
metagranite
10RC42
Cajamarca
metagranite
10RC43
Cajamarca
metagranite
10RC53
Cajamarca
metagranite
10RC66
Cajamarca
qtz-Schist
10RC69
Cajamarca
metagranite
10RC71
Cajamarca
pegmatite
DV65
Cajamarca
metagranite
DV82
Cajamarca
metagranite
DV02
Cajamarca
paragneiss
DV18
Cajamarca
gneiss
DV19
Cajamarca
quartzite
Abejorral
Abejorral
gneiss
Palmitas
Palmitas
gneiss
Amaga
Amaga
granite
La Honda
La Honda
granite
El Buey
El Buey
granite
Manizales
Manizales
granite
GSI1
Santa Isabel
gneiss
GN1
Nechi
gneiss
PALM-1
Palmas
migmatite
Sierra Nevada de Santa Marta
A14
St. M. mylonite
granite
A48
St. M. mylonite
granite
EAM-12-05
St. M. mylonite
granite
Plato-San Jorge Basin
Cicuco-2a
unknown
granite
Cicuco-3
unknown
granite
Lobita 1
unknown
granite
Guajira Peninsular
AVO-03
Uray Gneiss
gneiss
AVO-06
Uray Gneiss
gneiss
Amphibolites
Ecuador
10RC28
Chinchina
amphibolite
11RC04
Monte Olivo
amphibolite
11RC10
Monte Olivo
amphibolite
11RC14
Piedras
amphibolite
JR148
Piedras
amphibolite
Colombia
10RC39
Santa Elena
amphibolite
10RC39A
Santa Elena
amphibolite
10RC50
Tr. Intrusive
amphibolite
AC32B
El Picacho
plagiogranite
CMK040A
El Picacho
meta-gabbro
Padua
Padua
amphibolite

Latitude N-S dm's''

Longitude W dm's''

206

Pb/238U age 2 (Ma)

S 1 23' 51"
S 0 22' 33"
S 4 27' 43"
S 3 9' 24"
S 4 29' 02"
S 3 58' 41"
S 1 23' 57"
N 0 23' 24"

78 21' 15"
78 08' 32"
79 08' 52"
78 48' 45"
79 08' 55"
79 01' 15"
78 22' 08"
77 51' 44"

233.7
234.4
247.2
231.0

S 3 42' 16"

79 51' 07"

S 3 42' 58"

80 03' 18"

S 3 40' 41"

79 54' 14"

N 4 19' 24"
N 5 53' 13"
N 6 01' 08"
N 5 59' 17"
N 5 58' 34"
N 7 00' 56"
N 5 08' 20"
N 5 09' 27"
N 5 07' 34"
N 5 59' 16"
N 4 17' 16"
N 4 46' 42"
N 4 28' 19"
N 4 28' 19"

75 12' 07"
75 25' 28"
75 07' 28"
74 55' 37"
74 54' 02"
75 22' 28"
75 09' 47"
75 07' 57"
74 54' 38"
74 55' 34"
75 13' 59"
74 57' 54"
75 33' 18"
75 33' 18"

MSWD

0.8
0.9
4.3
1.9

1.1
0.8
3.0
2.1

235.0 1.5
207.6 9.2
227.3 2.2

3.0
1.9

237.7
226.0
223.2
229.3
225.7
227.5

5.2
1.3
2.2
2.4
6.5
0.8

4.6

277.6
236.1
234.1
244.6
245.0
236.4

1.6
3.3
1.2
2.4
2.0
1.8

1.2
3.7
1.2
2.3
0.6
3.0

255.7 1.5
236.0 0.6
240.9 1.5
275.8 1.5
238582
236.2 6.3
2311163
250 10#
240 4#
227.6 4.5

1.2
0.9
0.6
3.0

40

Ar/39Ar age 2 (Ma)

214.6 0.9m

221.8 1.0m

213.7 0.9m

0.6

1.4
218.7 0.3b
219.3 0.3m
229.7 0.5h

N 6 57' 34"
N 8 10' 13"
N 6 09' 14"

N 9 16' 25"
N 9 17' 39"
N 9 18' 30"

74 45' 13"
74 46' 55"
75 32' 36"

74 38' 53"
74 38' 52"
74 41' 31"

N 5 03' 05"
N 0 23' 24"
S 1 23' 56"
S 3 39' 9"
nr

75 34' 25"
77 51' 44"
78 22' 52"
79 50' 35"
nr

N 5 54' 06"
N 5 53' 52"
N 6 09' 26"

75 24' 31"
75 24' 37"
75 44' 31"

226.7 1.6
236.4 6.6
222 10#

1.2
2.1

288.1 4.5
276.5 5.1
264.9 4.0

1.0
1.8
0.0

241.6 3.9
241.6 3.9
239.6 2.9

3.9
6.0
0.6

247.6 4.1
245.6 3.9

0.5
0.5

224.7 1.9

0.8

231.9 3.2
222.7 6.3
221 17.0

1.6
1.9

239.7 2.4

1.9

216.6 0.4

0.7
243 4h

Abbreviations: b (biotite), h (hornblende), m (muscovite), wr (whole rock); A/CNK (Molecular Al2O3/CaO + Na2O + K20); (La/Yb)n (normalized to N-MORB)
87
Sr/86Sr 2 s.d. (ext. reproducibility) = 0.0007%; 143Nd/144Nd = b0.0005%; 206Pb/204Pb = 0.12 %.
Dates acquired by LA-ICPMS (Villagmez et al., 2011; Cochrane et al., 2014a), TIMS (Litherland et al., 1994; Aspden et al., 1995), SHRIMP (Vinasco et al., 2006; Restrepo et al., 2011).
Monazite date.
#Date obtained from the youngest zircon when a large spread of zircon ages were obtained due to xenocrystic contamination.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

18O () 2

Th/U zircon 2

A/CNK wr

(La/Yb)n wr

Publication

10.5 to 3.2
11.0 to +3.2
5.3 to 0.5
2.63 0.43

15.3 0.2
15.1 0.2
16.8 0.2

0.26
0.04
0.69
0.24

0.1
0.1
0.5
0.1

1.99
1.40
1.23
1.19
1.37

13.24
13.50
10.58
12.68
4.65

6.0 to +1.7
16.3 to 9.0

12.1 0.2
15.1 0.2

0.14 0.1
0.01 0.0

2.24

6.92

Cochrane et al. (2014a)


Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Litherland et al. (1994)

0.42 0.5

2.38

11.36

Cochrane (2013)
Riel et al. (2013)
Riel et al. (2013)
Riel et al. (2013)
Riel et al. (2013)
Aspden et al. (1995)

16.23
8.19
11.49
12.00
15.70
14.27
12.63
12.81

Cochrane et al. (2014a)


Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Villagmez et al. (2011)
Villagmez et al. (2011)
Villagmez et al. (2011)
Vinasco et al. (2006)
Vinasco et al. (2006)
Vinasco et al. (2006)
Vinasco et al. (2006)
Vinasco et al. (2006)
Vinasco et al. (2006)
Restrepo et al. (2011)
Restrepo et al. (2011)
Restrepo et al. (2011)

Hf zircon 2

Nd w.r 2

(87Sr/86Sr)i wr 2

(206Pb/204Pb)i wr 2

7.5 to +0.8

1.85
0.13
0.10

1.96 0.31
6.57 0.66
9.5 to 0.2
8.2 to +1.4
11.7 to 3.1
5.9 to +3.1

13.6
17.4
13.1
13.1

0.2
0.2
0.2
0.2

15.9 0.2

3.16 0.7
6.0 to +0.4
5.9 to +0.7
3.7 to +0.3

15.6 0.2

1.50

1.27
0.08
0.23
0.35
0.42
0.30

0.6
0.1
0.1
0.1
0.4
0.2

1.10
0.31
0.26
0.66

0.2
0.1
0.2
0.1

1.18
1.73
1.27
1.33
1.36
1.56
1.84
1.70

0.82
0.25
0.30

0.19
0.23
0.24
0.73
0.57

Cardona et al. (2010)


Cardona et al. (2010)
Cardona et al. (2010)

nr
nr
nr

Montes et al. (2010)


Montes et al. (2010)
Montes et al. (2010)

0.20
0.59

Weber et al. (2010)


Weber et al. (2010)

13.31 0.25

9.83

0.70354

17.520938

0.20 0.1

6.3 to +11.2
15.00 0.29

5.03
9.79

0.71470
0.70271

18.707878
17.754038

0.19 0.1
0.32 0.2

8.98
4.13
10.18
3.4
8.4

0.70430
0.70535
0.70243
0.70448

18.119529
18.298843
16.607997

4.8 to +10.0

0.66
0.61
0.63
0.61
0.52

1.41
2.59
1.71
0.81

Cochrane et al. (2014a)


Cochrane et al. (2014a)
Cochrane et al. (2014a)
Cochrane et al. (2014a)
Noble et al. (1997)

0.82
0.62
0.50
0.97
0.61

2.34
2.02
0.49
8.00
0.64

Cochrane et al. (2014a)


Cochrane et al. (2014a)
Cochrane et al. (2014a)
Martnez (2007)
Martnez (2007)
Vinasco et al. (2006)

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

Sample

Unit

Lithology

Jurassic intrusions and volcanic rocks


Ecuador
Eastern Cordillera
09RC57
Azafrn
Granodiorite
09RC59
Azafrn
Monzogranite
09RC65
Rosa Florida
Quartz diorite
09RC22
Abitagua
Granite
09RC60
Abitagua
Granodiorite
09RC61
Abitagua
Monzogranite
09RC43
Zamora
Granodiorite
09RC46
Zamora
Granodiorite
101E
Zamora
Granodiorite
138
DTR64/
69/184
MI90

Latitude N-S
dm's''

Longitude W
dm's''

206

Pb/238U
age 2 (Ma)

MSWD
(U/Pb)

S01 14' 37"


S01 14' 37"
N00 15' 4.0"
S01 14' 47"
S00 24' 33"
S00 22' 19"
S04 24' 0.0"
S03 34' 43"
S03 12' 15"

78 09' 59"
78 09' 27"
77 19' 23"
78 06' 48"
77 28' 51"
77 29' 45"
79 04' 49"
78 30' 59"
78 24' 05"

143.5
140.7
182.4
174.0
173.0
169.8
131.6
178.1

2.2
1.6
0.7
1.8
2.8
2
0.4
3

Zamora

Granodiorite

S03 11' 18"

78 25' 08"

Porphyry

Qtz diorite

S04 01' 28"

78 47' 25"

Misahualli

Andesite

Colombia
Central Cordillera
DV04
Ibagu

Diorite

1.3
0.7
0.6
1.2
1.3
1.1
1.1
1.4

40
Ar/39Ar
age 2 (Ma)

Ndi w.r 2

9.25
8.70
2.84
4.74
4.02
4.89
6.25
6.23

5.04
4.61
0.44
1.78
1.27
1.9
3.28
2.84

0.54(3.6)
0.37(2.1)
0.24(1.5)
0.26(0.9)
0.60(4.1)
0.23(1.6)
0.88(10)
0.23(0.8)

145.4 0.21

Th/U
zircon

A/
CNK
wr

(La/Yb)
n wr

Publication

1.07
1.08
0.9
1.09
1.07
1.11
1.75
1.06

18.05
22.21
14.79
18.93
32.17
19.59
16.82
19.57

0.88

16.99

Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Chiaradia
et al. (2009)
Chiaradia
et al. (2009)
Chiaradia
et al. (2009)
Romeuf et al.
(1995)

0.46
172.3 2.1h

N04 47' 00.2" 74 58' 31.4"

159.2 5.2h
h

Granodiorite

N04 24' 27.7" 75 16' 05.3"

DV06

Ibagu

Granite

N04 24' 08.9" 75 17' 40.3"

182.6 2.4h

DV07

Ibagu

Granite

N04 24' 25.4" 75 18' 04.5"

148.9 3.3h

DV09

Ibagu

Granite

N04 24' 29.7" 75 18' 11.8"

159.6 2.4

0.63

DV129
DV132
DV137
DV138
10RC02
10RC06
10RC08
10RC10
10RC78
10RC03
10RC07
CB0007A

Unnamed
Saldana
Unnamed
Saldana
Ibagu
Ibagu
Ibagu
Ibagu
Segovia
Saldana
Saldana
Las Minas

Granodiorite
Agglomerate
Granite
Rhyolite
Granite
Granite
Granodiorite
Granite
Monzogranite
Rhyodacite
Qtz porphyry
Monzodiorite

N01 10' 02.5"


N01 07' 55.5"
N01 04' 55.4"
N01 06' 45.0"
N04 10' 39"
N04 28' 12"
N04 14' 48"
N04 14' 31"
N06 17' 53"
N04 10' 39"
N04 28' 12"
N02 14' 37.5"

175.8
179.0
173.6
181.5
164.4
168.8
156.5
155.7
188.9
158.5
146.8
187.4

1.7
2.0
1.5
1.6
1.1
0.7
1.1
2.2
2.0
1.0
1.5
2.3

1.3
1.6
2.5
3
0.8
2.5
0.5
3.4
1.5
1.9
0.8
0.62

CB0010

Ibagu

Granite

N02 24' 04.2" 75 53' 46.8"

189.1 2.9

Garzn Massif
CB0001
Garzn

Granite

N02 11' 28.8" 75 35' 19.2"

Monzogranite N02 02' 24.9" 75 45' 34.8"

76 51' 32.7"
76 50' 54.4"
76 48' 34.1"
76 50' 18.6"
75 07' 19"
74 35' 07"
75 09' 47"
75 10' 09"
74 18' 0"
75 07' 19"
74 35' 06"
75 48' 22.2"

(87Sr/86Sr)i
wr 2

153.8 1.5h

Ibagu

Altamira

Hf zircon 2
(MSWD)

160.5 1.7h

DV05

CB0005

Lu-Hf
age 2
(Ma),
(MSWD)

166.0 10.0

0.29

153.1 2.0

151.8 0.9b
1.04 0.20(1.6)
2.90 1.70(48)
1.86 0.51(3.5)
0.01 0.43(2.3)
4.47 0.26(1.1)
2.68 0.19(1.5)
7.65 0.24(1.0)
7.44 0.23(1.4)
2.94 0.31(1.5)
3.76 0.45(2.9)
8.19 0.24(1.3)

1.60
0.32
3.86
3.71
3.69
1.12
4.24
3.32

1.07
1.07
1.04
1.03
1.00
1.04
1.11
0.89

22.47
28.45
8.77
18.35
10.88
12.66

0.59

4.32

0.87

13.35

173.9 1.6

1.5

5.32

179.0 2.2

0.99

2.32

0.91

23.28

30.11

Villagmez et al.
(2011)
Villagmez et al.
(2011)
Villagmez et al.
(2011)
Villagmez et al.
(2011)
Villagmez et al.
(2011)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
Bustamante et al.
(2010)
Bustamante et al.
(2010)
Bustamante et al.
(2010)
Bustamante et al.
(2010)

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

Table 2
Summary of geochronological data collected from Jurassic and Early Cretaceous rocks of the cordilleras of Ecuador and Colombia.

Granodiorite

N07 06' 03"

73 00' 36"

193.8 1.8

2.2

5.47 0.56

4.98 0.14

0.7088

0.77

1.09

29.10

4.44 0.16

0.7082

0.88

1.06

27.25

10VDL22

Pescadero

Granodiorite

N06 49' 48"

72 59' 27"

199.1 1.3

2.5

4.72 0.21

10VDL28

Onzaga

Granodiorite

N06 22' 31"

72 49' 06"

200.4 0.7

1.3

5.21 0.38

0.87

1.01

39.13

10VDL31

Onzaga

Granodiorote

N06 24' 28"

72 49' 08"

201.0 0.9

2.9

5.93 0.55

0.79

1.06

26.52

10VDL32

Mogotes

Granodiorite

N06 25' 22"

72 49' 29"

198.0 0.8

1.3

6.08 0.35

0.81

1.04

32.65

10VDL35

Rio Surata

Diorite

N07 10' 22"

73 05' 08"

201.1 1.4

1.9

5.75 0.57

1.17

1.11

15.73

Pegmatite

N07 10' 46"

72 59' 48"

208.8 1.2

2.2

6.55 0.85

0.03

1.05

22.62

10VDL39

13.06 3.55 0.7216

10VDL52

Paramo Rico

Tonalite

N07 13' 54"

72 53' 54"

199.8 1.2

1.2

4.41 0.20

0.97

1.12

6.23

10VDL54

Ocana

Granite

N08 09' 45"

73 17' 59"

195.8 1.5

1.2

5.01 0.39

0.80

1.23

2.25

10VDL59

Rio Negro

Tonalite

N07 17' 13"

73 08' 46"

196.0 1.1

1.5

6.11 0.43

1.29

1.05

32.12

10VDL61

Rio Surata

Granodiorite

N07 09' 59"

73 05' 17"

200.0 1.5

1.2

5.48 0.55

1.13

1.13

15.15

Early Cretaceous metasedimentary, volcanic and H-MP/LT metamorphic rocks


Ecuador
Eastern Cordillera and the Amotape Complex
09RC12
Alao-Paute
Sandstone
S02 03' 26.0" 78 36' 39.6"
163.7 1.6max
11RC13
Upano
Greenschist
S03 14' 54"
78 41' 39"
121.0 0.8
09RC63
Chingul
Monzogranite N 00 22' 51" 77 17' 39"
125.3 0.9
SEC16-1 Raspas
Blueschist
S03 35' 40.6" 79 55' 30.5"
SEC47-4

Raspas
Raspas
Raspas

Raspas
09PR47
Peltetec
04PR48
Peltetec
Colombia
Central Cordillera
DV176
Quebradagrande
DV20
Quebradagrande

132 5
123.9 1.4P

Metapelite
Metabasalt
Gabbro

129.3 1.3P
134.3 12.8Pl
134.7 0.9Pl

Diorite
Tuff

190B

Barragn

Pajarito

N05 27 '16.0'' 75 28' 28.2''


N04 29' 27.8" 75 34' 02.0"

112.9 0.8
114.3 3.8

N05 03' 13.3" 75 34' 04.6"


N04 16' 24.7" 75 47' 22.2"

149.2 6.1max

Gabbro

4.24

11.2
7.85

2.3
2.00

N05 32' 34.4"


N05 18' 29.3"

74 09' 10.7"

1.14

12.11 0.17(1.4)

0.48
0.56

2.87
3.12

1.13

32.81

112.0 3.7h

72 42' 7.7"

136.0 0.4h
h

120.5 0.6

Cochrane (2013)
Cochrane (2013)
Cochrane (2013)
John et al. (2010)
John et al. (2010)

120.7 0.3m

Schist

Gabbro

0.84
1.05
0.81

Metapelite
Metapelite

Quebradagrande Sandstone
Arqua
Amphibolite

Pj7

8.15 0.45(3.0)

0.7167

126.4 4.0
(4.1)
129.9 5.6
(2.0)

Metapelite

10RC27
DV89b

Eastern Cordillera
Pa5
Pacho

1.4
2.8

7.20 0.09

Van der Lelij


(2013)
Van der Lelij
(2013)
Van der Lelij
(2013)
Van der Lelij
(2013)
Van der Lelij
(2013)
Van der Lelij
(2013)
Van der Lelij
(2013)
Van der Lelij
(2013)
Van der Lelij
(2013)
Van der Lelij
(2013)
Van der Lelij
(2013)

0.70621

5.4

0.70413

Feininger (1980)
Gabriele
(2002)
Gabriele (2002)
This study
This study

Cochrane (2013)
Villagmez et al.
(2011)
Cochrane (2013)
Villagmez et al.
(2011)
Bustamante et al.
(2012)

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

Santander Massif
10VDL05 Corcova

Vsques et al.
(2010)
Vsques et al.
(2010)

Abbreviations: b (biotite), h (hornblende), m (muscovite), max (indicates maximum stratigraphic age), P (phengite), Pl (plagioclase); A/CNK (Molecular Al2O3/CaO + Na2O + K20); (La/Yb)n (normalized to N-MORB).
U-Pb zircon dates acquired by LA-ICPMS (Villagmez et al., 2011; Cochrane, 2013; Van der Lelij, 2013), TIMS Chiaradia et al. (2009).
The 40Ar/30Ar date of Bustamante et al. (2012) does not necessarily date the timing of crystallisation of muscovite.
The Lu-Hf dates are isochron dates obtained from more multiple garnet fractions, pyroxene, amphibole and whole rock.

10

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

2010; Van der Lelij, 2013). Field relationships (Litherland et al., 1994)
suggest that the mesosomal rocks of the Triassic migmatites and
S-type granites within Ecuador are considered to be sedimentary rocks
of the Palaeozoic, fossil bearing Chiguinda and Isimanchi units of the
Cordillera Real (Fig. 2). These sparsely studied sequences yield a detrital
zircon U-Pb age spectrum that has the same age peaks (Fig. 3c; Chew
et al., 2008), although the tectonic setting within which these sequences
were deposited is undetermined. The Brasiliano metamorphic belts
(Cordani et al., 2003) formed during the late Neoproterozoic amalgamation of Gondwana, and may have supplied some detritus to western
South America. However, these belts are located in eastern South
America, and a lack of evidence for detritus being sourced from the
intervening Amazonia Craton suggests that the Brasiliano Orogenic
belts were not a major source region (Chew et al., 2008). Finally,
all magmatism associated with Neoproterozoic extension is mac
(e.g. the Puncoviscana fold belt in northwestern Argentina; Omarini
et al., 1999), which led Chew et al. (2008) to suggest that it is unlikely
that these rocks were a major contributor of zircons to Palaeozoic
sequences along western South America.
4.2.2. Comparison with the ages of Permian and Triassic rocks in Venezuela
and Peru
Van der Lelij (2013) report concordant zircon U-Pb dates (LAICPMS) from four granitoid intrusions and a dacitic lava from the Merida
Andes of Venezuela (Fig. 1). These dates range between 202.0 1.8 Ma
(La Quinta Fm.) and 243.5 3.4 Ma, and overlap with dates obtained
from Colombia and Ecuador (Fig. 3d). No Permian concordant zircon
U-Pb dates have been reported from the Merida Andes. Rhyolites and
granites of the El Baul massif in Venezuela yield zircon U-Pb dates that
span between 283.3 2.5 Ma and 291.1 3.1 Ma (Viscarret et al.,
2009), and a zircon U-Pb age of 272.2 2.6 was obtained from a granitic
intrusion in the Paragauana Peninsula (Van der Lelij, 2013).
Voluminous, partly migmatised Late PermianTriassic magmatic
intrusions are exposed throughout the southern and central Eastern
Cordillera of Peru (Mikovi et al., 2009). Zircon U-Pb dates range
between 223 and 285 Ma (Fig. 3d; Mikovi et al., 2009; Reitsma, 2012),
with a peak at 240260 Ma. The crystallisation ages show a southward
younging trend, and the oldest plutons south of 11.5S are younger than
245 Ma. The Mitu Group of the central and southern Eastern Cordillera
of Peru hosts abundant Triassic sedimentary and volcanic sequences.
Volcanic tuffs in the south yield concordant zircon U-Pb (LA-ICPMS)
dates ranging between 234.3 0.3 Ma and 238.7 1.8 Ma (Mikovi
et al., 2009; Reitsma, 2012), and Chew et al. (2005) report a zircon
U-Pb age of 219.7 1.8 Ma from a rhyolite of the Mitu Group in central
Peru. Detrital zircons extracted from oxidised terrigeneous sedimentary
rocks of the Mitu Group yield minimum crystallisation dates ranging
between 217.2 4.1 Ma and 250.7 4.9 Ma, which constrain their
maximum stratigraphic ages (Reitsma, 2012). Several authors propose
that the Mitu Group was deposited within a rift (Mgard, 1978;
Laubacher et al., 1988; Reitsma, 2012), and Reitsma (2012) suggests
that the rift formed in a back-arc basin setting.
Romero et al. (2013) recently published a concordant zircon U-Pb
age of 243 0.1 Ma from a basalt exposed in Macab Island offshore
northern Peru (~8S).
4.3. Geochemistry of the granites and migmatites
4.3.1. Cordillera Real of Ecuador and Cordillera Central of Colombia
Major oxide, trace element and Rare Earth Element (REE) abundances and oxygen isotope compositions (Table 1) have been obtained
from Permian and Triassic granites and migmatitic leucosomes from
Colombia (Vinasco et al., 2006; Martnez, 2007; Cardona et al., 2010)
and Ecuador (Litherland et al., 1994; Cochrane et al., 2014a). These
rocks span the boundaries of calcic and alkali-calcic differentiation
trends on the modied alkali-lime index of Peacock (1931; Fig. 4a),
with a compositional range of 6278 wt% SiO2. The same rocks plot

within the high-K calc-alkaline and calc-alkaline elds when comparing


SiO2 with K2O (Fig. 5a). The Triassic anatectites have strongly peraluminous Aluminium Saturation Indices (ASI 0.972.38; calculated
using Maniar and Piccoli, 1989; Fig. 4b), while the Permian granitoids
tend to cluster at slightly lower peraluminous and mildly metaluminous
values (ASI 0.921.73, with a majority b 1.1). The Triassic anatectites
yield elevated 18O quartz (Fig. 4c), which along with their high ASI indices places these granites and leucosomes within the S-type granite
eld of Chappell and White (1974) and Harris et al. (1997). N-MORB
normalized trace element abundances of Triassic granitoids from
Ecuador and Colombia are identical (Fig. 4d), suggesting that there are
no signicant along-strike changes in the fractionation and assimilation
history of these high-SiO2 melts. The trace elements are enriched in
Light Ion Lithophile Elements (LILE), and negative Nb and Ta anomalies
are present in both the Permian and Triassic granitoids (Fig. 4e, f), suggesting that a subduction-derived component was incorporated into
these rocks. The Triassic granitoids yield slight negative Ba, Eu, Sr and
Ti anomalies, which suggest that plagioclase and Fe-Ti oxides have fractionated, and a positive Pb anomaly that may be derived from a
protolith within the continental crust. In contrast, the Permian granites
do not yield Ba, Eu and Sr anomalies, although they do have negative Ti
anomalies, suggesting that they evolved via a different fractionation
scheme. Trace element concentrations normalized to the composition
of average upper continental crust (Taylor and McLennan, 1995) plot
close to unity, corroborating the S-type character of these rocks
(Fig. 4f). REE abundances in the Triassic granites and leucosomes normalized to N-MORB reveal light-REE enrichment with (La/Yb)n ranging
between 2.3 and 19.8, with a mildly positive correlation with 206Pb-238U
crystallisation age (Fig. 4h). (La/Yb)n ratios from Permian granites yield
a larger range of 8.855.1, and the REE concentrations have a larger
range relative to N-MORB, compared to the Triassic rocks (Fig. 4g).

4.3.2. Comparison with Permian and Triassic rocks in Venezuela and Peru
Migmatised granitoids within the Eastern Cordillera of Peru that
crystallised during 285223 Ma are high-SiO2 granites, mildly metaluminous to peraluminous (ASI 0.91.1; Fig. 4b), and yield K2O/Na2O
ratios that mainly range between 0.8 and 1.2 (Mikovi et al., 2009).
Mikovi et al. (2009) report U-Pb zircon ages from monzogranites
in southern Peru (Cordillera de Carabaya) which range between 190
and 216 Ma. These intrusions are geochemically distinct from the
older Permo-Triassic group (223285 Ma) because they are strongly
peraluminous (ASI 0.981.42; Fig. 4b), and yield anomalously high
whole rock K2O/Na2O ratios (0.552.22, with a majority N 1.20). Geochemically, the Permian and Triassic migmatites and granites of the
Cordillera Real of Ecuador and the Cordillera Central of Colombia
(K2O/Na2O 0.772.93; ASI 0.922.38) resemble the Late Triassic granites
of the Eastern Cordillera of Peru. Mikovi et al. (2009) combined these
major element characteristics with iron oxide number, SiO2 (e.g. Frost
et al., 2001) and trace element abundances, to classify the PermoTriassic (285223 Ma) and Late Triassic plutons as late- to postorogenic.
The shift from a well characterised, calc-alkaline Carboniferous arc in
Peru to Permian post-tectonic alkali feldspar granites (Mikovi et al.,
2009) and ultimately alkaline bimodal volcanic rocks of the Mitu
Group is characteristic of lithospheric thinning (e.g. Xu et al., 2007).
Mikovi et al. (2009) suggest that the Permo-Triassic granitoids
(223285 Ma) formed by dehydration melting of the lower crust during
basaltic underplating (e.g. Sisson et al., 2005), which was driven by decompression subsequent to break-off of the Carboniferous slab. Granitoid intrusions and rift-related magmatism (Dalmayrac et al., 1980)
initiated in the Middle Triassic, forming the Mitu Rift and bimodal volcanic rocks of the Mitu Group (Reitsma, 2012). Finally, the highly
peraluminous Late Triassic granites formed by melting of the fertile uppermost crust, consisting of an igneous protolith and a substantial sedimentary component.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

4.4. Geochemistry of the amphibolites


A comparison of the abundance of K2O and SiO2 (Fig. 5a) in all of the
magmatic rocks that yield Triassic ages reveals the bimodal nature of
magmatism within northwestern South America as part of Pangaea between 216.6 0.4 Ma and 243 4 Ma (Fig. 3d). Amphibolitic dykes and
massive metagabbros from the Cordillera Real and Amotape Complex of
Ecuador (Chinchina, Monte Olivo and Piedras units) and the Cordillera
Central of Colombia (Santa Elena, Padua and Aburr) yield low K2O
(b 0.5 wt%) relative to SiO2 (4655 wt%), placing them within the tholeiitic eld of Peccerillo and Taylor (1976; Fig. 5a). However, a comparison of the immobile elements Th and Co (Fig. 5b) suggests that
amphibolitic dykes straddle the tholeiite and calc-alkaline elds, while
the massive metagabbros of the Aburr Ophiolite plot in the tholeiite
eld. These discrepancies suggest that the amphibolitic dykes may be
partially altered, although this was not visible in hand specimen or in
thin sections. The amphibolites and metagabbros have values of Zr/
TiO2 that are lower than 0.01 (Fig. 5c), placing them within the subalkaline basalt eld, implying that the tholeiitic nature is primary. The
amphibolitic dykes are enriched in Ti relative to V (Fig. 5d), and plot
in the MORB or back arc basin basalt (BABB) eld of Shervais (1982).
However, the massive metagabbros of the Aburr Ophiolite (El Picacho
metagabbros) that yield very low K2O abundances of 0.020.12 wt%
(Fig. 5a), plot closer to the arc eld. LILE abundances within the
amphibolitic dykes from Colombia and Ecuador (Fig. 5e) are enriched
(up to ~100 times) relative to N-MORB and lack signicant Nb and Ta
anomalies. The HFSE elements plot close to parity with N-MORB,
which is consistent with the tectonic discrimination plots. Similarly,
N-MORB normalised REE plots (Fig. 5f) for the amphibolitic dykes plot
close to MORB compositions, although the LREE are slightly enriched
with (La/Yb)n ratios varying between 0.59 and 3.16, while the massive
metagabbros of the Aburr Ophiolite yield approximately at REE patterns that are slightly depleted relative to N-MORB. N-MORB normalised La/Yb ratios from all of these rock sequences show a progressive
reduction with crystallisation age from 243 4 Ma to 216.6 0.4 Ma
(Fig. 5g). Finally, whole rock Ndi values for the amphibolitic dykes
and the metagabbros range between 3.40 and 10.18, and become
more juvenile with younger crystallisation ages (Fig. 5h). The most juvenile rocks are characteristic of MORB and BABB isotopic compositions.
A single amphibolite from the Monte Olivo unit in the Cordillera Real
of Ecuador yields a whole-rock 87Sr/86Sri of 0.7147 (Fig. 5i), which is extremely high relative to its 143Nd/144Ndi of 0.5126 and low La/Yb ratio of
1.71 (Cochrane et al., 2014a). This is consistent with low temperature
alteration, which has preferentially mobilized the LILE but had a minimal effect on the REE.
4.5. Zircon Hf isotope geochemistry
Cochrane et al. (2014a) report Hf isotopic compositions from zircons
extracted from eighteen migmatitic leucosomes, more massive granitoids and amphibolitic dykes throughout the Cordillera Real of Ecuador
and the Cordillera Central of Colombia (Table 1). The zircons, which
have been dated by LA-ICP-MS (U-Pb), yield a large range of weighted
mean Hfi values of +15 and 20 (Fig. 6), which are consistent with
crustal recycling and the addition of new continental crust (Collins
et al., 2011; Cochrane et al., 2014a).
4.5.1. Zircon Hf isotope geochemistry of the granites and
migmatitic leucosomes
Single leucosomes of migmatites and peraluminous granites generally yield high, intra-sample variations (e.g. Hfi + 3 to 11; Fig. 6a)
within coeval magmatic rims that surround variably aged xenocrystic
cores (Cochrane et al., 2014a), and within samples that lack older
cores. These variations are too large for magmatic zircons that
crystallised from a single, well-mixed source (e.g. Gerdes et al., 2002),
and Cochrane et al. (2014a) conclude that they are mainly derived

11

from multiple crustal sources, while some variation found in the


xenocrystic bearing zircons may be due to the fractionation of Hf isotopes between cores and overgrowths during melting (e.g. Gerdes and
Zeh, 2009). In contrast, a small proportion of granites yield Hfi (zircon)
values that statistically dene a single population, suggesting that they
were derived from a distinct, homogeneous source. No correlation is
found between the Hfi values obtained from the rims of xenocrystbearing and xenocryst- free zircons, and crystallisation age, which is
not surprising given the heterogeneous nature of the source rocks within the crust. Hfi values obtained from xenocrystic zircon cores (Fig. 6a)
span a large range and are representative of the sedimentary protoliths
that melted to form the anatectites. The large heterogeneity in Hfi corroborates the large range in 238U/206Pb dates of the protoliths.
4.5.2. Zircon Hf isotope geochemistry of the amphibolites
Cochrane et al. (2014a) report Hfi (zircon) values from four amphibolites which show a negative correlation with crystallisation age
(Fig. 6b). The two older amphibolitic dykes (240232 Ma) yield a
large range in Hfi, with juvenile values (7.4 to 10) obtained from
patchy or unzoned (cathodoluminescence) zircons, and crustal compositions (3.6 to 4.8) from zircons that exhibit oscillatory zoning, similar to the zircons extracted from the anatectites (Fig. 6c). The two
youngest amphibolites (225223 Ma) dene single populations of Hfi
(13.315.0) from unzoned zircons, which approach the depleted mantle
array. Crustal contamination of the mac melts during emplacement
was an important process in the petrogenesis of the older amphibolites
prior to ~225 Ma. However, there is no evidence for the assimilation of
signicant continental crust after ~225 Ma. This interpretation is supported by a negative correlation between Hfi (zircon) and (La/Yb)n,
(Fig. 6b) suggesting that the mac dykes trend towards MORB compositions with a depleted mantle source, during 240223 Ma.
4.5.3. Comparison with Zircon Hf isotope compositions in Peru
Hf isotopic compositions obtained from Permian and Triassic
peraluminous granitoids of the Eastern Cordillera of Peru yield no
clear trends with time, and Hft values range between 6 and 8
(Fig. 6d; Mikovi and Schaltegger, 2009). This range overlaps with
that obtained from Permian and Triassic granitoids of Ecuador and
Colombia, and individual, intra-sample spot analyses show a large
spread, reecting the heterogeneity of the source rocks. Mean zircon
Hfi values from the Early and Late Triassic of 0.02 1.56 and
1.96 1.56 suggest that this period was characterised by the addition of isotopically juvenile, mantle derived magmas, which were underplating previously attenuated continental crust (Mikovi and
Schaltegger, 2009). Calculated crustal Hf model ages for the granitoids
within the Eastern Cordillera of Peru range between 1.4 and 1.0 Ga
(Mikovi and Schaltegger, 2009), suggesting that the source rocks
were basement that formed within the Sunsas Orogeny during the
amalgamation of Rodinia.
4.6. Thermal histories during the Triassic
Numerous K/Ar dates have been obtained from Palaeozoic and
Triassic magmatic rocks in the Cordillera Real of Ecuador (Litherland
et al., 1994) and the Cordillera Central of Colombia (Feininger et al.,
1972; Hall et al., 1972; McCourt et al., 1984; Aspden et al., 1987).
However, these dates cannot be unambiguously interpreted as
crystallisation ages given the potential for isotopic disturbance by
thermally activated diffusive loss and uid assisted loss of the daughter
isotopes. Within Ecuador, muscovite, biotite and whole rock K/Ar dates
of the Tres Lagunas Granite and Sabanilla unit (migmatites) range
between 100 and 50 Ma (Litherland et al., 1994), revealing a clear
disturbance to the isotopic system. Unfortunately, we cannot extract
useful timeTemperature (tT) information from these data because
the degree of daughter isotope loss cannot be quantied. Nevertheless,

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

76

Jurassic
Metalluminous granitoids
(mainly granodiorite).
Continental arc intrusions
Triassic
Peraluminous granites and migmatites
(e.g. Tres Lagunas Granite, Sabanilla
Migmatite)
Amphibolites and ultramafic rocks
(e.g. Piedras unit, Monte Olivo unit)
Palaeozoic - Triassic
Undifferentiated para- and
ortho-, schists and gneisses
(Ecuador: Agoyn, Chiguinda, Piuntza
units.
Colombia: Cajamara Unit)

244.62.4

U-Pb zircon, monazite


LA-ICPMS, intrusions

277.33.0

U-Pb zircon
SHRIMP, intrusions

233.74.8 - 2.6 Ga

236.46.6
8N

OPF

236.41.8
Aburr Ophiolite

216.60.4
22210*
2404*
227.64.5
218.70.3
219.30.3
239.72.4
236.13.3
221.81.0
25010*
238 - 2800
224.71.9
Chinchina Stock

Santa Elena
Amphibolites
M

226.71.9
234.11.2
244.62.4
245.02.0*
213.70.9
240.91.5
227.64.5
2434
255.71.5
236.00.6

IF

236.26.3

U-Pb zircon, LA-ICPMS,


detrital zircons,
metasedimentary rocks

220 - 600
Cajamarca Amphibolites

229.70.5
220 - 1200

4N

CAF

40Ar/39Ar

213.70.9

plateau date
hornblende, muscovite,
biotite, intrusions

247.24.3*

Also dated by apatite


U-Pb for t-T analysis

277.61.6 275.81.5

78W

City
0

74

75

SJF

12

2N

Colombia

100 km
Monte Olivo
Amphibolites
1N

P
ue

Tres Lagunas
Granite

234.40.9

ed
lat
re

ag

Ib

207.69.2

PF

Ecuador

235.01.5
233.70.8

79W

227.32.2
231.93.2
LF

Moromoro Migmatite
& Piedras Amphibolite

222.76.3
22118
223-229
237.75.2
214.60.9

80W

227.50.8
247.24.3*
2345 - 2600

PF

2S

BF

231.01.9
3S

Piuntza
Unit
4S
Sabanilla
Migmatite
Zumba
Ophiolite
5S

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

McCourt et al. (1984) and Litherland et al. (1994) interpret these data to
reect continental collision at ~120 and 6555 Ma.
Spikings et al. (2000, 2001) and Villagmez and Spikings (2013)
present 40Ar/39Ar (white mica, biotite, alkali feldspar) and ssion track
(zircon, apatite) data from Triassic magmatic rocks in Ecuador and
Colombia, which were used to construct tT paths. However, the collision of the Caribbean Large Igneous Province with South America at
~75 Ma drove more than 350 C of cooling, and Triassic thermal histories were not preserved in the isotopic systems.
Recently, Cochrane et al. (2014b) published apatite U-Pb data from a
Triassic leucosome of the Sabanilla Unit (migmatite 09RC42; Table 1) of
southern Ecuador. The authors demonstrate that Pb was lost from the
apatite grains by thermally activated diffusion, and thus the dates can
be combined with grain sizes and the diffusion properties of Pb in apatite to generate a series of plausible tT paths (Fig. 7a) using a computed
Monte Carlo algorithm, at temperatures N350 C. Those paths reveal
rapid cooling subsequent to anatexis at ~250 Ma. The leucosome subsequently remained at temperatures lower than the Pb Partial Retention
Zone throughout the Triassic. The same method has been applied to a
peraluminous Triassic granite (10RC43; Table 1; Fig. 7b) from the Cajamarca Complex of central Colombia. Similarly, the computed tT paths
also reveal rapid cooling subsequent to Triassic anatexis, after which
the rock was colder than the apatite Pb Partial Retention Zone. Data
from the latter sample are new, although the methodology is identical
to that presented in Cochrane et al. (2014b), and the data are presented
in a supplementary le. Rapid cooling during the Triassic corroborates
the indistinguishable U-Pb dates from apatites with a large range in
grain size (Fig. 7b).
Rapid cooling from anatectic temperatures to less than ~ 380 C is
probably mainly a consequence of thermal relaxation subsequent to
the removal of the heat source at a local geographic scale. Some component of cooling may also be a consequence of exhumation, and given
that the samples remained colder than ~380 C throughout the remainder of the Triassic and the Jurassic, it is likely that these particular samples were at depths of 15 km within the crust after ~ 220 Ma. Rapid
cooling during ~250220 Ma temporally coincides with a signicant increase in zircon Hfi obtained from the amphibolitic dykes (Fig. 7a, b).
4.7. Interpretation: Permian and Triassic
4.7.1. Arc magmatism and metamorphism during 290240 Ma along
western Pangaea
The exposure of Permian and early Triassic (290240 Ma) granitoids
within the Northern Andes is restricted to the Sierra Nevada de Santa
Marta, Guajira Peninsula and the central Cordillera Central within
Colombia, while almost all intrusions within the Cordillera Real and
Amotape Terrane of Ecuador are 240 Ma. The Aluminium Saturation
Index of the alkali-calcic to calcic granites that crystallised during
290240 Ma straddles the peraluminous and metaluminous elds,
while the d18O values of 14%17 suggest that these rocks formed by
partial melting of sedimentary rocks. Zircons yield magmatic Th/U ratios
of 0.261.27, suggesting that they have not recrystallized during subsequent metamorphic events (Table 1; Fig. 8). The whole rock trace element abundances are characteristic of subduction related magmatism.
Magmatism during this period was not accompanied by mac dyke emplacement, and all rocks yield N58 wt% SiO2. The 290240 Ma granites
within the Sierra Nevada de Santa Marta (Cardona et al., 2010) and
the Cordillera Central of Colombia (Villagmez et al., 2011) are
interpreted to have formed above an east dipping Pacic subduction
zone beneath Pangaea (Cochrane et al., 2014a; Fig. 8).

13

Plate reconstructions of western Pangaea during the Permian to


Early Triassic (Elas-Herrera and Ortega-Gutirrez, 2002; Weber et al.,
2007) juxtapose Acatln, Oaxaquia and the Chortis Block against
north-western South America (Fig. 9). At the present time, Oaxaquia is
considered to underlie central and southern Mexico, including the
Maya Block. The Maya Block hosts extensive, undeformed Permian
granites, and deformed and foliated Permian granitic gneisses and
migmatites (Chiapas Massif; e.g. Solari et al., 2008), which underwent
rock uplift and erosion during the early Triassic (Schaaf et al., 2002).
The undeformed and deformed Permian granites yield concordant UPb (zircon) dates ranging between 289 and 255 Ma (Yanez et al.,
1991; Solari et al., 2001; Elas-Herrera and Ortega-Gutirrez, 2002;
Ducea et al., 2004; Weber et al., 2007; Kirsch et al., 2012;
Ortega-Obregon et al., 2013; Kirsch et al., 2014) and are considered to
form part of a continental arc (e.g. Torres et al., 1999). We suggest
that the Permian, peraluminous granitoids exposed within northwestern South America formed within the same tectonic regime, and
are a continuation of the Permian belt that is exposed in southern
Mexico (e.g. Centeno-Garcia and Keppie, 1999; Dickinson and Lawton,
2001; Kirsch et al., 2014). Remnants of Permian magmatism have also
been found within the Sierra de Perij (Dasch, 1982), Paraguana Peninsula (Van der Lelij, 2013) and the El Baul Massif in Venezuela (Viscarret
et al., 2009). The Permian granites and migmatites exposed in Colombia
formed in a different tectonic regime to similar lithologies that formed
during the Triassic, which was dominated by extension (see next
section).
Weber et al. (2007) report concordant zircon U-Pb dates from
migmatites of the Maya Block (Chiapas Massif) of 251.8 3.8254.0
2.3 Ma, which they interpret to be a result of MP-HT metamorphism
during compression, and stacking within an orogenic wedge. The collision event post-dates the amalgamation of Pangaea, which is recorded
along the diachronous Ouachita-Marathon suture (Fig. 9) that had
formed by the Early Permian. Weber et al. (2007) attribute the compressional event to closure of a marginal basin, which had previously formed
during extension that accompanied Permian arc magmatism. Similarly,
Cardona et al. (2010) attribute anatexis at ~250 Ma in the Sierra Nevada
de Santa Marta (Fig. 1) to a compressional event. The cause of compression has been attributed to either terrane accretion, subduction of thickened, topographically prominent and buoyant oceanic lithosphere (e.g.
Weber et al., 2007), or increased plate coupling during the waning
stages of the amalgamation of Pangaea (Cardona et al., 2010).
The high-SiO2, Permo-Early Triassic granites and alkali feldspar
granites of the Eastern Cordillera of Peru are interpreted to have formed
within a continental arc setting during extensive lithospheric thinning
(Sempere et al., 2002; Mikovi et al., 2009). Roll-back is considered to
have driven anti-clockwise rotation of the forearc, dextral displacement
and southward younging of the onset of extension along the Peruvian
arc, all of which commenced at ~280 Ma (Mikovi et al., 2009). Permian arc magmatism in Peru is consistent with arc magmatism that is
sporadically preserved within the Northern Andes, and more abundantly within the conjugate margin that is now exposed in southern Mexico.
Weber et al. (2007) propose that the Permian arc intrusions within the
Maya Block were also emplaced during lithospheric thinning. However,
their interpretation is based on the tectonic-switching model of Collins
(2002), who state that a thinned, hot lithosphere is a prerequisite for
subsequent crustal thickening and anatexis within a period of ~20 My.
Geochemical and isotopic evidence from Permo-Triassic intrusions in
Peru provides no evidence for compression at ~250 Ma (Mikovi et al.,
2009). Coeval compression and extension along different regions of the
western margin of Gondwana, within Pangaea, spatially correlates with

Fig. 2. Geology of the Cordillera Real and Amotape Complex of Ecuador, and the Cordillera Central of Colombia, showing the distribution of Palaeozoic and Triassic rocks. The Jurassic
continental arc is also shown for reference. Concordant Permian and Triassic zircon U-Pb and plateau 40Ar/39Ar dates and their uncertainties (2) obtained by various analytical methods
(see Table 1) are shown (see references in Table 1). The Palaeozoic and Triassic rocks within the Cordillera Real of Ecuador are grouped together within the Loja Terrane by Litherland et al.
(1994). Cities, I: Ibagu, L: Loja, M: Medellin, P: Pasto, Q: Quito. Faults. Faults: BF: Baos Fault, CAF: Cauca-Almaguer Faults, LF: Llanganates Fault, OPF: Ot-Pericos Fault, PF: Peltetec Fault.
Map compiled from Litherland et al. (1994) and Gmez et al. (2007).

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

14

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

A Frequency
9
Triassic

Permian

7
6
5

Anatectites
Colombia (n = 27)

Ecuador (n = 14)

3
2
1
0
200

225

275

250

300

206Pb/238U age (Ma)

Frequency
100

Palaeozoic metasedimentary

Frequency units (melanosome)


9 480 - 630 Ma
960 - 1200 Ma

Triassic anatectite
Colombia (n = 360)

Ecuador (n = 288)

Permian anatectite
Colombia (n = 104)

N = 99

10
4
3
2

0
360
420
480
540
600
660
720
780
840
900
960
1020
1080
1140
1200
1260
1320
1380
1440

1180

1120

1060

1000

880

940

820

760

700

640

520

580

460

400

340

280

220

206Pb/238U age (Ma)

206Pb/238U age (Ma)

Zircon U-Pb
age (Ma)2

Ecuador

300

Colombia

Venez Peru
EBM

SNSM

280
270

Permian

290
PP

260
GP
250

MA

230
220

Triassic

240

210

Metagranite,
migmatite

200

Amphibolite,
plagiogranite

190

South

Latitude (degrees)

10

12

North

Fig. 3. A) 206Pb/238U age histogram for the time of anatexis or metamorphic zircon growth for granites, metagranites and migmatites (leucosomes) in Ecuador and Colombia. B) 206Pb/238U
age histogram for Permian and Triassic S-type granitoids and migmatites (leucosomes) from the Cordillera Real (Ecuador) and the Cordillera Central (Colombia). Ages are single spot
zircon ages determined using LA-ICPMS (Villagmez et al., 2011; Cochrane et al., 2014a), SHRIMP and SIMS (Vinasco et al., 2006; Chew et al., 2008; Restrepo et al., 2011). C) 206Pb/
238
U age histogram for detrital zircons from the Palaeozoic Chiguinda and Isimanchi metasedimentary units of the Cordillera Real of Ecuador (Chew et al., 2008). D) A comparison of
Permian and Triassic concordant zircon and monazite U-Pb dates with latitude along the Cordillera Real of Ecuador, Cordillera Central, Guajira Peninsula and the Sierra Nevada
de Santa Marta of Colombia. The ranges of concordant zircon U-Pb dates obtained from granitoid intrusions and volcano-sedimentary rocks from Venezuela (Van der Lelij, 2013) and
the Eastern Cordillera of Peru (Mikovi et al., 2009; Reitsma, 2012) are shown for comparison. Data and citations are presented in Table 1. EBM: El Baul Massif, GP: Guajira Peninsula,
MA: Merida Andes, PP: Paraguana Peninsula, SNSM: Sierra Nevada de Santa Marta.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

Alkalic

2
0

l
lka

i-

l
ca

cic

l
Ca

c-

l
lka

ne

55

S-type
Lachlan FB

Calcic

60

2.0
1.5

Peru
Late Triassic

65

70

75

1.0

1.5

2.0

1.5

2.5

3.0

0.5

10

F
10

Ecuador - Triassic
Colombia - Permian

1000

Ecuador - Triassic
Colombia - Triassic

S-type

Metaluminous

1000

I-type

Peraluminous

Al/(Ca+Na+K)

80

2.0

1.0

Peru
Perm.- Mid. Triassic

0.5
0.5

SiO2 (wt%)

2.5

1.0

-2
50

Metalluminous

3.0

Peru
Late Triassic

Al/(Ca+Na+K)

Ecuador - Triassic

Peraluminous

3.5

Al/(Na+K)

Na2O + K2O - CaO (wt%)

4.0

Peru
Colombia - Triassic Perm.- Mid.
Colombia - Permian Triassic

11

12

13 14 15
18O (quartz)

16

17

18

Ecuador - Triassic
Colombia - Permian

100
100
10
10

0.10 N-MORB

Cs RbBa Th U Nb Ta La Ce Pb Pr Sr Nd Zr Hf SmEu Ti Tb Y TmYb 0.1

H
G

280

N-MORB
Rb

Ba

10 0

Th

Nb

La

Ce

Pb

Sr

Zr

Ecuador - Triassic
Colombia - Permian

270

Peru
Perm.- Mid. Triassic

Ecuador

260
250
240
230
220

0.1

0.01

Colombia

206Pb/238U

10

Ti

age2

0.1

UCC
Cs Rb Ba Th U Nb Ta La Ce Pr Sr Nd Zr Hf SmEu Ti Tb Y TmYb

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

N-MORB
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Yb Lu

210

Peru
Late Triassic

10

12

14

16

18

20

(L /Yb)

Fig. 4. Geochemical data from Permian and Triassic granites and migmatitic leucosomes from the Cordillera Real (Sabanilla and Tres Lagunas units), Amotape Complex (Moromoro unit) and the Cordillera Central (Cajamarca unit). Fields for PermoTriassic intrusions within the Eastern Cordillera of Peru are from Mikovi et al. (2009). Multi-element plots are normalised to N-MORB (Sun and McDonough, 1989) and Upper Continental Crust (UCC; Taylor and McLennan, 1995). Data from
Ecuador: Litherland et al. (1994), Cochrane et al. (2014a). Data from Colombia: Vinasco et al. (2006), Martnez (2007), Cardona et al. (2010), Cochrane et al. (2014a).
15

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

A
10

C
SiO2 wt%

V (ppm)

60

500

Shoshonitic

60

65

70

75

SiO2
0.01
0

10

20

30

Tholeiite

40

50

70

F
3.5

Ti/
V

OIB

0.01

00
Ti/V=1

100

0.1

N-MORB

10

Ti/1000 (ppm)

Zr/TiO2

Ecuador (dykes, sills)


Colombia (dykes, sills)
Colombia (Aburr Ophiolite)

3.0
2.5

10

200

2.0
1.5
1

1.0
0.5

0.1

N-MORB

0.0

Cs Rb Ba Th U Nb Ta La Ce Pb Pr Sr Nd Zr Hf Sm Eu Ti Tb Y Tm Yb

La

Ce

Pr

Nd

Sm

Eu

Gd

Tb

Dy

Ho

Er

Yb

Lu

230

220

240
8
230

(La/Yb)n

220

210

210

Altered
amphibolite

Nd

age (Ma)

N-MORB

240

206Pb/238U

206Pb/238U

age (Ma)

12

Nd

10

12

0
0.700

0.704

0.708

0.712

0.716

87Sr/86Sri

Fig. 5. Geochemical data from Triassic amphibolitic dykes and metagabbros of the Cordillera Real (Monte Olivo unit), Amotape Complex (Piedras unit) and the Cordillera Central (Cajamarca unit and the Aburr Ophiolite). Multi-element plots are
normalised to N-MORB (Sun and McDonough, 1989). Data from Ecuador: Litherland et al. (1994), Cochrane et al. (2014a). Data from Colombia: Martnez (2007), Cochrane et al. (2014a).

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

E
100

50

=
Ti/V

45
0.001

80

=2

50

Co ppm

Ecuador (dykes, sills)


Colombia (dykes, sills)

300

Bas/Trach/
Neph

60

MORB
BABB

alt

Sub-alkali
basalt

Basalt
Gabbro

55

0.1

Phonolite

li bas

Tholeiitic

55

Calc-alkaline

,
andesite
Basaltic , diorite
andesite

Calc-alkaline

50

400

0
45

ARC

Andesite

High-K
calc-alkaline

High-K
calc-alkaline

Colombia (dykes, sills)


Colombia (Aburr Ophiolite)

10

10 Ecuador (dykes, sills)

Alka

B
Th ppm

Ti/V
=

16

K2 O

dacite,
Rhylite, granodiorite
granite,

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

the location of terranes that were outboard of South America (Fig. 9).
This supports a hypothesis that compression at ~ 250 Ma was a consequence of the accretion of terranes that now form the basement of
Central America. Alternatively, varying stress regimes may be a consequence of varied displacement at the oceancontinent plate interface
to the north and south of the Huancabamba deection.
4.7.2. Initiating the disassembly of western Pangaea during 240200 Ma
The crustal anatectites of the Cordillera Central of Colombia and the
Cordillera Real of Ecuador that formed during 240225 Ma yield lower
whole rock (La/Yb)n (Fig. 8), lower zircon Th/U, and are signicantly
more peraluminous than the Permian granitoids in Colombia, the Eastern Cordillera of Peru (Mikovi et al., 2009), and the Maya Block of
southern Mexico (Weber et al., 2007). These data are indicative of increased partial melting of pelitic rocks after 240 Ma, forming metamorphic zircon (zircon Th/U values of b ~0.1; Rubatto, 2002; Hartmann and
Santos, 2004; Fig. 8) within the anatectites, and perhaps an overall
greater volume of magmatism compared to prior to 240 Ma, which is
reected by a peak in the quantity of dated samples during 240
225 Ma (Fig. 3a). Cochrane et al. (2014a) suggest that mac underplating elevated the geothermal gradient, driving uid expulsion from
the pelitic protoliths and lowering their solidus.
The trace element content of the amphibolitised tholeiitic dykes is
characteristic of a back-arc basin or MORB setting, and the progressive
trend in isotopic compositions towards the depleted mantle, combined
with progressive depletions in incompatible elements during 240
225 Ma suggests that they were emplaced within a lithosphere that
was thinning. Hafnium isotopic compositions in zircons shows that
mantle derived tholeiites emplaced during 240232 Ma assimilated isotopically evolved continental crust (Figs. 6 and 8), whereas there is little
evidence of the assimilation of signicant crust after 225 Ma, when Hfi
approaches depleted mantle compositions.
The geographically widespread occurrence of coeval tholeiitic,
amphibolitised dykes and crustal anatectites supports an extensional
setting, which probably formed within a region of increased heat ow
that is characteristic of an attenuated back-arc basin (e.g. Collins and
Richards, 2008). Cochrane et al. (2014a) suggest that the period between 240 and 225 Ma was dominated by progressive thinning of the
continental lithosphere during rifting and the disassembly of western
Pangaea (Fig. 8). Mac underplating and anatexis occurred because of
doming and decompression of the asthenosphere during extension,
and heat convection during the intrusion of mac magmas into the
crust (Fig. 9). This interpretation is supported by i) the tentatively
mapped Triassic Zumba Ophiolite in the southernmost Cordillera Real
of Ecuador, although its Triassic age is assumed based on its structural
position (Fig. 1; Litherland et al., 1994), and ii) late Middle to Upper
Triassic continental and marine volcano-sedimentary rocks of the
Piuntza Unit that were deposited in rift grabens (Litherland et al.,
1994; Fig. 2). Anatexis during low pressure metamorphism and extension has been recorded in several locations, including the eastern Mt.
Lofty Ranges (Oliver and Zakowski, 1995) and the southern Menderes
Massif, Turkey (Bozkurt and Park, 1994). Gerbi et al. (2006) utilize thermal modelling to show that crustal scale detachment faulting is required to provide sufcient heat for low-pressure anatexis. A suspect,
Triassic detachment fault has not been identied within the Eastern
(Ecuador) and Central Cordillera (Colombia), and it is likely that is
was reactivated during subsequent tectonism (see below).
No products of melting continental crust have been found younger
than 225 Ma, and the basaltic amphibolites that formed during and
after 225 Ma yield N-MORB isotopic and geochemical signatures
(Figs. 6 and 8). Cochrane et al. (2014a) suggest that the continental
crust was either extremely thin, or not present, after 225 Ma. Martnez
(2007) document a series of metagabbros, amphibolites and plagiogranites from the northern Cordillera Central, and utilize petrological
observations, isotopic and geochemical analyses and eld mapping
to conclude that they are part of an ophiolitic sequence (the Aburr

17

Ophiolite; Fig. 2) that formed within a back-arc basin. U-Pb dating of


magmatic zircons from the plagiogranite yields an age of ~ 216 Ma
(Table 1) suggesting that seaoor spreading started between ~ 223
and ~ 216 Ma. The youngest plateau 40Ar/39Ar date of 213.7 0.9 Ma
(Fig. 2; Table 1) obtained from an anatectite in Colombia records cooling
below ~300 C as the margin drifted further from the source of elevated
heat ow, and possibly exhumation during continued extension. Some
regions of the continental margin remained at temperatures that
were sufciently high to form metamorphic zircon rims until ~207 Ma
(Fig. 3d; Table 1; Cochrane et al., 2014a). The rift and transition to
drift occurred over a period of ~2025 My (Fig. 8), which is comparable
with the duration of other examples of rifting and the transition to a
drift phase, including the Lau-Havre-Taupo system (south Pacic;
Parson and Wright, 1996), and the west IberianNewfoundland conjugate margins (Russell and Whitmarsh, 2003). The combination of
bimodal magmatism, an increase in the depleted mantle component
of mac magmas, increased partial melting of the crust, geochemical
compositions of a back-arc basin or mid-ocean ridge setting, and the formation of oceanic lithosphere at ~216 Ma (Aburr Ophiolite) is entirely
inconsistent with collisional orogenesis.
Rifting starting at ~240 Ma is consistent with the geochronology and
geochemistry of highly peraluminous granites and volcanic tuffs in the
Eastern Cordillera of Peru. Mikovi et al. (2009) suggest that decompression melting generated basaltic magmas at the base of the lower
crust, resulting in crustal melting during 260200 Ma, ultimately giving
rise to the bimodal Mitu Group within a continental rift. The range
in dates of magmatic rocks of the Mitu Group (217238 Ma; see
Section 4.2.2) closely overlaps with the timing of rifting within
Ecuador and Colombia. Reitsma (2012) interpret the rift as a back-arc
basin on the assumption that the Triassic arc has been removed by tectonic erosion. There is no evidence for the formation of Triassic oceanic
lithosphere within the Eastern Cordillera of Peru, although the rift foundered during ~220190 Ma, resulting in the deposition of limestones of
the Pucar Group during thermal sag (Rosas et al., 2007). Extensive
tracts of continental TriassicPrecambrian continental crust, including
the Arequipa Terrane, lie outboard of the Mitu Group, and it is likely
that the rift failed to advance to a drift phase, and remains as an
aulocagen (Fig. 9). In contrast, no Triassic continental crust is found outboard of the Triassic anatectites within Colombia and Ecuador (Fig. 2),
and this region advanced to the drift phase. The recent discovery of Triassic basalt offshore northern Peru (8S; Romero et al., 2013) suggests
that this may also be a product of extension of the continental crust.
These rocks are exposed along the Outer Shelf High (Romero et al.,
2013), and no continental crust has been found to the west, suggesting
that the Central American conjugate margin may have extended into
Peru, within Pangaea.
Triassic U-Pb dates obtained from the amphibolites and crustal
anatectites from northern Colombia to southernmost Ecuador reveal
no trend with latitude (Fig. 3d), suggesting that extension and rifting
was not diachronous within the northern Andes. Similarly, Reitsma
(2012) compare U-Pb dates of detrital zircons within the Mitu
Group along the strike of the Eastern Cordillera in central and southern Peru, and conclude that the basal strata were coeval everywhere,
although Triassic plutons young from central towards southern Peru
(Mikovi et al., 2009). Rifting along western Pangaea commenced at
~ 240 Ma and affected more than 2500 km of the Gondwanan margin.
Evidence includes Middle to Late Triassic extensional basins that are
dispersed throughout large regions of South America, including
southern Chile, western Argentina (230246 Ma from the Cuyana
Basin; Franzese and Spalletti, 2001; Barredo et al., 2012), southern
Brazil (Zerfass et al., 2004) and Bolivia (Sempere et al., 2002), suggesting that western Gondwana was placed under tension (e.g.
Ramos, 2009) at ~ 240 Ma, culminating in the fragmentation of
western Pangaea by ~ 180 Ma.
We tentatively suggest that rifting along north western South
America at least initiated within a back-arc basin (Fig. 9). The arc

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

18

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

20
DM
Extension,
crustal growth

Hfi (zircon)

10

CHUR

0
-10
Crustal
Recycling

-20

Metagranite zircon cores


Metagranite zircon rims

-30

Amphibolite zircons
-40
150

250

350

550

750

206Pb/238U

950

1150

age (Ma)

DM
15

Hfi (zircon)

Hfi (zircon)

10
N-MORB

5
0
-5

Range of anatectites

-10
MS
MS
MSWD
0.78

M
MS
MSWD
1.60

10RC28

11RC14

220

11RC10

225

230

100 um

10RC39A
234.0 5.8
Hf = + 8.4

whole rock
(La/Yb)n

age (Ma)

238.3 8.7
Hf = + 8.6

240.0 3.5
Hf = - 4.1

236.9 7.1
Hf = - 4.8

100 m

240

237.5 7.1
Hf = + 7.4

241.0 4.0
Hf = - 3.6

10RC39A

235

206Pb/238U

MS
S
MSWD
1.90

100 um

MSWD
WD 1.90

235.3 6.1
Hf = + 9.1

232.7 6.3
Hf = +10.0

Granitoids, zircons, Eastern Cordillera, Peru

Hfi (zircon)

4
2
0
-2
-4
-6
-8
-10
200

Eastern Cordillera, Peru


Merida Andes, Venezuela
220

240
206Pb/238U

260

280

300

age (Ma)

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

rocks are preserved along the back-bone to the southern coast of


Mexico, and within the Chaucs complex of central Guatemala (Solari
et al., 2011). These granites yield zircon U-Pb dates spanning from the
Early Permian until ~ 218 Ma (Torres et al., 1999; Keppie et al., 2006;
Solari et al., 2011), and thus they were coincident with rifting. Furthermore, Helbig et al. (2012) describe transitional arc to MORB like metavolcanic rocks within the Acatlan Complex (southern Mexico), which
these authors assigned to a back arc setting in the late Triassic. This interpretation is consistent with Martnez (2007) who suggests that the
Aburr ophiolite formed in a back-arc.
We consider rifting between 240 and 216 Ma within northwestern South America to represent the initial stage of the disassembly
of Pangaea because i) it was geographically extensive along western
Gondwana, ii) it resulted in the formation of oceanic lithosphere to
the west of the future site of the Gulf of Mexico and the locus of subsequent disassembly within Pangaea, giving rise to the Central
Atlantic Magmatic Province at 202 Ma (e.g. Beutel, 2009), and iii) it
overlapped in time with rifting in south eastern North America
(~ 230 Ma; Schlische, 2002). However, it is unclear what caused the
locus of extension to displace eastwards, leading to fragmentation
of Pangaea.
The Pacic margin of northwestern Gondwana remained passive
until ~ 209 Ma, after which a prolonged period of latest TriassicEarly
Cretaceous magmatism (see Section 5) modied the margin of
Colombia and Ecuador, and southernmost Peru.

4.8. Conjugate margins to northwestern Gondwana


Several lines of evidence suggest that the basement to Central
American terranes (e.g. Oaxaquia, which is exposed within the Maya
Block) may have been the conjugate margin to northwestern Gondwana (Fig. 9). These include i) Grenville-aged granulite belts with similar
Pb isotope signatures (Restrepo-Pace et al., 1997; Ruiz et al., 1999; Cameron et al., 2004), ii) anorthosite complexes (Keppie and Gutierrez,
1999; Restrepo-Pace and Cediel, 2010), iii) similar Cambrian fauna
(Cocks and Torsvik, 2002; Restrepo-Pace and Cediel, 2010), and iv) Triassic bimodal magmatism and rift-related sedimentary rocks (Keppie
et al., 2006). Late Triassic (216197 Ma) tholeiitic basalts in the Guerrero composite terrane of Mexico are interpreted to have formed in a continental rift (Keppie et al., 2006), and the same authors attribute Late
Triassic rocks of the Oaxaquia Terrane to a back-arc setting. Similarly,
Helbig et al. (2012) suggest that late Triassic volcanic rocks with EMORB to MORB characteristics within the Acatln Complex formed
within a back-arc. Weber et al. (2007) and Cochrane et al. (2014a) use
these similarities to suggest that Oaxaquia and other Mexican terranes
were juxtaposed against northwest Gondwana. A potential obstacle to
this reconstruction is the lack of magmatic rocks in southern Mexico
with crystallisation ages spanning between 230 and 216 Ma. However,
an asymmetric distribution of magmatic rocks across continental rifts
has also been documented across the Nova ScotiaMorocco Rift (latest
TriassicJurassic; e.g. Dehler, 2012), where reduced magma production
is related to the proximity of the transform margin (Funck et al., 2004).
Alternatively, Ramos and Aleman (2000) and Ramos (2008) propose
that continental crust of Oaxaquia was located offshore central and
northern Peru, based on correlations between trilobite assemblages. In
this case, the reconstruction of Central American and Mexican basement
relative to South America within Pangaea will need to be revised.

19

Regardless of the true position of Oaxaquia, Triassic roll-back of


the east-dipping Pacic slab drove widespread extension along
N2500 km of western Gondwana, resulting in complete rifting of
continental crust in the region of northwestern South America.
Tensile forces failed to form oceanic lithosphere further to the
south (e.g. the Mitu Aulocagen; Fig. 9), perhaps because the orthogonal component of roll-back was less severe, or the continental
crust was stronger.
4.9. Rifting between North and South America
The relationship between the timing of rifting that separated North
and South America, and back-arc rifting along western Pangaea is unclear. Rifting commenced in south-eastern North America at ~230 Ma
(Schlische, 2002), and possibly also within far southern North America
(e.g. northern Florida; Beutel, 2009). Magmatic injection in the form of
giant dyke swarms (Central Atlantic Magmatic Province) accompanied
rifting within North America at 200 2 Ma (Beutel, 2009 and references therein), which coincides with ~ 202 Ma tuffs at the base of the
rift-related, la Quinta Fm. of the Merida Andes, Venezuela (Fig. 1; Van
der Lelij, 2013), and oceanic crust started forming between North and
South America at ~180 Ma. Beutel (2009) utilise nite element modelling of stress elds obtained from dyke and rift orientations to suggest
that the separation of North and South America was initiated by the migration of North America towards the northwest at ~230 Ma, giving rise
to amagmatic rifting within North America, followed by the displacement of South America towards the southwest at ~ 200 Ma, forming
coeval magmatism, rifting, and seaoor formation by ~180 Ma. The relationship between slab-retreat tensional forces that drove extension
along western Gondwana starting at 240 Ma and forces that separated
North and South America is unclear.
5. Latest TriassicLower Cretaceous: arc magmatism and
tectonic switching
Jurassic rocks within the Cordillera Real and Cordillera Central are
dominated by extensive granitoid batholiths that form the eastern
anks of the cordilleras (Fig. 1). A majority of these intrusions are not
metamorphosed or foliated, with the exception of parts of the aerially
extensive Ibagu Batholith in central Colombia, proximal to major
faults, and the Azafrn Batholith in Ecuador. The Jurassic intrusions
are accompanied by coeval volcanic rocks that are mainly exposed
in the foreland basin to the east, and extend from northern Chile to
Colombia (Romeuf et al., 1995). Some of the protoliths of metasedimentary and meta-volcanic rocks located to the west of the batholiths may also be Jurassic.
5.1. Historical perspective and occurrence
5.1.1. Latest TriassicJurassic granitoid intrusions
Jurassic plutons within the eastern ank of the Cordillera Real in
Ecuador form, from north to south, the undeformed and unmetamorphosed Rosa Florida, Abitagua and Zamora batholiths (Fig. 10). The
granitoids were rst described by Colony and Sinclair (1932) and Saur
(1950), and the British Geological Survey mapped these intrusions
(Litherland et al., 1994). These rocks are generally coarse, biotite quartz
monzonites, monzogranites and hornblende-biotite granites, granodiorites and minor diorites. The batholiths are mainly in faulted contact

Fig. 6. A) Hf (zircon) isotope data (Cochrane et al., 2014a) acquired from rims and xenocrystic cores of Permian and Triassic peraluminous granitoids, migmatitic leucosomes and
amphibolitic dykes and sills of the Amotape Complex, Cordillera Real and the Cordillera Central. Hfi was determined using zircon crystallisation dates determined using LA-ICPMS,
and the CHUR composition (176Lu/177Hf = 0.0336, 176Hf/177Hf = 0.282785; Bouvier et al., 2008). B) Variation in Hfi (zircon) of the amphibolites with zircon crystallisation date and
whole rock (La/Yb)n (N-MORB), showing a trend towards depleted mantle isotopic and geochemical signatures, with time. C) Representative cathodoluminescence images for
amphibolite 10RC39A (Cordillera Central) show that juvenile Hfi values (7.4 to 10) are yielded by patchy or unzoned zircons, whereas less juvenile values (3.6 to 4.8) are obtained
from oscillatory zoned zircons. D) Hfi (zircon) data from Permo-Triassic granitoids of the Eastern Cordillera of Peru (rims and cores; Mikovi et al., 2009), and the Merida Andes (mean
values; Van der Lelij, 2013).

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

20

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

09RC42, migmatitic leucosome, Sabanilla Unit, zircon U-Pb 247.24.3 Ma


Burial and elevated
Collision of
heat flow
Caribbean Plateau
Anatexis
16

Hfi (zircon)

400

constrained by
(U-Th)/He, FT and
40Ar/39Ar data.

-8

APbPRZ

Temperature (C)

200

40Ar/39Ar

muscovite
Quebradagrande Complex
Azafrn and Chingul batholiths
Zamora, Abitagua batholiths
Ibague, Segovia batholiths
Triassic amphibolitic dykes

600

800
Permian

300

250

Jurassic

Triassic

200

Cenozoic

Cretaceous

150
Time (Ma)

100

50

10RC43, granite, Cajamarca Complex, zircon U-Pb 245.02.0 Ma


Burial and elevated
Collision of
heat flow
Caribbean Plateau
Anatexis
16

Hfi (zircon)

200

400

constrained by
(U-Th)/He, FT and
40Ar/39Ar data.

APbPRZ

Temperature (C)

8
0

-8

206Pb/238U date (Ma)


240

40Ar/39Ar
muscovite

235
230

600

225

50

100

150

200

Effective diffusion radius (m)


Jurassic

Triassic

250

200

Cretaceous

150
Time (Ma)

Cenozoic

100

50

Cordillera Central

Cordillera Real

20

Temperature (C)

40

60

AHePRZ

80

APAZ

100

120
160

APAZ

140

ArPRZ (kspar)

200
240

AHePRZ

?
?

180

280

260

320
Cretaceous

140

120

220

ZPAZ

100

300

Cenozoic

80
60
40
Time (Ma)

20

120

ZPAZ

?
Cretaceous

100

Cenozoic

80

60
40
Time (Ma)

20

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

with contemporaneous volcanic rocks of the Misahuall Fm. to the east,


and they may intrude their volcanic pile in a few locations. Late Jurassic
Early Cretaceous schists of the Upano Fm. (the Salado Terrane;
Litherland et al., 1994) are juxtaposed against the western margin of
the Jurassic batholiths via the Cosanga Fault (Fig. 10). The foliated and
occasionally gneissic Azafrn Batholith and Chingul pluton are also
located to the west of the Cosanga Fault, within the Salado Terrane of
Litherland et al. (1994; Fig. 10). These plutons are dominated by diorites
and granodiorites, with minor granite, quartz monzonite and gabbro.
Litherland et al. (1994) present Rb/Sr isochron and K/Ar dates from
all of these intrusive phases, although more accurate estimates of the
crystallisation ages are provided by U-Pb zircon dates (Litherland et al.,
1994; Noble et al., 1997; Chiaradia et al., 2009; Cochrane, 2013). Geochemical analyses of these rocks have been presented by Litherland
et al. (1994) and Cochrane (2013).
Jurassic batholiths extend northwards from Ecuador, and dene
the eastern ank of the Cordillera Central in Colombia (Fig. 10). These
include the geographically extensive Ibagu Batholith, and Segovia
batholith, where the lithologies are dominated by hornblende tonalities
and granodiorites (Alvarez, 1983; Aspden et al., 1987). Jurassic intrusions also crop out in the Sierra Nevada de Santa Marta where tonalities
are also common (Alvarez, 1983), and the Santander Massif (Figs. 1 and
10), where the intrusions are mainly biotite monzonites, granodiorites
and granites (Irving, 1975). These intrusions are generally unfoliated,
and they crop out to the east of the Triassic anatectites. Numerous
authors presented a large quantity of K/Ar and Rb/Sr dates from these
intrusions in the 1980's (e.g. Goldsmith et al., 1971; see additional references in Aspden et al., 1987), although more accurate zircon U-Pb dates
have been published by Leal-Mejia et al. (2011), Villagmez et al.
(2011), Van der Lelij (2013) and Cochrane (2013).
5.1.2. Late JurassicEarly Cretaceous rocks to the west of the
Jurassic intrusions
A series of poorly dated, JurassicEarly Cretaceous meta-sedimentary
and meta-volcanic rocks, along with the Azafrn Batholith, is exposed to
the west of the Cosanga Fault in the Cordillera Real of Ecuador, and
similar sequences crop out within the Cordillera Central (Fig. 10). Within
Ecuador, Aspden and Litherland (1992) and Litherland et al. (1994)
utilise structural, metamorphic and geochronological relationships to
group these into several north-south trending, fault bounded terranes,
which accreted via dextral strike-slip transpression. However, Pratt
et al. (2005) re-interpret several of these faults as intrusive contacts,
and suggest that the Cordillera is an eroded and uplifted Palaeozoic
core, anked by autochthonous JurassicCretaceous sequences.
Meta-turbidites and meta-andesites of the Upano Unit, which were
originally described by Saur (1965), are located west of the Jurassic
batholiths (Fig. 11). Baldock (1982) and Pratt et al. (2005) suggest
that the metasedimentary rocks are equivalent to Cretaceous sedimentary rocks (Hollin and Napo fms.) that are exposed within the region of
the Amazon Foreland Basin (Fig. 1). However, Litherland et al. (1994)
regard the Upano Unit as having a distinctly different history based on
their metamorphic grade, and assign them, along with the Azafrn Batholith, to the fault bounded Salado Terrane. Sparse geochronological
analyses include pollen analyses (Riding, 1989), K/Ar dates (Feininger
and Silberman, 1982), and U-Pb zircon dates of the Azafrn Batholith

21

(Cochrane, 2013). Litherland et al. (1994) consider the Salado Terrane


to have formed within a marginal basin setting. Traversing westwards,
the Upano Unit is faulted against (Litherland et al., 1994; Llanganates
Fault; Figs. 10 and 11) poorly differentiated Palaeozoic metasedimentary rocks and Triassic anatectites (see Section 4) of the Loja
Terrane of Litherland et al. (1994). Pratt et al. (2005) interpret this contact as a metamorphosed intrusive contact. Further west, the Triassic
anatectites are faulted against greenschist grade para-graphitic schists,
and submarine meta-andesites and meta-agglomerates across the
Baos Fault (Fig. 11), which were rst described by Sheppard and
Bushnell (1933), and subsequently by numerous workers. There have
been few geochronological determinations, which include K/Ar dates
(Herbert and Pichler, 1983; Litherland et al., 1994) and Jurassic pollen,
spores and acritarchs (Riding, 1989). Bristow (1973) considered these
lithologies to be equivalent to sedimentary rocks that are now exposed
within the Western Cordillera of Ecuador (Yunguilla Fm.), although
Litherland et al. (1994) interpret these to represent a distinct island
arc terrane, and recognise forearc, arc and backarc sequences, which
they refer to as the Alao Arc (Figs. 10 and 11). The western margin of
these sequences is dened by the 12 km wide Peltetec fault zone,
which hosts inliers of steeply dipping, anastomosed blocks of peridotite,
olivine gabbros, spilitised dolerites, basalt and volcanoclastic rocks
(Fig. 11). These rocks were initially documented by Litherland et al.
(1987), and are described in detail in Fortey (1990) and Litherland
et al. (1994) who suggest that they dene an ophiolitic assemblage,
which they refer to as the Peltetec Unit. Pratt et al. (2005) describe a
stratigraphic transition from the greenschists of the Alao Arc into
metasedimentary rocks located west of the Peltetec Unit, and thus
they rule out the Peltetec Fault as a terrane boundary. This publication
presents new 40Ar/39Ar dates from plagioclase extracted from basalts
of the Peltetec Unit, and we are not aware of any other geochronological
data. Finally, the Peltetec Fault denes the eastern boundary of a series
of gently dipping slates and quartzites, which are exposed in isolated inliers. These marine rocks were rst reported by Litherland et al. (1994),
and host JurassicEarly Cretaceous ammonites. Litherland et al. (1994)
named these the Guamote Sequence and suggest that they form part
of allochthonous continental crust of the Chaucha Terrane (Figs. 10
and 11), whereas Pratt et al. (2005) suggest that they are a stratigraphic
continuation of the Alao volcanic sequence.
The same sequence of rocks is also exposed with the Amotape
Terrane of southern Ecuador (Fig. 10), which is rotated ~65 clockwise
relative to the Cordillera Real. Here, the Triassic anatectites are faulted
against a complicated mlange of deformed and metamorphosed igneous rocks that are considered to represent an accretionary prism. The
mlange hosts eclogites and blueschists of the Raspas unit (Litherland
et al., 1994; Arculus et al., 1999; Bosch et al., 2002; John et al., 2010),
which may be a tectonic equivalent of the Peltetec Unit.
Metamorphosed igneous and sedimentary rocks of the Quebradagrande Unit are juxtaposed, via the San Jernimo Fault, against the Triassic Cajamarca Unit in the Cordillera Central of Colombia, and thus they
are located in a similar structural position as the Alao Terrane in Ecuador
(Figs. 1 and 10). The Quebradagrande Unit was rst described by Botero
(1963) and is exposed in highly deformed inliers along the entire length
of the Cordillera Central. Lithologies are dominated by low-grade metamorphosed basalts, andesites and pyroclastic rocks (e.g. Nivia et al.,

Fig. 7. Time (t)Temperature (T) plots for A) a Triassic leucosome (09RC42), and B) a Triassic peraluminous granite (10RC43) of the southern Cordillera Real, and the northern Cordillera
Central, respectively (locations shown on Fig. 2). 09RC42 is from Cochrane et al. (2014b), and 10RC43 is new data (see supplementary data). 40Ar/39Ar muscovite plateau dates are also
shown, with the closure temperature range determined from the diffusion parameters presented in Harrison et al. (2009). The tT paths are determined (see text) by a using a computed
Monte Carlo algorithm constrained by the diffusivity and activation energy of diffusion of Pb in apatite (Cherniak et al., 1991), apatite grain size and U-Pb date (determined using TIMS).
The tT paths colder than 350 C and younger than 75 Ma are constrained by multi-phase 40Ar/39Ar, ssion track and (U-Th)/He data (Spikings et al., 2000, 2001, 2005, 2010). The blue line
is the best-t solution. The red dashed line has not been computed, and only serves to illustrate heating of a Palaeozoic sedimentary protolith during anatexis. The relationship between
diffusion radius (related to grain size) and U-Pb date is shown for granite 10RC43, and the indistinguishable dates corroborate fast cooling through the APbPRZ (apatite Pb Partial Retention
Zone), and no subsequent reheating into it. H (zircon) are also shown, and reveal i) the trend towards Depleted Mantle during the Triassic, and ii) steadily increasing values throughout
the Jurassic to the middle of the Cretaceous. C) tT envelopes (1 envelope describes 1 hand specimen) for Early Cretaceous and older rocks of the Cordillera Central of Colombia, and the
Cordillera Real of Ecuador constrained by 40Ar/39Ar, ssion track and (U-Th)/He data (Spikings et al., 2000, 2001, 2010; Villagmez and Spikings, 2013). APAZ: Apatite Partial Annealing
Zone, AHePRZ: Apatite Helium Partial Retention Zone, ArPRZ: Argon Partial Retention Zone, ZPAZ: Zircon Partial Annealing Zone.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

22

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

AMPHIBOLITES

CRUSTAL ANATECTITES
Th/Uzircon
0.001

0.1

(La/Yb)n
10 8 10 12 14 16 18 20 20

Hfizircon
15

10

(La/Yb)n

Ndi
5 12 10

Tectonic
stage

190

DM

200

N-MORB

Jurassic

180

Active
margin

210
Aburr
ophiolite

Time (Ma)

230

Triassic

220

DRIFT/
Passive
margin

RIFT

240

250
Ecuador
Zircon

260

Whole rock

Colombia
Permian

Zircon

270

Whole rock

Early
anatexis
Continental
Arc?

280

290

Fig. 8. Geochronological and geochemical summary of data from Permian and Triassic crustal anatectites, tholeiitic basaltic sills and dykes, and the Aburr Ophiolite. These data dene
i) Permian to earliest Triassic anatexis (290240 Ma) during arc magmatism, ii) bimodal magmatism during lithospheric thinning and continental rifting (240223 Ma), and iii) a passive
margin stage during which oceanic lithosphere was forming (223 Ma209 Ma). Data are from Vinasco et al. (2006), Martnez (2007) and Cochrane et al. (2014a).

2006; Villagmez et al., 2011; Rodriguez and Zapata, 2013), and marine
metasedimentary rocks that are covered by marine and terrestrial
meta-sedimentary rocks of the Abejorral Fm., which hosts Hauterivian
to lower Albian fossils (Gonzlez, 1980). The highly deformed nature
of the Quebradagrande Unit lead Maya and Gonzales (1996) to assign
to term Quebradagrande Complex, which will be used in this review.
Fossil evidence suggests that the sedimentary component was deposited during the Berriasian to Aptian (145112 Ma using the timescale
of Gradstein et al., 2004; see Nivia et al., 2006 and references therein).
Other geochronological analyses include K/Ar dates (Toussaint and
Restrepo, 1978), and concordant U-Pb (zircon) dates (Nivia et al., 2006;
Villagmez et al., 2011; Cochrane, 2013). Geochemical analyses of the
meta-volcanic rocks are presented by Nivia et al. (2006), Villagmez
et al. (2011), Rodriguez and Zapata (2013) and Cochrane (2013). Tectonic interpretations vary from mid-ocean-ridge (e.g. Gonzlez, 1980;
Bourgois et al., 1987), oceanic arc (Villagmez et al., 2011), continental
arc (Cochrane, 2013) and an intra-cratonic marginal basin (Nivia et al.,
2006). Vsquez and Altenberger (2005) document small, Early Cretaceous intrusions within the Colombian Eastern Cordillera (Fig. 1),
which they relate to rifting.
The Quebradagrande Complex is in contact with faulted slices of
mac igneous rocks that have been metamorphosed under mediumto high-pressure and low temperature conditions. These rocks are exposed in discontinuous, fault-bounded lenses along the strike of the
Cordillera Central, and include the Arqua (e.g. Restrepo and Toussaint,
1976), Jambal (Orrego et al., 1980; Bustamante et al., 2011) and

Barragn (e.g. Bustamante et al., 2012) complexes, amongst others


(Fig. 10). These authors describe eclogite and blueschist facies metamorphic rocks (epidote glaucophane schists and chlorite-lawsonite
schists), associated with amphibolites, meta-ultramac rocks, serpentinites and protocataclasites. Rocks exposed in the Jambal region are
faulted against the eastern margin of the Quebradagrande Complex
(Bustamante et al., 2008), although the Arqua and Barragn sequences
crop out to the west of the Quebradagrande Complex, across the SilviaPijao Fault, and their western margin is limited by the Cauca-Almaguer
Fault (Fig. 10). Early radiometric dates of these rocks were K/Ar analyses
(Feininger and Silberman, 1982), and numerous 40Ar/39Ar dates have
been reported (Bustamante et al., 2008; Villagmez et al., 2011;
Bustamante et al., 2012). The Arqua and Barragn complexes occupy
an equivalent tectonic position as the Peltetec and the Raspas
complexes in the Cordillera Real and Amotape Terrane of Ecuador,
respectively.
5.2. Geochronology
5.2.1. Latest Triassic and Jurassic intrusions: Cordillera Real, Cordillera
Central and the Santander Massif
Goldsmith et al. (1971), McCourt et al. (1984), Aspden et al. (1987)
and Litherland et al. (1994) present Rb/Sr isochron and K/Ar dates from
all of these intrusive phases, although they are not considered to be accurate crystallisation ages given the high potential for daughter isotope
loss. More accurate estimates of the crystallisation ages are provided by

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

A
240 Ma, extension begins
Slab
rollback

23

A Permian - 240 Ma
Ouachita-Marathon-Suture
Laurentia

Pangaea
Back-arc
basin formation
Permian
arc

Mexico

Oceanic crust
Lithosphere

A
Pacific
Ocean

Asthenosphere

240 - 225 Ma
Slab
rollback

A
Ch

Gondwana

Peru

Pangaea
Attenuated and underplated
crust

Arc
240- 210(?) Ma
Laurentia

225 - 216 Ma

NW Gondwana
Central American
Basement blocks
(Maya?)
Continental break-up
and ophiolite formation

Roll
back

Mexico

Ch

A
Mit

u(

Gondwana
au
loc
ge
n)

Fig. 9. Schematic reconstruction and cross sections for northwestern South America within western Pangaea during 240216 Ma, showing the location of the Permian arc, and the Triassic
rift axis. The Permian arc is preserved along the Gondwanan conjugate margin in Mexico, and the Eastern Cordillera of Peru, although few remnants remain in the region of the Northern
Andes. The approximate location of the Mitu Group is shown. The reconstruction is based after Grajales-Nishimura et al. (1999), Golonka and Bocharova (2000), Dickinson and Lawton
(2001), Elas-Herrera and Ortega-Gutirrez (2002), Weber et al. (2007) and Cochrane et al. (2014a).

concordant U-Pb zircon dates (Fig. 10; Table 2), which range between
194 and 209 Ma in the Santander Massif (Van der Lelij, 2013), 147
189 Ma in the Cordillera Central (Bustamante et al., 2010; Villagmez
et al., 2011; Cochrane, 2013) and 141182 Ma in the Cordillera Real
(Chiaradia et al., 2009; Cochrane, 2013). Leal-Mejia et al. (2011) report
a range of zircon U-Pb dates from the Santander Massif and Cordillera
Central (210149 Ma), although they do not present their data, and
thus they cannot be evaluated. 40Ar/39Ar plateau dates from retentive
phases (hornblende; closure temperature 500550 C) are reported
here because they may be reasonable estimates of the time of crystallisation, although they are frequently younger than the U-Pb zircon
dates (Table 2). A comparison of zircon U-Pb age and latitude (Fig. 12)
suggests that the timing of the onset of magmatism may become younger towards southern Ecuador. However, a better dened trend is seen
when the dates are compared with their distance from the equivalent
Silvia-Pijao and Peltetec Faults (Fig. 11b), indicating that Jurassic
magmatism becomes younger as it approaches the approximate
location of the contemporaneous plate margin. Clearly, latest Triassic
magmatism initiated far from the trench at ~ 210 Ma, and these rocks
are currently exposed within the Santander Massif (209194 Ma;
Fig. 10; Van der Lelij, 2013) and in the region of Mocoa in the far southern Cordillera Central (Leal-Mejia et al., 2011). Magmatism migrated
westwards at ~ 195 Ma and stabilised within the region that is now
exposed within the Central Cordillera, throughout the Jurassic. The
older magmatic belt (N189 Ma) is not exposed, or did not form in
Ecuador. The earliest Cretaceous (Gradstein et al., 2012) intrusions of
the foliated Azafrn Batholith (141144 Ma; Table 2) occur to the
west of the mainly unfoliated Jurassic intrusions (Fig. 11) of the Segovia,

Ibagu, Rosa Florida, Abitagua and Zamora batholiths (145189 Ma).


The range in dates of the unfoliated Jurassic intrusions closely overlaps
with the age of Jurassic arc magmatism recorded in southern Peru
(Fig. 12), Patagonia and northern Chile (195147 Ma; zircon U-Pb and
K/Ar dates; Scheuber and Gonzalez, 1999; Rapela et al., 2005; Castro
et al., 2011).
5.2.2. Early Cretaceous magmatic and sedimentary rocks: Cordillera Real
and Cordillera Central
A majority of dates obtained from the Upano, Alao Arc, Quebradagrande Complex and the mac and ultramac rocks of the Peltetec,
Arquia, Barragn and Jambal complexes are K/Ar and Rb/Sr dates
(E.g. Feininger and Silberman, 1982). A meta-andesite of the Upano
Unit yields a zircon U-Pb crystallisation age of 121.0 0.8 Ma
(Cochrane, 2013). Cochrane (2013) present a U-Pb zircon age for the
Chingul Batholith of 125.3 0.9 Ma (Fig. 10), which is considered to
reside in the same Salado Terrane (Litherland et al., 1994) as the
Upano Unit, and the Azafrn Batholith (141144 Ma; Table 2).
Cochrane (2013) presents 238U-206Pb dates from the rims and cores of
detrital zircons extracted from a quartzite of the Upano Fm. (Fig. 13),
which reveals a large spread in ages from rims and cores, and the youngest age of 143.3 9.9 Ma constrains the maximum stratigraphic age to
the latest Jurassic. These dates suggest that the contacts between the
Azafrn Batholith and parts of the Upano Unit are unconformable,
supporting the interpretation of Pratt et al. (2005). Age peaks occur at
500600 Ma, 9001200 Ma, 1500 Ma and the oldest is ~ 2 Ga. The
maximum stratigraphic age is consistent with pollen assemblages that
have a poorly resolved Early JurassicCretaceous age (Riding, 1989).

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

24

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

No zircon U-Pb dates have been obtained from the meta-volcanic


rocks of the Alao Arc, located west of the Triassic anatectites (Loja Terrane) and the Baos Fault (Fig. 10) within Ecuador. Litherland et al.
(1994) report imprecise K/Ar (hornblende) dates of 115 12 Ma and
142 36 Ma, and pollen assemblages suggest that the volcanoclastic
sedimentary rocks are Middle Jurassic (Riding, 1989). If the pollen occurrence is accurately calibrated against absolute time, then the K/Ar
hornblende dates are partially reset. Cochrane (2013) presents 238U206
Pb dates from the rims and cores of detrital zircons extracted
from a quartzite of the Alao Arc (Fig. 13), yielding a minimum age of
163.7 1.6 Ma (Table 2), constraining its maximum stratigraphic age.
Similar detrital age populations are found compared to the detrital
dates from the Upano Unit, suggesting that they were supplied by the
same source regions. Furthermore, Early Ordovician acritarchs suggest
that they have been reworked into the Alao arc sequence (Litherland
et al., 1994), corroborating the detrital zircon U-Pb age spectrum
(Fig. 12). We conclude that no reliable age dates exist for the Alao arc,
although it is Middle Jurassic, or younger. Villagmez et al. (2011))
and Cochrane (2013) report concordant zircon U-Pb dates of magmatic
rocks of the Quebradagrande Fm. of 114.3 3.8 Ma (tuff) and 112.9
0.8 Ma (diorite) (Fig. 10; Table 2). Toussaint and Restrepo (1994) report
a K/Ar (whole rock basalt) date of 105 0.8 Ma, although it is likely to
be at least partially reset. The U-Pb dates are consistent with Berriasian
to Aptian (145112 Ma; Nivia et al., 2006 and references therein) fossil
dates, and a maximum stratigraphic age of ~149 Ma obtained from detrital zircons (Fig. 13). Vsquez et al. (2010) report hornblende 40Ar/
39
Ar (plateau) dates of 120.5 0.6 Ma (tuff) and 136.0 0.4 Ma (diorite) from small gabbroic intrusions located in the Eastern Cordillera of
Colombia (Figs. 1 and 11; Table 1).
Dating the crystallisation ages of the mac and ultramac protoliths
of the MP-HP/LT rocks of the Arqua, Barragn, Jambal, Peltetec and
Raspas complexes is problematic due to i) metamorphic overprinting,
and ii) the lack of high-U, low-(common)Pb minerals, which are amenable to U-Pb dating. No U-Pb dates have been published, although this
could be addressed by recent developments in dating baddeleyite (e.g.
Chamberlain et al., 2010), which is stable in mac rocks. Bustamante
et al. (2012) present a weighted mean 40Ar/39Ar date from three plateau
white mica dates of 120.7 0.6 Ma (Fig. 10; Table 2) from a single
muscovite schist of the Barragn Complex, which they suggest records
retrogression during exhumation. This Early Cretaceous date is similar
to a plateau hornblende 40Ar/39Ar date of 112.0 3.7 Ma (Villagmez
et al., 2011) obtained from the Arqua Complex, which has been
interpreted as the time of retrogression. The only other interpretable
dates from Colombia are plateau 40Ar/39Ar white mica (paragonite and
phengite) dates from blueschists of the Jambal Complex, which range
between 68 and 62 Ma (Bustamante et al., 2011), which were interpreted to record the mylonitic event that was responsible for exhuming
the rocks. Numerous dates have been obtained from eclogites and
blueschists of the Raspas Complex in the Amotape Complex (Ecuador).
Phengite K/Ar (Feininger, 1980) and 40Ar/39Ar dates (Gabriele, 2002) of
132 5 Ma and 123129 Ma (Fig. 10) were interpreted to record cooling
during retrogression of the eclogites. However, these dates must be
interpreted with caution, given the propensity for phengite in highpressure rocks to incorporate excess 40Ar, which is not resolvable by inverse isochron analysis (e.g. Sherlock and Kelley, 2002). John et al.
(2010) report Lu-Hf isochron (garnet, whole rock, amphibole, pyroxene)
ages of 126.4 4.0 and 129.9 5.6 Ma, while the older estimate is considered to be more accurate, given its MSWD of 2.0 (Table 2). These
dates are interpreted to record garnet growth (John et al., 2010),
and overlap with the phengite 40Ar/39Ar dates of Gabriele (2002).
Mac and ultramac assemblages of the Peltetec sequence have
only experienced greenschist grade metamorphism, and we present
a plagioclase 40Ar/39Ar plateau dates of 134.7 0.9 Ma and 134
13 Ma (Table 2; details in the supplementary material; Fig. 14),
which we interpret as the minimum crystallisation age of the basaltic
protolith.

5.2.3. Comparison with Peru and the Merida Andes of Venezuela


Jurassic magmatic rocks in Peru are limited to the coast of the
Arequipa Terrane (Fig. 1), and in the southernmost Cordillera de
Carabaya (southern Peru, east of the city of Cusco). There is a gap in
exposure of Jurassic magmatic rocks between southernmost Ecuador
at 5S, and the Paracas Peninsula at ~ 17S (e.g. Sempere et al., 2002;
Boekhout et al., 2012). The Jurassic arc along coastal southern Peru is
the northern termination of a continuous belt that extends to central
Chile (28S; Oliveros et al., 2006). The Ilo Batholith intrudes the
Grenvillian basement of the Arequipa Terrane, and the uppermost levels
intrude volcanic rocks of the Chocolate Fm. The batholith yields concordant zircon U-Pb (LA-ICPMS) dates ranging between 152 and 173 Ma
(Boekhout et al., 2012; Fig. 12b), while the Chocolate Fm. yields zircon
U-Pb ages ranging between 216 and b135 Ma (Boekhout et al.,
2013a). Demouy et al. (2012) obtained concordant zircon U-Pb dates
ranging between 161 and 200 Ma from diorites and gabbros from the
western Arequipa Terrane. Mukasa (1986) report zircon U-Pb dates
between 184 and 188 Ma from a gabbro- tonalite unit located in southern coastal Arequipa, and Mikovi et al. (2009) present zircon U-Pb
dates from four syenites within the Eastern Cordillera, inboard of the
Arequipa Terrane (Cordillera de Carabaya; Allincapac Complex), which
range between 173 and 195 Ma. These dates suggest that Jurassic arc
magmatism may have been active in southern Peru since ~ 216 Ma,
and throughout the Jurassic, which overlaps with the entire span of Jurassic dates found in the northern Andes (Fig. 12). Demouy et al. (2012)
use the distribution of U-Pb dates from Jurassic plutons to suggest that
the arc axis within Arequipa migrated oceanward at ~ 175 Ma. The
Jurassic intrusions host few xenocrystic cores within the zircons, with
ages clustering at approximately ~ 500600, ~ 9001050, and older
(Fig. 12b; Demouy et al., 2012).
Early Cretaceous granitoid intrusions in Peru are concentrated along
the coastline, forming the Coastal Batholith within the Western
Cordillera of Peru, and along coastal Arequipa, although none have
been found in the Eastern Cordillera of Peru (e.g. Mikovi et al.,
2009). Early Cretaceous zircon U-Pb crystallisation ages from the Coastal Batholith located north of the Arequipa Terrane are ~ 100 Ma
(Mukasa, 1986) and 115117 Ma (de Haller et al., 2006), which are similar to the ages obtained from the Quebradagrande Complex in
Colombia (Table 2; Fig. 12). Boekhout et al. (2012) present concordant
zircon U-Pb dates ranging between 106 and 110 Ma from coastal Arequipa, which were originally mapped as the Ilo Batholith, although
they form a distinctly younger age group than the Jurassic granitoids.
A compilation of these dates suggests that there is a magmatic gap in
southern Peru between ~110 and ~132 Ma.
No Cretaceous magmatic rocks have been documented in the Merida
Andes, and this period is characterised by the deposition of sandstones
and marine mudstones (Van der Lelij, 2013).
5.3. Geochemistry
5.3.1. Latest Triassicearliest Cretaceous granitoids
Latest Triassic to Jurassic (209141 Ma) granitoids from the Cordillera
Real, Cordillera Central and the Santander Massif, are metaluminous
(Fig. 15a) and plot in the calc-alkaline to alkali-calcic eld of Frost et al.
(2001; Fig. 15b), and in the high-K calc-alkaline and calc-alkaline elds
when comparing immobile elements (Fig. 16a). Negative Nb, Ta and Ti
anomalies, combined with enriched LILE relative to HFSE, and LREE
relative to HREE suggest that these rocks formed in a subduction related
setting and they are interpreted as continental arc intrusions (Fig. 15c, d).
Granitoids from the Santander Massif, which were probably located
furthest from the plate margin, are slightly more enriched in LILE and
are more peraluminous that the younger granitoids which intruded closer to the palaeo-trench. The magmas within the Santander Massif may
have assimilated more continental crust, perhaps because they intruded
through thicker crust. Finally, the trace and major elements reveal no
along-strike geochemical differences in the magmas that formed within

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

the Northern Andes (Fig. 15). Ndi (whole rock) values for the latest
TriassicJurassic granitoids range between 7.2 and 5.3, and the intrusions within the Santander Massif host distinctly less radiogenic Nd
isotopic compositions than younger intrusions located further west
(Fig. 16b). Ndi (whole rock) values from the unfoliated intrusions of
the cordilleras Real and Central (red and black symbols in Fig. 16b) dene no particular trend. H (zircon) has a large range of 6 to 9.25
(Fig. 16c), and a well-dened trend indicates that the isotopic composition becomes more juvenile as the rocks become younger, although a
large range in Hfi (zircon) can be found in the unfoliated intrusions of
the cordilleras Real and Central, at any given time. The youngest, earliest
Cretaceous foliated intrusions (Chingul and Azafrn) are located to the
west of the older granitoids, and yield the most juvenile isotopic compositions. These trends suggest that the latest Triassicearliest Cretaceous
granitoid intrusions become more juvenile as the intrusions migrate
towards the palaeo-plate margin. The arc intrusions were intruding
through continental crust that was becoming thinner between
~194 Ma and ~ 189 Ma, and during the earliest Cretaceous, starting at
~143 Ma (Azafrn and Chingul intrusions). This could be interpreted
as extension along the plate margin, or the migration of the arc axis
towards the trench, or a combination of both.
5.3.2. Early Cretaceous igneous rocks
Dacites, andesites, basalts and gabbros of the Quebradagrande
(Colombia) and Alao (Ecuador) sequences are metaluminous and span
a larger range in Aluminium Saturation Index than the Jurassic granitoids (Fig. 17a). Tectonic discrimination diagrams suggest that these
rocks formed in a variety of tectonic environments, spanning from
calk-alkaline arc to island arc tholeiite, ocean plateau tholeiite and
MORB compositions (Fig. 17b, c). These observations corroborate i)
the N-MORB normalised REE plot (Fig. 17e) which reveals both NMORB-like compositions, and rocks that are enriched in LREE, which is
more characteristic of subduction-related rocks, and ii) N-MORB normalised trace element plot (Fig. 17d), which shows that some samples
have characteristic negative Nb, Ta and Ti anomalies, while these are
missing in the rocks that yield almost at REE patterns. These rock sequences are exposed within faulted blocks, and distinguishing between
rock sequences that form the Quebradagrande Complex and the juxtaposing Arqua Complex is difcult in the eld. Therefore, we suggest
that some of the basalts that yield N-MORB signatures may be a
structurally detached component of the Arqua or Peltetec complexes, which yield MORB and E-MORB signatures (see later), and
are now intercalated within arc rocks of the Quebradagrande and
Alao sequences. Ndi (whole rock; Fig. 16d) from the volcanic rocks
of the Quebradagrande and Alao sequences ranges between 0.64
and 7.63 (Cochrane, 2013), and there is a general reduction in (La/
Yb)n as Ndi (whole rock) becomes more radiogenic (Fig. 16d). The
least radiogenic basalts within these volcanic sequences are more juvenile than the most radiogenic Nd isotopic compositions obtained
from the Jurassic granitoids. Similarly, the single Hfi (zircon) measurement from the Quebradagrande Complex (Cochrane, 2013) is
more radiogenic that the same measurements from the Jurassic
granitoids (Fig. 16c).
Blueschists and amphibolites of the Barragn and Arqua complexes
yield at REE (N-MORB normalised) multi-element plots ((La/Yb)n
0.744.68), and their trace element abundances lack strongly negative
Nb, Ta and Ti anomalies (Fig. 17f, g), contrasting with arc-related andesites and basalts of the Quebradagrande Unit, which is faulted against
their eastern margin. These features are consistent with a tholeiitic fractionation trend (Fig. 16a), and juvenile Ndi (whole rock) values of 3.2
to 9.6 (Arqua Complex only; Fig. 16d), which are more radiogenic
than the Early Cretaceous volcanic rocks of the Quebradagrande unit.
The MP-HP/LT metamorphosed rocks plot in the MORB to E-MORB
eld when comparing Nb/La with (La/Sm)n (Fig. 17c), and island arc
to ocean-plateau tholeiite eld when comparing La/Yb and Zr/Th
(Fig. 17b). Bustamante et al. (2012) suggest that the protoliths to the

25

blueschists and amphibolites of the Barragn Complex were normal


mid-ocean ridge basalts. Similarly, Villagmez et al. (2011) suggest
that the protolith to the amphibolites of the Arqua Complex may
have formed at a mid-ocean ridge.
Metamorphosed ultramacmac rocks of the Raspas and Peltetec
complexes in Ecuador are geochemically similar to the Barragn and
Arqua units in Colombia. Eclogites of the Raspas Complex yield at
REE proles when normalised against N-MORB ((La/Yb)n 0.692.20;
Arculus et al., 2002; Bosch et al., 2002; John et al., 2010), and their
LILE contents are not signicantly enriched relative to the HFSE, while
negative Nb, Ta and Ti anomalies are missing (Fig. 17h, i). The eclogites
plot in the N-MORB eld when comparing Nb/La with (La/Sm)n
(Fig. 17c), and in the ocean plateau tholeiite eld when comparing La/
Yb and Zr/Th (Fig. 17b). Ndi (whole rock) values for the eclogites are
juvenile, and range between 6.9 and 10.8 (Bosch et al., 2002; Fig. 16d),
corroborating their depleted chemical compositions. In contrast, the
blueschists from the Raspas Complex have enriched LILE relative to
their HFSE, and elevated LREE relative to their HREE ((La/Yb)n 6.0
7.9; John et al., 2010; Fig. 17h, i). The blueschists plot in the seamount
eld when comparing Nb/La with (La/Sm)n (Fig. 17c), and in island
arc tholeiite eld when comparing La/Yb and Zr/Th, although they lack
Nb, Ta and Ti anomalies when normalised against N-MORB. John et al.
(2010); Fig. 17h interpret the eclogites as subducted oceanic lithosphere that was typical of N-MORB, whereas the protoliths to the
blueschists were considered to be seamounts. Bosch et al. (2002) suggest that the mac and ultramac rocks metamorphosed under highpressure conditions and originally formed part of an oceanic plateau,
while Arculus et al. (2002) report that the protoliths originated as
both N-MORB, and within an oceanic plateau, and equilibrated with
peak conditions at 1.32.0 GPa and 600 C. Finally, altered gabbros
and basalts of the Peltetec unit, which is exposed within the Peltetec
Fault Zone (Ecuador; Fig. 11) have received less attention, and the
only available geochemical data are from Litherland et al. (1994), and
new data that are published in this review (supplementary material).
These rocks are generally depleted compared to the other mac rocks
(4550 wt% SiO2; Cochrane, 2013; Fig. 17h) that lie in a similar structural position in Ecuador and Colombia, with lower trace element and REE
abundances, and they yield slightly enriched LREE relative to HREE
((La/Yb)n 2.34.6; Fig. 17i), while there is a general absence of distinctive Nb, Ta and Ti anomalies when normalised against N-MORB. The
trace element abundances plot close to unity when normalised against
N-MORB, and it is very likely that the LILE trends have been disturbed
by alteration. The gabbros and basalts plot within the calc-alkaline to
tholeiitic eld when comparing immobile elements (Fig. 16a), and a
majority of the rocks plot within the MORBseamount eld when comparing Nb/La against (La/Sm)n (Fig. 17c). Ndi (whole rock) values for
the Peltetec unit range between 1.1 and 1.2 (this study; Fig. 16b), and
they are less radiogenic than those obtained from other mac rocks
(4550 wt% SiO2; Cochrane, 2013) that lie in a similar structural position in Ecuador and Colombia. We interpret these data to suggest that
the protolith to these greenschist facies rocks formed either within an
ocean plateau, or perhaps within a back-arc basin setting, as transitional
crust.
5.3.3. Comparison with magmatic rocks from Peru
Granitoids of the Jurassic Ilo Batholith (U-Pb zircon 152173 Ma; see
above) within the Arequipa Terrane of southern Peru are metaluminous
to slightly peraluminous (Fig. 15a), and show a strong enrichment in
LILE with negative Nb, Ta and Ti anomalies (Boekhout et al., 2012).
Ndi (whole rock) values range between 5.1 and 2.4 (Fig. 16),
and Hfi (zircon) range between 9.5 and 7, suggesting that mantle derived melts assimilated a signicant amounts of heterogeneous Sunsasaged (~ 1 Ga) continental crust (Demouy et al., 2012; Boekhout et al.,
2013b), with perhaps variable depths of melting. The few inherited
xenocrystic cores in Jurassic intrusions corroborate the assimilation of
continental crust (Fig. 12b). These data are consistent with a subduction

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

26

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

76

75

74

73

195.81.5
Segovia
Batholith

Cretaceous (Early)

159.25.2

SJF

149.26.1
114.33.8
Barragn
120.70.3
Arqua
112.03.7

Metalluminous granitoids
Continental arc intrusions

City

IF

Ibagu
Batholith

100 km

3N

CAF

187.42.3
175.81.7
179.02.0
173.61.5

P
Chingul
Pluton

181.51.6

157.97.3

Rosa Florida
Batholith
0

LF
PF

80W

Raspas
126.44.0
129.95.6
Raspas
123.91.4
129.31.3
1325

Abitagua
Batholith
Y

Peltetec
134.70.9
134.313

182.40.6
169.81.1
173.01.3

CF

79W

189.12.9

Ibagu
Batholith

125.30.9

140.71.7
143.51.3
168.82.2
155.06.1

198.00.8
201.00.9
200.40.7

168.80.7
146.81.5
148.93.3
166.010
182.62.4
159.62.4
164.41.1
156.51.1
155.72.2
158.51.0

5N

Slates, quartzites
Guamote Sequence (Ecuador)

199.81.2
208.81.2
200.01.5
198.30.8
199.11.3

188.92.0

SPF

112.90.8

Jurassic - latest Triassic

Azafrn
Batholith

155.66.2

Metalluminous granitoids
Continental arc intrusions
Azafrn Batholith, Chingul Pluton
Upano Unit (basalts and turbidites)

78W
2N

196.01.1
201.11.4
B

Metalluminous volcanic rocks


Alao Arc (Ecuador)
Quebradagande Complex (Colombia)

202.21.0

Santander
Massif

8N

Ultramafic - mafic H-MP/LT


metabasalts, metagabbros.
Arqua, Barragn complexes (Colombia)
Raspas, Peltetec complexes (Ecuador)

163.71.6
143.39.9

CF

U-Pb zircon
LA-ICPMS

277.33.0

U-Pb zircon
TIMS

233.74.8

U-Pb zircon
(youngest detrital zircon)
LA-ICPMS
metasedimentary rocks

213.70.9

40Ar/39Ar plateau date


hornblende, muscovite,
phengite, plagioclase

244.62.4

Lu-Hf isochron

174.01.2

244.62.4

3S

PF

121.00.8

BF

178.11.4
145.40.2

131.61.1

153.81.5
160.51.7
Zamora
Batholith

5S

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

27

Passive E. Cret. arc


Margin?
Guamote
Alao Terrane
Sequence
121 S
PF V
V V
N

BF

Pz - Triassic
basement
Loja
Terrane

Jurassic - earliest Cretaceous arc

LF

CF
Upano Fm.

3 km

Abitagua Batholith

Salado Terrane

Tungurahua
volcano

159 S
E. Cret. arc
Passive
Margin?

Guamote Sequence

Alao Terrane
V

153 S
V

PF V V
V

V
V

159 S

Peltetec Unit

B
Age (Ma)

Cretaceous ultramafic - mafic complexes

250

Peltetec (Ecuador)
Raspas (Ecuador)
Arqua Unit (Colombia)
Barragn Unit (Colombia)

200
Gz

150

Cretaceous Volcanic Rocks


Alao Arc (Ecuador)

100
50

Quebradagrande Complex (Colombia)

100

200

300

Distance from suture (km)

400

Cretaceous Intrusions
Eastern Cordillera (Colombia)
Jurassic Intrusions
Azafrn and Chingul (Ecuador)
Zamora, Abitagua (Ecuador)
Ibagu, Segovia (Colombia)
Santander Massif (Colombia)

Fig. 11. A) Strip maps across the central Cordillera Real of Ecuador showing the terrane terminology of Litherland et al. (1994). B) Relationship between zircon U-Pb age and distance from
the Peltetec Fault. The Peltetec Fault is considered to represent the Jurassic-Early Cretaceous palaeo-margin, and is equivalent to the Silvia-Pijao Fault in Colombia. Citations for the age data
are provided in Table 2. Gz: Garzn Massif.

environment within continental crust, and in this sense they appear


to be similar to the Jurassic granitoid intrusions exposed within the
Northern Andes, which yield the same crystallisation ages, although
the Nd isotopic compositions of the Peruvian rocks along the coast of
Arequipa (Fig. 1, inset) are less radiogenic (Fig. 16b). The Ilo Batholith
is currently exposed along the coastline of southern Arequipa, whereas
coeval granitoid intrusions within the Northern Andes are exposed 30

150 km inland from present-day position of the Jurassic palaeo-margin


(Fig. 11b). Middle Jurassic nepheline syenites of the Allincapac igneous
complex in the Cordillera de Carabaya are more alkali than all Jurassic
igneous rocks found in the northern Andes (Fig. 15a, b), and formed
from SiO2 under-saturated magmas (Mikovi et al., 2009), which are
considered to have intruded through thinned continental crust in the
Jurassic back-arc (Mikovi et al., 2009).

Fig. 10. Geology of the Cordillera Real and Amotape Complex of Ecuador, and the Cordillera Central and Santander Massif of Colombia, showing the distribution of JurassicEarly Cretaceous
arcs, obducted macultramac rocks, and the Guamote Sequence (the Chaucha Terrane of Litherland et al., 1994). Exposure of the Peltetec Unit is too small to show on the map. The
Palaeozoic sequences and Triassic anatectites are shown for reference. Concordant Jurassic and Cretaceous zircon U-Pb, plateau 40Ar/39Ar and Lu-Hf dates and their uncertainties
(2; see references in Table 2) are shown. Sections Y-Y' and Z-Z' are shown (see Fig. 11). Cities, B: Bucaramanga, I: Ibagu, L: Loja, M: Medellin, P: Pasto, Q: Quito. Faults: BF: Baos
Fault, CF: Cosanga Fault; CAF: Cauca-Almaguer Fault, IF: Ibagu Fault, LF: Llanganates Fault, OPF: Ot-Pericos Fault, PF: Peltetec Fault, SJF: San Jeronimo Fault, SPF: Silvia-Pijao Fault.
Map compiled from Litherland et al. (1994) and Gmez et al. (2007).

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

28

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

The Coastal Batholith of Peru (~11537 Ma) is bimodal, and is dominated by metaluminous tonalites (N 70 wt% SiO2) with enriched LILE
that are typical of calc-alkaline rocks. Hfi values from the Coastal Batholith range between 5.2 and ~10.0 from rocks with U-Pb zircon ages of
~115 Ma (Polliand et al., 2005; de Haller et al., 2006; Schaltegger unpublished data) along the coastline in Central Peru, and between 5 and
7.5 (100110 Ma) within Arequipa (Demouy et al., 2012; Boekhout
et al., 2013b), which are less juvenile than a diorite of similar age
(~ 113 Ma; Cochrane, 2013) from the Quebradagrande Complex
(Table 2; Fig. 16c). Ndi (whole rock) from Early Cretaceous granitoids
along coastal Arequipa range between 4.4 and 0.1 (Boekhout et al.,
2012), which is also less radiogenic than magmatic rocks of a similar
age in the Northern Andes (Fig. 16b).
5.4. The tectonic setting during the latest TriassicJurassic (210145 Ma)
A combination of eld studies, concordant zircon U-Pb dates, geochemistry and isotopic data clearly shows that an I-type, metaluminous,
high-K to calc-alkaline continental arc formed within northwestern
South America at ~ 209 Ma, due to subduction of Pacic oceanic lithosphere beneath western South America. This time period marks the formation of a new subduction zone inboard of the Central American
terranes that were the Permo-Triassic conjugate margin to northern
South America within Pangaea. Arc magmatism during 194209 Ma
was focussed within the rocks of the Santander Massif, currently located
280350 km inboard of the Silvia-Pijao Fault (Fig. 18a). The Peltetec
Silvia Pijao Fault is the western boundary of the Quebradagrande Complex, and is considered here to represent the JurassicEarly Cretaceous
continental margin (Fig. 19). These early arc magmas assimilated large
quantities of continental crust, and are highly enriched in LILE, LREE
and non-radiogenic Nd and Hf isotopes. These rocks were coeval with
arc magmatism within southern Peru. The lack of arc magmatism within
Venezuela during 194209 Ma suggests that it was located more than
400 km from the Pacic active margin, while the inter-American gap
had not yet opened.
The arc axis migrated westwards after ~194 Ma, giving rise to a longlived calc-alkaline, metaluminous arc (141189 Ma) which is dened
by the Segovia, Ibagu, Rosa Florida, Abitagua, Azafrn, Chingul and Zamora batholiths throughout Colombia and Ecuador, and several volcanic
formations (e.g. the Misahuall Fm. in Ecuador; Figs. 18a and 19). These
intrusions yield more radiogenic Hf and Nd isotopic compositions
(Fig. 16c), and are more metaluminous (Fig. 15a) than during 194
209 Ma suggesting that they either are derived from a more juvenile
source, or have assimilated less continental crust. A comparison of age
and longitude suggests that these intrusions are also young to the
west over a (present day) distance of ~100 km (Figs. 11b and 18a), traversing the cordilleras Real and Central, with the oldest intrusions occurring in the Segovia Batholith (Colombia), and the youngest in the
foliated Azafrn Batholith (Ecuador). This trend suggests that the arc
axis migrated to the west during 189141 Ma, albeit at a much slower
rate than during 194189 Ma. Westward migration of the Middle
Late Jurassic arc axis by 100 km is similar to the quantity measured by
Boekhout et al. (2012) and Demouy et al. (2012) in the region of Arequipa (from the city of Arequipa towards the coast), although migration at
that location is considered to have started at ~175 Ma. Migration of the
arc axes may be due to either slab steepening and migration relative to
the trench, or migration of the trench and subducted slab to the west,
extending the margin. Given that the younger intrusions yield more
juvenile Nd and Hf isotopic compositions, and are located further to
the west, we suggest that the crust was extending, resulting in progressively less continental contamination of magmas derived from the mantle wedge above a subduction zone that was retreating. Unlike the other
Jurassic intrusions, which are generally unfoliated, the younger and
earliest Cretaceous Azafrn and Chingul intrusions are foliated, and
are located to the west of the Cosanga Fault in Ecuador (Fig. 10). We
hypothesise that the Cosanga Fault may have originated within an

extensional system. This interpretation is consistent with i) Sempere


et al. (2002), who utilize sedimentary facies to demonstrate that
the margin of Arequipa was extending throughout the Jurassic, and
Boekhout et al. (2012) report subsidence rates of ~3.5 km/My, ii) Jurassic back arc magmatism through thin continental crust (173195 Ma;
Mikovi et al., 2009) in the Cordillera de Carabaya, southern Eastern
Cordillera of Peru, and iii) reconstructions of Pindell and Kennan
(2009), who draw rift zones extending from Central Peru towards
northern Colombia throughout the Jurassic period (Fig. 18b). Jurassic
extensional basins are recognised to the east of the Jurassic intrusions,
which are characterised by grabens with varying rates of subsidence,
along with marine environments (e.g. Bogota Basin) in Colombia
(Fabre, 1984; Toussaint and Restrepo, 1994). Subsidence gave rise to a
2000 m thick sequence of Sinamurian limestones and sandstones of
the Santiago Fm. (Litherland et al., 1994) in the southern Sub-Andean
Zone of Ecuador (Fig. 1), and in northern Peru (Jaillard et al., 1990).
The oldest crystallisation ages at any given location (Fig. 12a) show a
general younging trend from northern Colombia to southern Ecuador,
suggesting that the onset of subduction may have been diachronous
north of the Huancabamba Deection. However, arc magmatism in
Peru may have started as early as 216 Ma, and occurred until b 135 Ma
(Fig. 18a), as recorded by the Chocolate Fm. and the Ilo Batholith.
Calc-alkaline, metaluminous arc intrusions along coastal Arequipa
were accompanied by coeval, alkali back-arc magmatism in the
Cordillera de Carabaya during 173195 Ma (Fig. 18a). This time period
broadly corresponds with a westward jump of the arc axis in the Northern
Andes during 194189 Ma, although coeval arc-back arc relationships
have not been found in the Northern Andes. Clearly, the potentially
diachronous onset of arc magmatism in the Northern Andes cannot be
extended across the Huancabamba Deection.
The tT paths generated from apatite U-Pb dates (Fig. 7) shows
that the current surface of the Northern Andes was cooler than 380 C
throughout the Jurassic, and thus very little information can be obtained
from the U-Pb data. One sample reveals reheating to temperatures N 380
C commencing at a poorly constrained interval time of ~ 150 Ma,
whereas the other sample reveals no reheating into the apatite lead
partial retention zone (APbPRZ; ~ 380550 C). We interpret this as
burial of fault blocks in southern Ecuador at ~ 150 Ma, whereas other
fault blocks were not buried, depending on their structural setting
within the extensional system.

5.4.1. Why is there a gap in the Jurassic arc in Peru?


The Jurassic arc is recorded along the entire length of the Northern
Andes, although it is missing between ~ 6S and ~ 12S (northern
Peru). This may be because an arc never formed in northern Peru
(Fig. 18a), implying there was insufcient subduction to create
magma due the angle of the South American Plate margin relative to
the plate convergence direction. Alternatively, there was a Jurassic arc
along the entire Peruvian margin, but it was removed from northern
Peru by either i) tectonic erosion, or ii) lateral displacement. The Jurassic
arc axis in Arequipa crops out along the coastline, ~ 50 km from the
present-day trench, and thus we hypothesise that the Jurassic forearc
has been removed, and did originally form along the entire margin.
The Peruvian and northern Chilean margin has undergone rapid tectonic erosion since at least the Eocene (Clift et al., 2003; Clift and Hartley,
2007). The highest rates are found north of the present day intersection
of the Nazca Ridge with South America, which coincides with the
northern extent of the Arequipa Terrane. Clift et al. (2003) estimate
that at subduction of the Nazca Ride beneath Peru has removed ~110
km of continental crust since the ridge started subducting at ~ 16 Ma
(Rosenbaum et al., 2005) or ~ 11 Ma (Hampel, 2002), and that 38 km
was eroded during the preceding ~35 Ma. The Nazca Ridge has not yet
subducted beneath the Arequipa Terrane, which may account for the
preservation of the Jurassic arc, and the spatial coincidence supports
this hypothesis.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

29

A
Zircon U-Pb
age (Ma)2
210

Ecuador

Colombia

Venez Peru

200

170

Segovia

Rosa
Florida

Zamora

Abitagua

Jurassic

180

160

Cordillera Real,
Cordillera Central

150

no igneous ativity

190

Santander
Massif
Garzn
Massif

140
130
4

South

Frequency
80
70
60

North

Frequency

Cordillera Real
Cordillera Central
Santander Massif
Peru (Arequipa)

4
3

50
40

30
1

20

206Pb/238U

age (Ma)

206Pb/238U

2060
2180

1340
1460
1580
1700
1820
1940

740
860
980
1100
1220

220

210

200

190

180

170

160

140

150

130

110

120

0
100

380
500
620

10

age (Ma)

Fig. 12. A) A comparison of latest TriassicJurassic concordant zircon U-Pb dates with latitude along the Cordillera Real, Cordillera Central, and the Santander and Garzn massifs of
Colombia. The ranges of concordant zircon U-Pb dates obtained from granitoid intrusions and volcano-sedimentary rocks from the Eastern Cordillera (Mikovi et al., 2009) and Arequipa
Terrane (Boekhout et al., 2012; Demouy et al., 2012) are shown for comparison. Data and citations are presented in Table 2. B) Age frequency histogram for LA-ICPMS in-situ data (single
spots) for the latest TriassicEarly Cretaceous. C) Age frequency histogram for LA-ICPMS in-situ data (single spots) for the DevonianPrecambrian, revealing a paucity of xenocrystic zircon
cores in the Jurassic arc.

An alternative explanation for the Jurassic gap may be provided by


palaeomagnetic data obtained from Jurassic tuffs (K/Ar dates only;
Bayona et al., 2010) from northern Colombia (Sierra Nevada de Santa
Marta and the Santander Massif; Fig. 1). These data suggest that the
tuffs acquired their permanent remnant magnetization at southern
palaeo-latitudes that corresponded with northern Peru and Ecuador
during the Early Jurassic, and the fault blocks were at least within the
vicinity of Colombia by the Late JurassicEarly Cretaceous. Bayona
et al. (2010) conclude that there has been signicant along-margin
northward translation of terranes relative to the Amazonian Craton.
Therefore, the hypothesis that the Jurassic arc of northern Peru may
have displaced northwards, and is currently partly exposed within
Colombia should also be tested. Northward migration of terranes during
the late Jurassic could also account for ~63 of clockwise rotation of the
Amotape Terrane (southern Ecuador; Fig. 1), which started at ~115 Ma
(Mourier et al., 1988; Jaillard et al., 1999). Clockwise rotation would

require a pivot point and a dextral shear force, which may have been
provided by the northward transcurrent migration of crust from Peru.
Within this context, Jurassic quartzites and slates of the Guamote
Sequence, located to the west of the Peltetec Fault in Ecuador (Fig. 11)
may be a remnant of a dismembered terrane (the Chaucha Terrane of
Litherland et al., 1994) that was displacing towards to the north. This
hypothesis is consistent with the reconstructions of Pindell and
Kennan (2009), who invoke a lateral translation of the Taham Terrane
of the Cordillera Central of Colombia, during the JurassicEarly Cretaceous (Fig. 18b). Lateral translation of terranes along active margins
has been documented in Nova Scotia, within the Rheic Ocean Wilson
Cycle (Gibbons et al., 1996). Factors that oppose this hypothesis are i)
the timing of northward displacement proposed by Bayona et al.
(2010) is based around K/Ar dates, which rarely record crystallisation
ages and probably do not relate to the time that the remnant magnetisation was acquired, and ii) our compilation of zircon U-Pb dates

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

3 sandstones
Youngest single zircon:
155 6.1 Ma, N = 262

20
N
10

Alao Arc (Ecuador)


14
1 sandstone
Youngest single zircon:
12
163.7 1.6 Ma, N = 58
10
N 8
6
4
2

7
6
5
N4
3
2
1
0

1 sandstone
Youngest single zircon:
143.3 9.9 Ma
N = 59

2700

Peltetec Fault

1 sandstone
Youngest single zircon:
149.2 6.1 Ma, N = 50

6
5
4
N
3
2
1
0

2400
2100
1800
1500

Age (Ma)

Upano Unit (Ecuador)

1200
900
600
300
0

2700
2400
2100
1800
1500
1200
900
600
300
0

Quebradagrande Complex (Colombia)


Silvia-Pijao Fault

Guamote Sequence (Ecuador)


30

Llanganates Fault

30

Age (Ma)

3000
2700
2400
2100
1800
1500
1200
900
600
300
0
Age (Ma)
Fig. 13. Age frequency histograms (zircon LA-ICP-MS) for three sandstones of the Guamote Sequence, and single sandstones from the Quebradagrande complex, Alao Arc and the Upano
Unit. The youngest U-Pb ages constrain the maximum stratigraphic ages of the rocks.

from the Jurassic arc in Colombia shows that they progressively


young to the west, and the Jurassic arc is not duplicated in Colombia
or Ecuador (Fig. 11b). Consequently, we conclude that the Jurassic arc
in Northern Peru either never formed, or has been removed by subduction erosion.

5.5. The tectonic setting during the Early Cretaceous (145115 Ma)
The compiled zircon U-Pb dates, geochemistry and isotopic data
clearly show that I-type, metaluminous, high-SiO2 (N 75%; Fig. 15) arc
rocks formed the Azafrn and Chingul Batholiths, and parts of what is

40Ar/39Ar date
2 (Ma)

36Ar/40Ar

1400

Metabasalt 09PR47

1200
1000

0.04
0.03

Inverse isochron age 115.35 43.66 Ma


40Ar/36Ar intercept 319.7 57.2
MSWD 2.47
step used to calculate age
step not used

800
0.02

600
134.2612.84 (4 steps)

400

0.01
200
0

20

40
%39Ar

60

80

100

0.00
0.00

0.02

0.04

0.06

39Ar/40Ar

released

40Ar/39Ar date
2 (Ma)

36Ar/40Ar

Metagabbro 09PR48

400

Inverse isochron age 134.12 8.65 Ma

0.04 40Ar/36Ar intercept 298.4 42.8

MSWD 2.25
0.03

300

0.02

134.710.89 (5 steps)

200

0.01
100

20

40

60

%39Ar released

80

100
0.00
0.00

0.02
0.04
39Ar/40Ar

0.06

Fig. 14. 40Ar/39Ar age spectra (plateau dates) and inverse isochron plots from plagioclase extracted from meta-basalts of the Peltetec Unit, Ecuador. Steps with older dates are dened by
gas that is contaminated with excess 40Ar (step heating data are presented in the supplementary les).

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

intrusions (Fig. 16). Scant geochronological data suggest that the


volcanic rocks within the Quebradagrande Unit erupted at ~ 114 Ma,
while the volcanoclastic rocks were deposited after ~149 Ma (Fig. 13).
No accurate age estimates of the Alao arc are available although sedimentation could have occurred after ~ 164 Ma (Fig. 13). K/Ar dates of
115 12 Ma and 142 36 Ma suggest that it could be coeval with
the Quebradagrande Complex. This tectonic correlation is supported
by its outboard position relative to the Palaeozoic basement, Triassic
anatectites and Jurassic intrusions, and we tentatively assign an Early
Cretaceous age to the Alao arc.
The Alao Arc hosts large volumes of quartz-zircon-tourmaline
rich arenites (Cochrane, 2013), and the U-Pb ages of detrital zircons
(Fig. 13) reveal a derivation from cratonic South America. Therefore,
the Alao Arc is interpreted to have formed above an east-dipping subduction zone along the thinned fringe of a continental margin, giving
rise to isotopically juvenile, mac volcanic rocks (Cochrane, 2013;
Fig. 19). Extension was sufcient to form MORB-like rocks with tholeiitic signatures above the subduction zone. Alternatively, the tholeiitic basalts may form part of the Peltetec Unit (Fig. 11), and were structurally
emplaced against the calk-alkaline rocks during subsequent compression. This interpretation differs from that of Litherland et al. (1994),
who draw the Alao Arc as a Middle Jurassic Island Arc, although this is
inconsistent with the presence of abundant Precambrian zircons. Similarly, the volcanic rocks of the Quebradagrande Complex are associated
with sedimentary rocks (Abejorral Fm.; E.g. Gmez-Cruz et al., 1995)
that host abundant quartz, and Nivia et al. (2006) interpreted this
sequence as a thinned marginal basin along a continental margin
(Fig. 19). Villagmez et al. (2011) suggest that the arc rocks erupted
through highly attenuated continental crust because i) some basalts
yield geochemical signatures that approach seamounts (T-MORB),

mapped as the Zamora Batholith (granodiorite 09RC43; Table 2; Fig. 19)


during 144132 Ma. These units are exposed to the west of the older arc
intrusions in Ecuador, are bound against the Cosanga Fault (Fig. 11)
to their east, and exhibit a weak to strong foliation (Pratt et al., 2005).
The Hf and Nd isotopic compositions (Fig. 16) indicate that they
crystallised from melts that were more juvenile than the older, unfoliated Jurassic intrusions and they are considered by Cochrane
(2013) to have erupted through thinned crust (Fig. 19). This interpretation is consistent with Pratt et al. (2005) who suggest that the foliations
formed during pure shear, and Litherland et al. (1994) who suggest that
magmatism within the Salado Terrane occurred within a marginal
basin (Fig. 19). U-Pb dates of detrital zircons sandstones of the Upano
Unit suggest that this marginal basin was forming after 143 Ma
(Fig. 13), which is consistent with the crystallisation ages of the Azafrn
and Chingul plutons. Extension during 144132 Ma coincides with a
region-wide angular unconformity in the Upper Magdallena Valley
Basin (Fig. 1; Jaimes and de Freitas, 2006), the Oriente Basin in
Ecuador (Balkwill et al., 1995) and in northern Peru (Jaillard et al.,
1990). The older Jurassic intrusions to the east (N 145 Ma; e.g. the
Abitagua Batholith) were not affected by the extensional system, and
remain unfoliated.
Mac, calk-alkaline to tholeiitic volcanic rocks of the Alao and
Quebradagrande Complexes are preserved to the west of the Azafrn
and Chingul Batholiths, and west of an uplifted core (Pratt et al.,
2005) of Palaeozoic basement and Triassic migmatites, via the Baos
Fault (Fig. 11). Major, trace element and REE compositions suggest
that these rocks formed within arcs (Fig. 17), although some samples
yield an afnity with MORB. The volcanic rocks are located outboard
of the Jurassic intrusions (Fig. 11b), and yield juvenile Hf and Nd isotopic
compositions and depleted trace elements relative to the Jurassic

C
4.0

10000
Ecuador (Abitagua, Zamora, Rosa Florida)

Peraluminous

Peru (Ilo Batholith)


Jurassic arc

3.5

Ecuador (Azafrn, Chingul)


Colombia (Ibagu, Segovia)
Colombia (Santander Massif)

1000

Metalluminous

3.0

Al/(Na+K)

31

2.5
2.0
1.5

100
Ecuador, Colombia
Triassic anatec tes

10
1

1.0
Peru (Allincapac) Jurassic back-arc

0.5
0.5

1.0

1.5

2.0

2.5

3.0

0.1

Al/(Ca+Na+K)

Cs Rb Ba Th U Nb Ta La Ce Pb Pr Sr Nd Zr Hf Sm Eu Ti Tb Y Tm Yb

100
Azafrn, Chingul Batholith
Rosa Florida, Abitagua, Zamora
Ibagu, Segovia

Na2O + K2O - CaO (wt%)

10

10
Peru (Allincapac)
Jurassic back-arc

8
6

4
Alkalic

Al

ial

ca

lci

Ca

-2
50

55

lc-

0.1

ka
al

lin

La

e
Calcic

60

65

Ce

Pr

Nd

Sm

Eu

Gd

Tb

Dy

Ho

Er

Yb

Lu

Peru (Ilo Batholith)


Jurassic arc

SiO2 (wt%)

70

75

80

Fig. 15. Geochemical data from Jurassic granites and granodiorites from the Cordillera Real, Cordillera Central and the Santander Massif (Litherland et al., 1994; Romeuf et al., 1995;
Bustamante et al., 2010; Cochrane, 2013; Van der Lelij, 2013). Fields are shown for Jurassic intrusions within the Arequipa Terrane (Boekhout et al., 2012; Demouy et al., 2012) and the
Allincapac Complex Mikovi et al. (2009) of Peru. Multi-element plots are normalised to N-MORB (citation).

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

32

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

B
Ndi (whole rock)

A
Th (ppm)

100

Gz

10

High-K calc-alkaline

Gz

0
Calc-alkaline

1
,
esite
and ite
r
altic
Bas site, dio
e
and
e,
dacit iorite
lite,
d
Rhyo e, grano
it
gran

-2

0.1

0.01
0

20

-4

Peru
Arequipa Terrane

-6

Tholeiite

40
Co (ppm)

Peru
Coastal Batholith
Arequipa Terrane

60

80

-8
100

120

160

140

180

200

Time (Ma)

C
Hfi (zircon)

16
14

(La/Yb)n
35

DM
Peru
Arequipa Terrane
and back-arc

12
10
8
6
4
2

30
25
20
15

0
-2
-4

10

-6
-8
100

Peru
Coastal Batholith
Arequipa Terrane and
North of Arequipa

5
0

120

140

160

180

200

Time (Ma)

220

-10

-5

Ndi (whole rock)

10

15

Cretaceous ultramafic - mafic complexes


Peltetec (Ecuador)
Raspas (Ecuador)
Arqua Unit (Colombia)
Barragn Unit (Colombia)
Cretaceous Volcanic Rocks
Alao Arc (Ecuador)
Quebradagrande Complex (Colombia)
Cretaceous Intrusions
Eastern Cordillera (Colombia)
Jurassic Intrusions
Azafrn and Chingul (Ecuador)
Zamora, Abitagua (Ecuador)
Ibagu, Segovia (Colombia)
Santander Massif (Colombia)
Fig. 16. Geochemical and isotopic data from Jurassic intrusions, Early Cretaceous volcanic rocks, and M-HP/LT mac and ultramac rocks of the Amotape Complex and Cordillera Real of
Ecuador, and the Cordillera Central, Santander Massif and the Eastern Cordillera of Colombia. See captions for Figs. 15 and 17 for citations. Fields for the Coastal Batholith of Peru are from
data in de Haller et al. (2006), Demouy et al. (2012), Boekhout et al. (2013b), and unpublished data from U. Schaltegger (University of Geneva). Fields for the Arequipa Terrane and Jurassic
back-arc of Peru are from data in Mikovi et al. (2009), Demouy et al. (2012), and Boekhout et al. (2013b).

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

ii) most volcanic rocks erupted in submarine conditions, and iii) no continental detritus is found to the west of the Quebradagrande Complex.
Pindell and Kennan (2009) draw the Quebradagrande Arc as an oceanic
arc above an east dipping subduction zone until 125 Ma, although this
date is not based on robust geochronological evidence from northwestern South America.
The simplest explanation for the geochemical, isotopic and geochronological trends obtained from the Jurassic and Early Cretaceous igneous rocks is that they formed above the same east-dipping subduction
zone, which was retreating oceanward during 209194 Ma, and after
~ 145 Ma until ~ 114 Ma (E.g. Cochrane, 2013; Fig. 19). Extension of
the continental crust during ~ 145114 Ma was sufcient to generate
mac magmas with T-MORB geochemical characteristics and marine
environments, which are found intercalated within the Alao and
Quebradagrande sequences. Kennan and Pindell (2009) refer to this
extensional feature as the Colombian Marginal Seaway. This time corresponds with heating of some fault blocks within the southern Cordillera
Real (Fig. 7a) to temperatures of up to ~500 C, which was synchronous
with a signicant increase in HFi (zircon) in the magmatic rocks.
Heating is interpreted to be a consequence of sedimentary burial during
extension, combined with an increase in geothermal gradients, which
may have been signicant in places due to the proximity of magma.
Other fault blocks (Fig. 7b; e.g. northern Colombia) were not being
heated at this time, presumably because they did not reside in a part
of the extensional system that was being buried. Finally, some faulted
units (e.g. south of the Ibagu Fault in the Cordillera Real; Villagmez
and Spikings, 2013) were cooling at ~140 Ma (Fig. 7c), perhaps because
they were exhumed during extension.
This interpretation is similar to that of Toussaint and Restrepo
(1994), Cooper et al. (1995), Sarmiento-Rojas et al. (2006) and Pindell
and Kennan (2009), who suggest that the rocks of the Eastern Cordillera
of Colombia (Fig. 1) underwent back-arc extension (NNESSW extensional axis; e.g. Mora et al., 2006), during the Cretaceous, associated
with an arc that formed rocks which are now exposed within the Cordillera Central (Quebradagrande Complex). Vsquez and Altenberger
(2005) and Vsquez et al. (2010) report that gabbroic intrusions exposed in the Eastern Cordillera formed within a rift setting during
135 Ma and 121 Ma (Table 2; Fig. 11). The MORB and OIB-type chemistry of these rocks (Vsquez et al., 2010) suggests that back-arc extension
was signicant.
The Guamote Sequence of Litherland et al. (1994) does not host any
magmatic rocks, and is composed of metasedimentary rocks that were
deposited after 155 Ma (Fig. 13). The Braziliano and Sunsas aged U-Pb
ages of detrital zircons shows that the sediments were derived from
cratonic South America. Furthermore, their detrital age signature is indistinguishable from that obtained from arenites within the Alao Arc,
Quebradagrande Complex and the Upano Unit (Fig. 13). The Guamote
Sequence is currently separated from the Alao Arc by ultramacmac
rocks of the Peltetec Complex, although it is not unreasonable to suggest
that it once formed a part of the South American Margin, and was
emplaced either by i) strike-slip displacement from more southern
latitudes (see Section 5.7), or by ii) rifting away from north-western
South America, followed by re-accretion (Fig. 19).
Litherland et al. (1994) suggest that prior to 140 Ma, continental
crust of the Chaucha Terrane lay outboard of a west-facing island arc
(Alao Arc), which was separated from South America by oceanic crust
that was subducting beneath South America, forming the Azafrn
Pluton. Subsequently, the same authors suggest that these terranes collided together during 140120 Ma, during the compressive Peltetec
Event. Similarly, Villagmez and Spikings (2013) suggest that the collision of a series of seamounts or oceanic plateau blocked the Jurassic subduction zone, terminating the Jurassic arc, and that a new subduction
zone formed outboard of the hypothetical plateau, forming the westfacing Quebradagrande Arc. However, the data compiled here suggest
that i) arc magmatism during the JurassicEarly Cretaceous formed
above a single subduction zone, ii) the Quebradagrande Complex was

33

a continental arc that erupted through thinned continental crust, and


iii) extension prevailed during the Early Cretaceous. No evidence exists
for a Jurassic oceanic Plateau within Colombia or Ecuador.
The Arqua, Barragn, and Raspas complexes are considered to be
equivalent because i) they are located in similar structural positions,
outboard of Triassic anatectites and volcanic arc rocks, ii) they are composed of basalts and gabbros that have been metamorphosed to varying
degrees under M-HP/LT conditions, iii) they yield peak metamorphic
and retrogression dates that are similar, and iv) they yield MORB
seamount geochemical signatures. The Raspas Complex now forms
part of the para-autochthonous Amotape Terrane, which detached
from rocks that are now exposed in the Cordillera Real. The Raspas Complex hosts HP-LT rocks of oceanic plateau and MORB afnity (Arculus
et al., 1999; Bosch et al., 2002; John et al., 2010), which were exhumed
following peak metamorphism at 130126 Ma (John et al., 2010).
Geochemical and isotopic compositions also suggest that the Barragn
(HP-LT) and Arqua (MP-LT) units originated as MORB and seamounts
(Bustamante et al., 2012), while their 40Ar/39Ar dates suggest that
they were exhuming and cooled below ~ 400 C during 120112 Ma.
We consider these units to have been metamorphosed within the
same subduction zone, which was exhumed to variable degrees alongstrike of northwestern South America (Fig. 19). Exhumation could
have been triggered by forced return ow (Gerya et al., 2002), which
may have occurred during a compressive event, combined with the
inherent buoyancy of oceanic plateau rocks relative to MOR-derived
lithosphere.
Rifting of northwestern South America during the Early Cretaceous
may have been sufcient to detach continental slivers (e.g. the Chaucha
Terrane in Ecuador), and we suggest that the greenschist facies ultramacmac rocks of the Peltetec Complex formed during advanced
continental rifting as E-MORB crust. The Peltetec Complex has not
been metamorphosed in a subduction zone, and its current structural
position between the Guamote Sequence and the Alao Arc in Ecuador
indicates they were obducted during a compressional event. Their
40
Ar/39Ar dates suggest that they formed at ~ 134 Ma, and hence they
formed before the subduction channel that hosted the Arqua, Barragn
and Raspas complexes was exhumed.

5.6. Compression during the Early Cretaceous


Dense mineral assemblages and detrital zircon ssion track dates
with very short lag-times suggests that rocks deposited within the
Upper Magdalena Valley (Fig. 1; Vergara and Prssl, 1994; Sarmiento
and Rangel, 2004) and Oriente (Ruiz et al., 2007; Martin-Gombojav
and Winkler, 2008) basins during 120115 Ma were derived from an
exhuming cordillera located to the west. Restrepo et al. (2009) suggest
that Aptianmiddle Albian sandstones of the Abejorral Fm., which are
located along the western ank of the Cordillera Central, were derived
from a proto-Central Cordillera to the east. TimeTemperature paths
for some faulted blocks suggest that they cooled rapidly at 117
107 Ma (Fig. 7c), while other faulted units appear to have continued
the heating trend from earlier in the Cretaceous (Fig. 7a). Villagomez
et al. (2013) interpret these data as evidence for compression, which exhumed some fault blocks, forming a proto-Cordillera Central. This time
period coincides with the timing of retrogression of the east-dipping
subduction zone during exhumation, and compression at this time
obducted blueschists and eclogites of the Raspas, Arqua and Barragn
complexes onto the South American margin. Compression may have
occurred due to an increase in the convergence rates of the oceanic
Caribbean Plate (Kennan and Pindell, 2009) and South America, as a
consequence of the opening of the South Atlantic, which started at
~ 120 Ma (Eagles, 2007). Detrital zircon ssion track dates from the
Oriente Basin of Ecuador (Fig. 1; Ruiz et al., 2004) suggest that a second
compressive event may have occurred within Ecuador at ~ 100 Ma,
which would be coeval with compression in Peru (Mgard, 1984).

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

34

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

8
7
6
5
4
3
2
1
0

La/Yb
30

Peraluminous
Metalluminous

Al/(Na+K)

Triassic
anatectes
Jurassic
arc

25

Jurassic
Batholiths

20

Calc-alkaline
arc
Azafran
Batholith

15

0.5

1.0

1.5

2.0

Alao
Quebradagrande

10

2.5

Al/(Ca+Na+K)

Barragn
Arqua
Raspas
Peltetec

Island arc
tholeiite
Ocean plateau tholeiite

Nb/La

1.6
1.4

E-MORB

seamounts

200

1.2

400
Zr/Th

600

800

1.0
0.8

Quebradagrande Complex

1000

MORB

Alao Arc

0.6
connental arc

0.4

100

oceanic
arc

0.2
0
0

0.5

1.0

1.5

2.0

2.5

3.0

(La/Sm)n

E
10

10

0.1
1

0.1

N-MORB
Cs Rb Ba Th U Nb Ta La Ce Pb Pr Sr Nd Zr Hf Sm Eu Ti Tb Y Tm Yb

N-MORB
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Yb Lu

F
Barragan Complex
100

Arqua Complex

Blueschist
Amphibolite
Amphibolite

10

1
10

0.1

N-MORB
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Yb Lu

0.1

N-MORB
Cs Rb Ba Th U Nb Ta La Ce Pb Pr Sr Nd Zr Hf Sm Eu Ti Tb Y Tm Yb

Raspas Complex

Blueschist
Eclogite

Peltetec Complex

Basalt, gabbro

100

10

10

1
0.1

N-MORB
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Yb Lu

0.1 N-MORB
Cs Rb Ba Th U Nb Ta La Ce Pb Pr Sr Nd Zr Hf Sm Eu Ti Tb Y Tm Yb

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

Maresch et al. (2009) report metamorphic zircon dates of 116


106 Ma from anatectites within HP-LT metagabbros at La Rinconada,
Margarita Island (southern Caribbean). Those rocks are interpreted as
back-arc basin crust that was subducted and retrogressed during 116
106 Ma. The coincidence in the timing of metamorphism in northwestern South America and Margarita Island suggests that they may have
formed in the same tectonic setting. However, Maresch et al. (2009)
suggest that the HP-LT rocks at Margarita Island formed on a westdipping slab of continental back-arc crust. The same arc polarity is
shown by Pindell et al. (2005) at 119 Ma, although Pindell and
Kennan (2009) suggest that the polarity of the subduction zone ipped
during 125120 Ma, during the transition from their Trans-American
Arc, to the Caribbean Arc, and the separation of North America from
Gondwana. However, it is difcult to account for the outboard position
of M-HP/LT metamorphic rocks relative to the Alao and Quebradagrande arcs within an east-facing arc system, and the simplest explanation of the very large amount of data obtained from Jurassic and Early
Cretaceous rocks of the northern Andes suggest that subduction was
east-dipping beneath South America.
5.7. The Chaucha Terrane and the Taham Terrane
As previously described, the Chaucha Terrane includes the Guamote
sequence (Fig. 11), which hosts quartzites and slates that contain Jurassic ammonites, some of which were deposited after 155 6.1 Ma
(Fig. 13). These rocks lie outboard of the Peltetec Fault, and probably
represent a rifted fragment of South America, which re-accreted during
compression at 120110 Ma, or perhaps at a later time. The N-S extent
of these units beneath the volcanic cover rocks is unknown, and other
suspect occurrences include i) a small inlier of slates within the northern Interandean Depression in Ecuador (Fig. 1; near the town of
Chota, 015N), ii) a faulted sliver of graphitic schists in the southern
Western Cordillera of Ecuador (230S), and iii) undated granites of
the Manu Inlier, located between the Amotape Terrane and the Eastern
Cordillera of Ecuador (330S). The Chaucha Terrane is shown on the
tectonic reconstructions of Litherland et al. (1994) as a large continental
block that is exotic to South America. Similarly, Pindell and Kennan
(2009) show the Chaucha Terrane on their reconstructions (Fig. 18b)
as a continental microplate, with sufcient dimensions to displace the
Jurassic subduction zone westwards by N100 km. However, the distance
between their Jurassic subduction zone and South America (Fig. 18b) is
too large to account for the distribution of Jurassic arc rocks within
Ecuador. We consider all of these exposures to be para-autochthonous
to South America, and equivalent to lithologies exposed to the east of
the Peltetec Fault. Their current dispersion over a strike-distance of
~ 300 km may be a result of strike-slip segmentation of several rifted
continental slivers during the amalgamation of the Caribbean Large Igneous Province at ~ 75 Ma (Vallejo et al., 2006; Spikings et al., 2010),
and we do not consider these rocks to belong to a separate terrane,
and use of the term Chaucha Terrane is misleading. In our interpretation, the latest JurassicEarly Cretaceous Guamote sequence formed
within the forearc of the Alao Arc and Upano Units during extension,
above an east-dipping subduction zone.
The Taham Terrane (Antioquia Terrane in Pindell and Kennan,
2009) of northern Colombia (Restrepo and Toussaint, 1982) is widely
considered to be fault bounded by the Ot-Pericos and Palestina faults
to the east, the San Jeronimo Fault to the west, and the Ibagu Fault in
the south (Fig. 1). It is considered to be a distinct terrane because i) it
is separated from Sunsas-aged (~1 Ga) crust in the east, and ii) it does
not expose Jurassic arc rocks, unlike the region to the east. The basement rocks in this suspect terrane yield Ordovician (Villagmez et al.,

35

2011) and Triassic (Cajamarca Complex; e.g. Cochrane et al., 2014a)


U-Pb dates, which are intruded by the ~ 90 Ma Antioquia Batholith.
However, in our opinion these are insufcient to assign a terrane
status to these rocks, and it is not proven that they are derived from
several hundred km to the south, as is advocated in many models (e.g.
Kennan and Pindell, 2009). The simplest interpretation of the data presented here is that the current west-to-east juxtaposition of M-HP/LT
rocks, an Early Cretaceous arc, Palaeozoic and Triassic basement, latest
Jurassic arc and Jurassic arc, is a consequence of JurassicEarly Cretaceous attenuation of the margin, followed by a switch to compression
starting at ~ 120115 Ma. The assemblage of Palaeozoic and Triassic
rocks in Colombia is identical in composition, eld relationships
and age to those found throughout the Loja Terrane (a term used by
Litherland et al., 1994) of Ecuador, and they are interpreted here as
basement that was exhumed during JurassicEarly Cretaceous extension. Jurassic intrusive rocks are not found in these regions because
they did not overlie the arc axis at that time (Fig. 19). Furthermore, UPb zircon ages spanning between 900 and 1700 Ma, with a peak at
~ 1200 Ma (orthogneiss, La Miel Unit; Villagmez et al., 2011) within
Ordovician rocks suggest that Precambrian rocks may underlie this
suspect terrane. Consequently, we do not include these rocks as terranes
in our reconstruction (Figs. 18a and 19). Similar to the case of the
Chaucha Terrane, models which place the Jurassic subduction zone outboard of a hypothetical Taham Terrane (e.g. Pindell and Kennan, 2009;
Fig. 18b) create a distance between the trench and the Jurassic arc in
South America that is too large.
5.8. Comparison with Peru (145115 Ma)
With the exception of far northern Peru, no intrusive or volcanic
rocks have been documented within Peru that yield zircon U-Pb
dates that lie between 145 and 115 Ma. The period between 145 and
110 Ma was characterised by extensional tectonic events along the
Peruvian margin, generating deep sedimentary basins in northern
Peru during 145130 Ma (Chicama Basin), and in Central Peru during
130110 Ma (Jaillard and Soler, 1996). Kennan and Pindell (2009)
draw highly oblique and sinistral plate convergence directions between
South America and the Farallon Plate at 130 Ma, which would result in
slow subduction north of the Huancabamba Deection, and no net
convergence along coastal Peru, accounting for a lack of magmatism in
that region. Pindell and Kennan (2009) draw orthogonal subduction
between their Caribbean Plate and the Peruvian coastline at 125120
Ma, although this is unlikely given the lack of arc rocks of this age.
The Aptian calc-alkaline Celica continental arc (Feininger and
Bristow, 1980; Lebrat et al., 1987) formed in northwestern Peru and
far southwestern Ecuador, and its forearc Celica-Lancones Basin is preserved (Fig. 1; Jaillard et al., 1999). The magmatic rocks are poorly
dated, and they may have formed above the same subduction zone
that gave rise to the Quebradagrande and Alao arcs. These rocks may
represent the earliest manifestation of magmatism in Peru that lead to
the formation of the Coastal Batholith starting at ~115 Ma.
6. The tectonic history of northwestern South America during
11575 Ma
Only small volumes of magmatism are recorded in Ecuador and
Colombia during ~115100 Ma, which corroborates the highly oblique
dextral convergence angles of the Caribbean Plate with South America
(Pindell and Kennan, 2009). Barragn et al. (2005) report plateau
(whole rock) 40Ar/39Ar dates of 11082 Ma from low volumes of alkali
basalts in the Oriente Basin of Ecuador (Fig. 1). The same authors

Fig. 17. Geochemical data from Early Cretaceous volcanic rocks (Litherland et al., 1994; Nivia et al., 2006; Villagmez et al., 2011; Cochrane, 2013; Rodriguez and Zapata, 2013) and
M-HP/LT mac and ultramac rocks (Litherland et al., 1994; Arculus et al., 2002; Bosch et al., 2002; John et al., 2010; Villagmez et al., 2011; Bustamante et al., 2012; Cochrane, 2013)
of the Amotape Complex and Cordillera Real of Ecuador, and the Cordillera Central of Colombia. La/Yb and Zr/Th tectonic discrimination elds are from Jolly et al. (2001), and the
Th-Co classication of igneous rocks and tectonic environments is based on Hastie et al. (2007). Multi-element plots are normalised to N-MORB (Sun and McDonough, 1989).

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

36

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

189 - 145 Ma

209 - 194 Ma
North
America

Yucatan

Yucatan

20N

Central and
southern
Mexico

Santander
Massif
20N
S

P-SP

120mm/a

10N

120mm/a

Arc exposed today


Arc not formed, burried
or tectonically eroded

10N

A
Z

Present day coastline


0N

0N

Arequipa
Terrane

CC
IB
Nazca
Ridge

300km

100W

Nazca
Ridge

300km

90W

100W

90W

190 Ma

158 Ma
North
America

Central and
Southern
Mexico

Yucatan

Yucatan

20N

Ch
20N
120mm/a

120mm/a

Ch

T
T

10N

10N

C
C

0N

0N

1000 km

1000 km
110W

100W

90W

110W

100W

90W

Fig. 18. A) Subduction zones along western Pangaea during the Jurassic determined using an arc-trench distance of 300 km, and constant slab-dip. These subduction zones are derived
assuming that the Tahami Terrane in autochthonous. Palaeopositions, plate motion and reconstructions for Yucatan and central and southern Mexico are taken from Pindell and Kennan
(2009), and reconstruction at 189145 Ma is from their reconstruction for 158 Ma. Black arrows indicate amount of lateral migration of the subduction zone between 194 and 189 Ma
(Colombia and Ecuador) and after ~175 Ma (southern Peru). Grey line is present day coastline, and position of the Nazca Ridge. A: Abitagua Batholith, CC: Cordillera de Carabaya, I: Ibagu
Batholith, IB: Ilo Batholith, P-SPF: Peltetec-Silvia Pijao Fault (this is the JurassicEarly Cretaceous palaeomargin; Vallejo et al., 2006), S: Segovia Batholith, Z: Zamora Batholith. B) Subduction zones and reconstruction of Pindell and Kennan (2009), which assume that the Taham Terrane of Colombia is allochthonous, and that the Chaucha Terrane (Litherland et al.,
1994) exists. Blue line is a rift axis. C: Chaucha Terrane, Ch: Chortis Block, T: Taham Terrane.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

189-145 Ma
Continental arc volcanism

37

145 - 141 Ma
Extension of the continental margin
Foliated
arc (E.g. Azafrn Batholith)

Jurassic Batholith
(E.g. Abitagua, Ibagu)

E
rollback
active
arc

active
Triassic anatectites,
arc
undifferentiated
Palaeozoic rocks

141-115 Ma Margin attenuation, westward migration of arc, intra-arc basins, rifted continental slivers?
Guamote
Sequence

Peltetec
Unit

Alao Arc

Upano
Unit (marine)

rollback

Raspas Complex
Arqua Complex
Barragn Complex

Quebradagrande
Complex
marine

rollback
active
arc
asthenospheric
upwelling forming
transitional crust

active
asthenospheric
arc
upwelling forming
transitional crust

115-100 Ma Closure of fore/intra/back arc basins, obduction of M-HP/LT rocks, rock uplift and exhumation
Future Caribbean
suture (75 Ma),
Ingapirca Fault
compression

Peltetec Fault Zone

Cosanga
Fault
marine

Future Caribbean
suture (75 Ma),
Cauca-Almaguer
Fault

Silvia-Pijao
Fault

Ot-Pericos
Fault
marine

compression

Fig. 19. Schematic models for the tectonic evolution of the northwestern South American margin, which t the geochronological, geochemical and sedimentological data. These models
propose that the Jurassic arc axis during 185145 Ma did not drift. Roll-back starting at 145 Ma caused the arc axes to migrate oceanward and thinned the crust, leading to calk-alkaline and
tholeiitic arc magmatism, occasionally T-MORB basalts and marine sedimentary environments. Extension in some parts of the margin may have caused continental slivers to rift
away, forming extensive tracts of transitional crust (Peltetec Unit) in intra-arc basins, perhaps accounting for the Guamote Sequence. Compression starting at 120 Ma obducted exhumed
M-HP/LT rocks outboard of the Alao and Quebradagrande arcs, and entrained ultra-mac and mac rocks of the Peltetec Unit between those arcs and re-accreted continental slivers.

suggest that the basalts formed during asthenospheric upwelling


caused by the detachment of an east-dipping slab subsequent to terrane
collision, implying that an east-dipping subduction zone existed between an accreting terrane and South America. Our model suggests
that the long-lived JurassicEarly Cretaceous subduction zone simply
retreated westwards, and was not pinched-off between two crustal
blocks (Fig. 19). We suggest that if any asthenospheric upwelling occurred, it was driven by steepening of the slab during its prolonged residence in the mantle in the absence of subduction after ~110 Ma, along a
highly oblique margin.
Plutons of the Coastal Batholith in Central Peru and coastal Arequipa
started intruding at ~115 Ma. The magmatic units along the coastline of
central Peru are considered to have formed by shallow melting of
mantle-derived basalts (Atherton, 1990; Atherton and Petford, 1996),
which were originally derived from partial melting in the mantle
wedge above a subduction zone (e.g. McCourt, 1981). The intrusions
crystallised within a deep (~10 km) marginal basin that formed during
continental margin extension over an along-strike distance of ~1600 km
(Atherton and Petford, 1996), which formed the Casma marginal basin

during the Albian. Extension is supported by gravity data, which reveal


an arch-like structure of dense material within a rift setting, which is
considered to be imbricated oceanic crust (Jones, 1981). Jaillard and
Soler (1996) report that extension was punctuated with compressional
events, and that volcanic activity which fed the Casma Basin ceased by
the Late Albian, during the compressional Mochica Phase of Mgard
(1984). Abundant subduction related magmatic activity along Peru
starting at ~ 115 Ma is accounted for in the reconstructions of Pindell
and Kennan (2009) by the north-east directed subduction of their
newly formed Caribbean Plate beneath South America, forming a
mainly strike-slip boundary north of the Huancabamba Deection,
and amagmatic tectonics in the northern Andes.
Acidic magmatism in the Northern Andes between 100 and 75 Ma is
dominated by the large Antioquia Batholith (Fig. 10; concordant zircon
U-Pb ages of 9585 Ma; Villagmez et al., 2011; Fig. 20) in the northern
Cordillera Central, which intrudes through undifferentiated Palaeozoic
rocks and the Triassic Cajamarca Fm. Other scattered continental arc
intrusions that yield zircon U-Pb ages in this age range include the
Crdoba Batholith (80 Ma; Villagmez et al., 2011) of the Cordillera

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

38

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

Central, and the Tangula Batholith (zircon U-Pb 92 1 Ma;


Schtte et al., 2010) of the southern Cordillera Real, which intrudes
continental crust. Villagmez et al. (2011) and Villagmez and
Spikings (2013) utilise geochemical data to suggest that these intrusions formed by minor subduction of oceanic crust beneath South
America. Pindell and Kennan (2009) utilise plate reconstructions to
suggest that the Antioquia Batholith formed too close to the oceancontinent boundary (b 100 km) to be arc magmas, and instead
formed by melting of a slab-tip during subduction initiation,
although they acknowledge that this is speculative.
6.1. The formation of the Caribbean Large Igneous Province and its collision
with South America.
6.1.1. Geochemistry and geochronology
The basement of the forearcs and Western Cordillera of Ecuador and
Colombia consists of ultramac and mac rocks, and their chemical
compositions suggest that they formed above an oceanic hot-spot
(Fig. 1; Reynaud et al., 1999; Lapierre et al., 2000; Hughes and
Pilatasig, 2002; Kerr et al., 2002; Mamberti et al., 2003, 2004; Kerr and
Tarney, 2005; Hastie and Kerr, 2010). Within Ecuador, the bulk of the
plateau rocks are represented by the Pion (forearc), Pallatanga and
San Juan (Western Cordillera) formations, and in Colombia the same lithologies have been given numerous names, although here we will use
the terms Volcanic Formation for the mac rocks of the Western Cordillera and forearc (Kerr et al., 1997), and Amaime Fm. for the basalts
exposed within the Cauca-Pata Valley (Fig. 1). No geochemical differences are found between the basement of the forearc and the Western
Cordillera.
Radiometric data from the San Juan Fm. include a Sm-Nd isochron
date of 123 13 Ma (Lapierre et al., 2000), a weighted mean 40Ar/39Ar
(hornblende) date of 99.2 1.3 Ma from a U-shaped age spectrum
(Mamberti et al., 2004), and a concordant U-Pb zircon date of 87.10
1.66 Ma (Vallejo et al., 2006). The U-shaped 40Ar/39Ar age spectrum reveals the presence of excess 40Ar, and this date is discarded. The discrepancy between the Sm/Nd and U-Pb dates suggests that the densely
faulted San Juan Unit is mapped incorrectly, and it includes intercalated
but unrelated sequences. Jaillard et al. (2004) suggest that the older
Sm/Nd date is from an Early Cretaceous, allochthonous oceanic plateau,
and they name it the San Juan-Multitud Terrane. We suggest that the
foliated gabbros that yield the older Sm/Nd age are a detached fragment
of the anastomosed Peltetec Complex. Luzieux et al. (2005) report a plateau hornblende 40Ar/39Ar from the Pion Fm. of 88.0 1.6 Ma, suggesting that it is equivalent to the basement of the Western Cordillera.
Amphibolites of the Totoras Amphibolite in the central Western
Cordillera of Ecuador formed by metamorphism of an oceanic plateau
at 800850 C and 69 kbar (Jaillard et al., 2004; Beaudon et al., 2005).
Vallejo et al. (2006) report a hornblende plateau 40Ar/39Ar date of
84.69 2.23 Ma, and suggest that this records the timing of retrogression through 550500 C, at the base of the oceanic plateau. Concordant
zircon U-Pb ages (gabbros; Villagmez et al., 2011) and plateau 40Ar/39Ar
dates of basaltic groundmass (Kerr et al., 1997) suggest that the oceanic
plateau rocks of Colombia crystallised during 10092 Ma, which is
slightly older than the ages from Ecuador. The crystalline ages and geochemical compositions reveal a strong afnity with the oceanic hotspot related rocks of the Caribbean Large igneous Province (e.g. Sinton
et al., 1998), and they are considered to be detached allochthons of
that province (e.g. Kerr et al., 1997).
Field studies and geochemical data show that the oceanic plateau
rocks were intruded by a primitive arc sequence (Fig. 20). The oldest
of these arc units are acidic intrusions of the Pujil Granite (Ecuador; zircon U-Pb 85.5 1.4; Vallejo et al., 2006), the Buga Granite (Colombia;
zircon U-Pb 90.6 1.3 Ma92.1 1.3 Ma; Villagmez et al., 2011),
the Santa Fe Tonalite (Colombia; Sm-Nd age 98.0 9.1 Ma; Weber
et al., 2011), and the Altamira Gabbro (zircon U-Pb 88.9 1.5; Zapata
et al., 2011), all of which intrude the hot-spot related basement basalts.

These rocks are equivalent to the Aruba Tonalite (Island of Aruba; zircon
U-Pb ages range between 89 and 87 Ma; Wright and Wyld, 2004; Van
der Lelij et al., 2010), which intrudes Turonian basalts that erupted
above an oceanic hot-spot (White et al., 1999). The Rio Cala Group
(Fig. 20; Vallejo et al., 2009) stratigraphically (high-Mg andesites,
basalts, and turbidites) overlies the oceanic plateau within Ecuador,
and its intra-oceanic character has been determined by isotopic and
geochemical studies (Cosma et al., 1998; Mamberti, 2001; Mamberti
et al., 2003; Allibon et al., 2005; Vallejo et al., 2006), combined with
dense mineral assemblages that reveal no input from differentiated continental crust (Hughes et al., 1998; Vallejo et al., 2009). No U-Pb dates
have been obtained from the igneous rocks of the Rio Cala Group, although the basal La Portada Fm. yields Santonian (85.8983.5 Ma;
Gradstein et al., 2004) biostratigraphic ages (Kerr et al., 2002). Similarly,
Campanian radiolaria have been found intercalated within island arc
lavas of the Ricaurte Arc in southern Colombia (Western Cordillera;
Spadea et al., 1989). The association of SantonianMaastrichtian intraoceanic island-arc rocks overlying oceanic hot-spot derived rocks is documented throughout the circum-Caribbean region (e.g. Frost and Snoke,
1989; Donnelly et al., 1990).
The similar geochemistry and dates obtained from the Western
Cordilleras and the forearcs suggest that these rocks were derived
from a single terrane (e.g. Luzieux et al., 2006). We refer to the accreted
allochthons within the forearcs and Western Cordilleras of Ecuador and
Colombia as the Caribbean Large Igneous Province (CLIP), which includes the equivalent Pallatanga-Pion and Calima terranes (Figs. 1
and 20). The basaltic basement of the Choc-Panam Terrane located
north of the Garrapatas Fault (Fig. 1) is younger and accreted to South
America during the Miocene, and thus it is not included in this
discussion.
6.1.2. Time of initial accretion with South America
Estimates of the timing of accretion of the CLIP onto northwestern
South America are 8565 Ma (Aspden et al., 1987; Lebrat et al., 1987;
Kerr et al., 2002; Spikings et al., 2005), or 7565 Ma (Spikings et al.,
2001; Hughes and Pilatasig, 2002; Jaillard et al., 2004; Luzieux
et al., 2006; Vallejo et al., 2006; Spikings et al., 2010; Van der Lelij
et al., 2010; Villagmez and Spikings, 2013).
Thermochronological analyses of the Cordillera Real and the Cordillera
Central reveal the onset of extremely rapid cooling at 7573 Ma (Fig. 7b,
c; Spikings et al., 2001, 2010; Villagmez and Spikings, 2013). Cooling is
interpreted to be due to exhumation at rates of 11.6 km/My during
7565 Ma, and high cooling rates continued until ~ 55 Ma. Recent
(Cochrane et al., 2014b) and new (this study) tT paths generated from
apatite U-Pb data reveal the onset of rapid cooling at 8075 Ma
(Fig. 7a). Palaeomagnetic data from the Pion and San Lorenzo blocks
of coastal Ecuador (Fig. 1) record 4050 of clockwise rotation during
7370 Ma (Luzieux et al., 2006), which was synchronous with rapid
exhumation of the same basement rocks located closer to South
America (Spikings et al., 2005).
Redbeds of the Tena Formation, located in the Subandean zone
(Fig. 1) of Ecuador are the oldest within the foreland basin to host
signicant quantities of metamorphic mineral grains derived from
high elevations to the west (Ruiz et al., 2004). Furthermore, ssiontrack dates of detrital zircons within the Tena Fm. are indistinguishable from their depositional ages, revealing extremely high exhumation rates in the Cordillera Real. At the same time, there is a reduction
in the detrital supply from cratonic South America to the east. Within
the forearc of Ecuador, the Late CampanianMaastrichtian (70
65 Ma; Gradstein et al., 2004) Yunguilla Fm. was being deposited in
a basin oored by oceanic crust of either the Pallatanga Unit
(Hughes and Pilatasig, 2002), or the San Juan-Multitud Terrane
(Jaillard et al., 2004). The turbidites were sourced from metamorphic
rocks of the Eastern Cordillera (Vallejo et al., 2009). Jaillard et al.
(2004) described coeval quartz-free pelagic cherts, and quartz rich
turbidites of the Yunguilla Fm. in the Western Cordillera of

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

Ecuador. Both of these sequences overlie oceanic mac rocks, and


Jaillard et al. (2004) suggest that they dene two separate terranes
that accreted before 71 Ma, and during 6965 Ma. Within
Colombia, rapid cooling and exhumation during 7565 Ma was synchronous with the deposition of the siliciclastic, El Cobre,
Monserrate, La Tabla and Cimarrona Fms. during the Campanian
Maastrichtian in the retro-foreland Magdalena Valley (Villamil,
1999), and the Nogales Fm. in the forearc, which hosts a large proportion of metamorphic detritus derived from the Cordillera Central
(Moreno and Pardo, 2003).
These data were used by Luzieux et al. (2006), Vallejo et al.
(2006), Spikings et al. (2010) and Villagomez et al. (2013) to suggest
that mac rocks that originated above an oceanic hot-spot at 99
87 Ma formed a single terrane, which along with its overlying arc accreted against the margin of northwestern South America at ~ 75 Ma
(Fig. 20). This interpretation implies that the intra-oceanic arc (E.g.
Rio Cala Group) formed by west-dipping subduction beneath the
buoyant hot-spot derived rocks, prior to its collision with South
America, which is consistent with the models of Burke (1988), Kerr
et al. (1997), Spikings et al. (2001), Van der Lelij et al. (2010), and
Zapata et al. (2011). However, this is inconsistent with the reconstruction of Pindell and Kennan (2009), who draw an east dipping
subduction zone at 84 Ma beneath South America. The reconstructions of Pindell and Kennan (2009) suggest that the Rio Cala Arc is
a continental arc, which is inconsistent with geochemical, isotopic
and sedimentological data. Estimates of collision at ~ 85 Ma (e.g.
Kerr et al., 2002; Spikings et al., 2005) were based around a misinterpretation of thermochronological data.

6.1.3. The nature of the CLIPSouth America suture


Within Ecuador, the suture between the 90 and 87 Ma CLIP and
the pre-existing continental margin is mainly obscured beneath
volcanic rocks and intermontane basins within the Interandean Depression (Fig. 1). The suture may be represented by the Pujil Fault
zone, which exposes the serpentinized Pujil Melange, whose matrix
is the Pallatanga Unit. Two gabbroic inliers (E.g. Guayllabamba
Basin) occur within the Interandean Depression, although the ages
of these rocks are unknown. Foliated metamorphic crust erupts as
xenoliths within volcanoes along the western side of the Interandean Depression (Bruet, 1987), suggesting that the basement to
the valley may be a complex assemblage of dissected Early Cretaceous margin, and accreted CLIP. The Arqua Complex in Colombia
is faulted against the Amaime Fm., which oors the Cauca-Pata
Valley (Fig. 1) across the Cauca-Almaguer Fault. According to our
model (Fig. 19), the Cauca-Almaguer Fault is equivalent to the
Ingapirca Fault (Litherland et al., 1994; Fig. 1) in Ecuador (the western margin of the Guamote sequence). No reliable age dates have
been obtained from the basalts of the Amaime Fm., although their
chemical composition suggests that they formed above an oceanic
hot-spot (Kerr et al., 1997), and are identical to the Volcanic,
Pallatanga and Pion fms. The nature of the contact between the
Buga Batholith (90.6 1.3 Ma92.1 1.3 Ma; Villagmez et al.,
2011) and the Amaime Fm. is uncertain, although it may be intrusive,
in which case the Amaime Fm. is older than ~ 91 Ma. We suggest that
these rocks form part of the CLIP, and the suture in Colombia is the
Cauca-Almaguer Fault.
Spikings and Crowhurst (2004) and Winkler et al. (2005) show that
the Interandean Valley formed in Ecuador at ~ 6 Ma within a
dextral transcurrent setting (e.g. Winter and Laven, 1989), implying that the accreted CLIP was closer to the Early Cretaceous
palaeomargin before ~ 6 Ma. We suggest that the CLIP-South
America suture, which started forming at ~ 75 Ma, now exists as
a complex melange zone that forms the basement of the Interandean Depression, and is perhaps more intact as the CaucaAlmaguer Fault within Colombia.

39

7. Conclusions
1. Geochemical and isotopic analyses suggest that high-temperature
metamorphism within Central American terranes (e.g. Maya
Block) at 250 Ma occurred during compression driven by terrane
accretion (e.g. Weber et al., 2007) along central Western Pangaea.
Extension prevailed along the Peruvian margin at 250 Ma, and
thus it is likely that no continental terranes lay outboard of Peru at
that time.
2. Magmatic underplating and anatexis of continental crust during
240225 Ma occurred during progressive thinning of the continental lithosphere during rifting along western Pangaea. Rifting advanced to complete separation of continental crust by ~ 216 Ma,
and the formation of oceanic lithosphere between the conjugate
margins of northwestern South America and basement terranes of
Central America (e.g. Oaxaquia). The rifting event is recorded by
amphibolitised tholeiitic basaltic dykes and extensive tracts of
migmatites and S-type granites within the conjugate margins. The
rift axis propagated southwards, and extension is recorded along
western Peru (the Mitu Aulocagen), Bolivia, western Argentina,
Chile and southern Brazil. Rifting along northwestern South
America started as a back-arc basin to a Permian arc, and represents
the early break-up of western Pangaea, leading to the separation of
North and South America by ~180 Ma.
3. Metaluminous, I-type arc magmatism commenced in northwestern
South America at ~209 Ma, due to east-dipping subduction of the
Farallon Plate. The arc axis migrated oceanward at some time during 194189 Ma, formed a long/lived continental arc during 189
144 Ma and the arc axis may have migrated ~ 100 km oceanward
during this time at a very slow rate. Coeval arc magmatism along
the Peruvian margin (~ 216135 Ma) also started to migrate
oceanward at ~175 Ma, resulting in coeval arc and back-arc rocks.
Arc migration is considered to be a result of slab-retreat along the
western margin of South America, which caused the continental
margin to extend, thinning the continental crust and generating
progressively more isotopically juvenile arcs. The onset of latest
TriassicJurassic subduction beneath Colombia and Ecuador may
young towards the south, although this trend cannot be extended
towards southern Peru, across the Huancabamba Deection.
4. The present-day gap in Jurassic arc rocks north of the Arequipa
Terrane in Peru is considered to be a result of tectonic erosion
(Clift et al., 2003), and thus we propose that an arc did form in
that region. We suggest that the alternative hypothesis that the
arc was tectonically displaced northwards and now forms part of
Colombia (Bayona et al., 2010) is unlikely because the Jurassic arc
in Colombia and Ecuador is not temporally duplicated.
5. Trench retreat of the east dipping-subduction zone accelerated
along northwestern South America at ~ 144 Ma, and extension
during 144115 Ma formed syn-tectonic granitoid intrusions within Ecuador, attenuated the continental margin forming thin intraarc basins characterised by transitional crust, and resulted in an
oceanward migration of the arc axes, which became progressively
more isotopically juvenile and geochemically depleted. Arc rocks
of the Quebradagrande Complex and Alao arc erupted through
thin continental crust during the Early Cretaceous within a marine
environment. Rapid extension may have rifted some narrow continental slivers (e.g. the Guamote Sequence) from the margin.
Back-arc magmatism is sporadically preserved within the Eastern
Cordillera of Colombia (136121 Ma). Highly oblique and sinistral
convergence directions between the Peruvian margin and the
Farallon Plate lead to a magmatic gap in Peru during ~135115 Ma.
6. The distribution and composition of sedimentary rocks, combined
with detrital thermochronology suggests that the margin of northwestern South America was placed under compression at ~115 Ma.
Compression of the attenuated, weak, hot crust juxtaposed arc
rocks with transitional crust, forming a proto-cordillera, which

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

40

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

80

60

70

50

40

80

70

60

50

40

Caribbean
Plate

10

Rio Cala
Arc

Mainly amagmatic
within South America

-cord

10

illera

Oceanic
plateau

Ar

Oceanic
plateau

ProtoCaribbean

CA/NA

CA/NA

Proto

CA/SA
T

10

CA/SA

South America

10

South America

90

80

70

95-85 Ma
20

60

50

20
Caribbean
Plate

20

Late Cretaceous arcs


A: Antioquia Batholith, Ar: Aruba Batholith,
B: Buga Granodiorite, P: Pujil Granite,
T: Tangula Batholith
Early Cretaceous subduction zone

CA/NA
Ar

10

M-HP/LT rocks (Arqua, Barragn, Raspas complexes)


Peltetec Complex
Early Cretaceous arcs

CA/SA

Northeast facing

PP

Oceanic
plateau

P
T

West facing, continental


(Quebradagrande Complex, Alao Arc)

South America

tal
as lith
co tho
ba

75-70 Ma

tal
as lith
c o tho
ba

tal
as lith
c o tho
ba

100 Ma

10

Fig. 20. Plate reconstruction for northwestern South America from after the initial formation of oceanic plateau rocks of the Caribbean Large Igneous Province at 100 Ma until the collision
of the plateau rocks and their overlying arc with South America starting at 75 Ma, modied after Villagmez et al. (2011). Relative positions of South and North America and arcs along
the northern margin of the plateau are from Pindell and Kennan (2009). The plates are positioned relative to the Indo-Atlantic hot-spot reference frame (Mller et al., 1993). The
reconstruction draws the east-facing Rio Cala Arc forming above the oceanic plateau as it approaches South America. PPC: Pion-Pallatanga-Calima terrane. Relative convergence direction
CA/NA Caribbean Plate/North America, CA/SA: Caribbean Plate/South America.

supplied detritus towards the fore- and backarc. M-HP/LT rocks are
faulted against the western margin of these compressed sequences,
and represent a subduction channel that started to exhume from
peak eclogitic conditions at 130126 Ma. These eclogites and
blueschists retrogressed through ~400 C at 120112 Ma, and it is
likely that they were obducted onto the margin during compression
that started at ~115 Ma. These rocks originally formed parts of the
same slab, and varying trench-parallel metamorphic facies reect
exhumation from varying depths. Models which invoke westdipping slabs after 125 Ma do not account for the spatial juxtaposition of M-HP/LT rocks and their associated arcs.
7. We suggest that there is very little evidence for the existence of
large allochthonous continental terranes (Taham and Chaucha)
outboard of northwestern South America during the Jurassic
Early Cretaceous. These suspect terranes are not required to t the
data presented here, and the geochemical, isotopic, sedimentological and thermochronological data can be accounted for by having a
single east-dipping slab that is retreating oceanward at variable
rates during 189115 Ma.
8. Highly oblique dextral convergence angles between the margin of
northwestern South America and the Caribbean Plate (Kennan

and Pindell, 2009) resulted in only very minor intra-plate


magmatism (Barragn et al., 2005) during 115100 Ma, which
may have occurred during the steepening of an antiquated slab.
Net convergence with an orthogonal component along the
Peruvian margin started forming the Coastal Batholith at ~115 Ma
within a marginal basin that formed during extension.
9. Oceanic plateau rocks of the Caribbean Large Igneous Province
erupted through the Farallon Plate at near equatorial latitudes
(Luzieux et al., 2006) during 10087 Ma, and these rocks migrated
approximately eastwards relative to South America. Subduction
beneath northwestern South America during this period was
restricted to northern Colombia and southern Ecuador, distal from
the leading edge of the approaching plateau. However, a majority
of the margin was amagmatic, and the intervening oceanic lithosphere was consumed by west-dipping subduction beneath the
oceanic plateau, forming the Rio Cala intra-oceanic arc, and numerous scattered intrusions. Collectively, the plateau and overlying arc
are referred to as the Caribbean Large Igneous Province.
10. The Caribbean Large Igneous Province rst collided with South
America at ~ 75 Ma, resulting in the detachment and clockwise
rotation of allochthons that form the present-day basement of the

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

forearc plains and the Western Cordillera. Other detached


allochthons form the basement to the Leeward Antilles along the
southern Caribbean plate boundary. Allochthons with South
America are sutured against the Early Cretaceous margin via the
Cauca-Almaguer (Colombia) and Ingapirca (Ecuador) faults. Within
Ecuador, the suture is represented by a melange, which is mainly
buried beneath the Interandean Depression.

Acknowledgements
We thank Arturo Egez, Etienne Jaillard, Alfredo Buitron, Byron
Pelicita and Luis Lopez for assistance in the eld in the cordilleras of
Ecuador, and Ecopetrol S.A., Andres Mora, Andreas Kammer, Agustin
Cardona, Jaime Corredor, Jaime Castellanos, Wilson Casallas, and
Luis Quiroz for their assistance during eld work in Colombia. The
manuscript was improved by the thorough and helpful reviews of Victor
Ramos, Maria Helbig and an anonymous reviewer.
Appendix A. Supplementary data
Supplementary data to this article can be found online at http://dx.
doi.org/10.1016/j.gr.2014.06.004.
References
Allibon, J., Monjoie, P., Lapierre, H., Jaillard, E., Bussy, F., Bosch, D., 2005. High Mgbasalts in the Western Cordillera of Ecuador: evidence of plateau root melting
during Late Cretaceous arc magmatism, In: Sempr, T. (Ed.), Proceedings of
the Sixth International Symposium on Andean Geodynamics, Program and Abstracts, IRD ditionsUniversitat de Barcelona, Instituto Geolgico y Minero de
Espaa, Barcelona, Spain, p. 3335.
Alvarez, J.A., 1983. Geologia de la Cordillera Central y e1 occidente colombiano y
petroquimica de 10s intrusivos granitoides Mesocenozoicos. Boletin Geological
INGEOMINAS, Bogota, Colombia, p. 26.
Arculus, R.J., Lapierre, H., Jaillard, E., 1999. Geochemical window into subduction and accretion processes: Raspas metamorphic complex, Ecuador. Geology 27, 547550.
Aspden, J.A., Litherland, M., 1992. The geology and Mesozoic collisional history of the Cordillera Real, Ecuador. Tectonophysics 205, 187204.
Aspden, J.A., McCourt, W.J., Brook, M., 1987. Geometrical control of subduction-related
magmatism: the Mesozoic and Cenozoic plutonic history of Western Colombia. Journal of the Geological Society 144, 893905.
Aspden, J.A., Bonilla, W., Duque, P., 1995. The El Oro metamorphic complex, Ecuador: geology and economic mineral deposits. Nottingham, British Geological Survey, Overseas Geology and Mineral Resources, 67, p. 63.
Atherton, M.P., 1990. The Coastal Batholith of Peru: the product of rapid recycling of new
crust formed within rifted continental margin. Geological Journal 25, 337349.
Atherton, M.P., Petford, N., 1996. Plutonism and the growth of Andean Crust at 9S from
100 to 3 Ma. Journal of South American Earth Sciences 9, 19.
Bahlburg, H., Vervoort, J.D., Du Frane, S.A., Bock, B., Augustsson, C., Reimann, C., 2009.
Timing of crust formation and recycling in accretionary orogens: insights learned
from the western margin of South America. Earth-Science Reviews 97, 215241.
Baldock, M.W., 1982. Geology of Ecuador. Explanatory Bulletin of the national geological
map of the Republic of Ecuador, scale 1: 1,000,000 Direccion General de Geologia y
Minas, Quito, Ecuador, 11, p. 54.
Balkwill, H.R., Paredes, F.I., Rodriguez, G., Almeida, J.P., 1995. Northern part of Oriente
Basin, Ecuador: reection seismic expression of structures. In: Tankard, A.J., Suarez,
R., Welsink, H.J. (Eds.), Petroleum Basins of South America: American Association of
Petroleum Geologists, 62, pp. 559571.
Barragn, R., Baby, P., Duncan, R., 2005. Cretaceous alkaline intra-plate magmatism in the
Ecuadorian Oriente Basin: geochemical, geochronological and tectonic evidence.
Earth and Planetary Science Letters 236, 670690.
Barredo, S., Chemale, F., Marsicano, C., vila, J.N., Ottone, E.G., Ramos, V.A., 2012. Tectonosequence stratigraphy and U-Pb zircon ages of the Rincn Blanco Depocenter, northern Cuyo Rift, Argentina. Gondwana Research 21, 624636.
Bayona, G., Jimenez, G., Silva, C., Cardona, A., Montes, C., Roncancio, J., Cordani, U., 2010.
Paleomagnetic data and K-Ar ages from Mesozoic units of the Santa Marta massif: a
preliminary interpretation for block rotation and translations. Journal of South
American Earth Sciences 29, 817831.
Beaudon, E., Martelat, J.E., Amortegui, A., Lapierre, H., Jaillard, E., 2005. Metabasites de la
cordillere occidentale d'Equateur, temoins du soubassement oceanique des Andes
d'Equateur. Comptes Rendus Geoscience 337, 625634.
Beutel, E.K., 2009. Magmatic rifting of Pangaea linked to onset of South America plate motion. Tectonophysics 468, 149157.
Boekhout, F., Spikings, R., Sempere, T., Chiaradia, M., Ulianov, A., Schaltegger, U., 2012. Mesozoic arc magmatism along the southern Peruvian margin during Gondwana breakup and dispersal. Lithos 146-147, 4864.

41

Boekhout, F., Sempere, T., Spikings, R., Schaltegger, U., 2013a. Late Paleozoic to Jurassic
chronostratigraphy of coastal southern Peru: temporal evolution of sedimentation
along an active margin. Journal of South American Earth Sciences 47, 179200.
Boekhout, F., Roberts, N.M.W., Gerdes, A., Schaltegger, U., 2013b. A Hf-isotope perspective
on continent formation in the south Peruvian Andes. The Geological Society of
London Special Publications 389. http://dx.doi.org/10.1144/SP389.6.
Bosch, D., Gabriele, P., Lapierre, H., Malfere, J.-L., Jaillard, E., 2002. Geodynamic signicance
of the Raspas Metamorphic Complex (SW Ecuador): geochemical and isotopic constraints. Tectonophysics 345, 83102.
Botero, A., 1963. Contribucin al conocimiento de la geologa de la zona central de
Antoquia. Anales Facultad de Minas (Medelln) 57, 101.
Bourgois, J., Toussaint, J.-F., Gonzales, H., Azema, J., Calle, B., Desmet, A., Murcia, L.A.,
Acevedo, A.P., Parra, E., Tournon, J., 1987. Geological history of the Cretaceous
ophiolitic complexes of Northwestern South America (Colombian Andes).
Tectonophysics 143, 307327.
Bozkurt, E., Park, L.R.G., 1994. Southern Menderes Massif: an incipient metamorphic core
complex in western Anatolia, Turkey. Journal of the Geological Society 151, 213216.
Bristow, C.R., 1973. Guide to the geology of the Cuenca Basin, southern Ecuador.
Ecuadorian Geological and Geophysical Society, Quito.
Burke, K., 1988. Tectonic evolution of the Caribbean. Annual Review Earth and Planetary
Sciences 16, 201230.
Bustamante, C., Cardona, A., Bayona, G., Mora, A., Valencia, V., Gehrels, G., Vervoort, J.,
2010. U-Pb LA-ICP-MS geochronology and regional correlation of Middle Jurassic intrusive rocks from the Garzn Massif, Upper Magdalena Valley and Central Cordillera,
southern Colombia. Boletin de Geologa 32, 93109.
Bustamante, A., Juliani, C., Hall, C.M., Essene, E.J., 2011. 40Ar/39Ar ages from blueschists of
the Jambal region, Central Cordillera of Colombia: implications on the styles of accretion in the Northern Andes. Geologica Acta 9, 351362.
Bustamante, A., Juliani, C., Essene, E.J., Hall, C.M., Hyppolito, T., 2012. Geochemical constraints on blueschist-facies rocks of the Central Cordillera of Colombia: the Andean
Barragn region. International Geology Review 54, 10131030.
Cameron, K.L., Lopez, R., Ortega-Gutirrez, F., Solari, L.A., Keppie, J.D., Schulze, C., 2004. UPb geochronology and Pb isotopic compositions of leached feldspars: constraints on
the origin and evolution of Grenville rocks from eastern and southern Mexico. Geological Society of America Memoirs 197, 755769.
Cardona, A., Cordani, U., Sanchez, A., 2007. Metamorphic, geochronological and constraints from the pre-Permian basement of the eastern Peruvian (10S): a Paleozoic
extensionalaccretionary orogen? 20th Colloquium on Latin American Earth Sciences, April 1113, Kiel, Germany, pp. 2930.
Cardona, A., Valencia, V., Garzn, A., Montes, C., Ojeda, G., Ruiz, J., Weber, M., 2010. Permian
to Triassic I to S-type magmatic switch in the northeast Sierra Nevada de Santa Marta
and adjacent regions, Colombian Caribbean: tectonic setting and implications within
Pangea paleogeography. Journal of South American Earth Sciences 29, 772783.
Castro, A., Moreno-Ventas, I., Fernndez, C., Vujovich, G., Gallastegui, G., Heredia, N.,
Martino, R.D., Becchio, R., Corretg, L.G., Daz-Alvarado, J., Garca-Arias, M., Liu, D.-Y.
, 2011. Petrology and SHRIMP U-Pb zircon geochronology of Cordilleran granitoids
of the Bariloche area, Argentina. Journal of South American Earth Sciences 32,
508530.
Cawood, P.A., McCausland, P.J.A., Dunning, G.R., 2001. Opening Iapetus: constraints from
the Laurentian margin in Newfoundland. Geological Society of America Bulletin
113, 443453.
Centeno-Garcia, E., Keppie, J.D., 1999. Latest Paleozoic-early Mesozoic structures in the
central Oaxaca Terrane of southern Mexico: deformation near a triple junction.
Tectonophysics 301, 231242.
Chamberlain, K.R., Schmitt, A.K., Swapp, S.M., Harrison, T.M., Swoboda-Colberg, N.,
Bleeker, W., Peterson, T.D., Jeffferson, C.W., Khudoley, A.K., 2010. In situ U-Pb SIMS
(IN-SIMS) micro-baddeleyite dating of mac rocks: method with examples. Precambrian Research 183, 379387.
Chappell, B.W., White, A.J.R., 1974. Two contrasting granite types. Pacic Geology 8,
173174.
Cherniak, D.J., Lanford, W.A., Ryerson, F.J., 1991. Lead diffusion in apatite and zircon using
ion implantation and Rutherford Backscattering techniques. Geochimica et
Cosmochimica Acta 55, 16631673.
Chew, D.M., Schaltegger, U., Koler, J., Whitehouse, M.J., Gutjahr, M., Spikings, R.A.,
Mikovc, A., 2007. U-Pb geochronologic evidence for the evolution of the
Gondwanan margin of the north-central Andes. Geological Society of America Bulletin 119, 697711.
Chew, D.M., Magna, T., Kirkland, C.L., Miskovic, A., Cardona, A., Spikings, R., Schaltegger, U.,
2008. Detrital zircon ngerprint of the Proto-Andes: evidence for a Neoproterozoic
active margin? Precambrian Research 167, 186200.
Chiaradia, M., Vallance, J., Fontbot, L., Stein, H., Schaltegger, U., Coder, J., Richards, J.,
Villeneuve, M., Gendall, I., 2009. U-Pb, Re-Os and 40Ar/39Ar geochronology of the
Nambija Au-skarn and Pangui porphyry Cu deposits, Ecuador: implications for the Jurassic metallogenic belt of the northern Andes. Mineralium Deposita 44, 371387.
Clift, P.D., Hartley, A.J., 2007. Slow rates of subduction erosion and coastal underplating
along the Andean margin of Chile and Peru. Geology 35, 503506.
Clift, P.D., Pecher, I., Kukowski, N., Hampel, A., 2003. Tectonic erosion of the Peruvian
forearc, Lima Basin, by subduction and Nazca Ridge collision. Tectonics 22. http://
dx.doi.org/10.1029/2002TC001386.
Cochrane, R., 2013. U-Pb thermochronology, geochronology and geochemistry of NW
South America: rift to drift transition, active margin dynamics and implications for
the volume balance of continents(PhD thesis) Terre & Environment, 118. University
of Geneva, Switzerland, p. 191.
Cochrane, R., Spikings, R., Gerdes, A., Ulianov, A., Mora, A., Villagmez, D., Putlitz, B.,
Chiaradia, M., 2014a. Permo-Triassic anatexis, continental rifting and the disassembly
of western Pangaea. Lithos 190191, 383402.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

42

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

Cochrane, R., Spikings, R.A., Chew, D., Wotzlaw, J.-F., Chiaradia, M., Tyrell, S., Schaltegger,
U., Van der Lelij, R., 2014b. High temperature (N350 C) thermochronology and
mechanisms of Pb loss in apatite. Geochimica et Cosmochimica Acta 127, 3956.
Cocks, L.R.M., Torsvik, T.H., 2002. Earth geography from 500 to 400 million years ago: a
faunal and palaeomagnetic review. Journal of the Geological Society 159, 631644.
Collins, W.J., 2002. Hot orogens, tectonic switching, and creation of continental crust. Geology 30, 535538.
Collins, W.J., Richards, S.W., 2008. Geodynamic signicance of S-type granites in circumPacic orogens. Geology 36, 559562.
Collins, W.J., Belousova, E.A., Kemp, A.I.S., Murphy, J.B., 2011. Two contrasting Phanerozoic
orogenic systems revealed by hafnium isotope data. Nature Geoscience 4, 333337.
Colony, R.J., Sinclair, J.H., 1932. Metamorphic and igneous rocks of eastern Ecuador. Annals of the New York Academy of Sciences 34, 153.
Cooper, B., et al., 1995. Basin development and tectonic history of the Eastern Cordillera
and Llanos Basin. Colombia. American Association of Petroleum Geologists 79,
14211443.
Cordani, U.G., Brito-Neves, B.B., D'Agrella, M.S., 2003. From Rodinia to Gondwana: a review of the available evidence from South America. Gondwana Research 6, 275283.
Cosma, L., Lapierre, H., Jaillard, E., Laubacher, G., Bosch, D., Desmet, A., Mamberti, M.,
Gabriele, P., 1998. Petrographie et geochimie des unites magmatiques de la Cordillre
Occidentale d'Equateur (030S): implications tectoniques. Bulletin de la Societe
Geologique de France 169, 739751.
Dalmayrac, B., Laubacher, G., Marocco, R., 1980. Gologie des Andes pruviennes.
Caractres gnraux de l'volution gologique des Andes pruviennes: Travaux et
Documents de l'ORSTOM, 122, p. 501.
Dasch, L., 1982. U-Pb Geochronology of the Sierra de Perij, Venezuela(PhD thesis) Case
Western Reserve University (183 pp.).
De Haller, A., Corfu, F., Fontbot, L., Schaltegger, U., Barra, F., Chiaradia, M., Frank, M.,
Alvarado, J.Z., 2006. Geology, Geochronology, and Hf and Pb Isotope data of the
Ral-Condestable Iron Oxide-Copper-Gold Deposit, Central Coast of Peru. Economic
Geology 101, 281310.
Dehler, S.A., 2012. Initial rifting and breakup between Nova Scotia and Morocco: insight
from new magnetic models. Canadian Journal of Earth Sciences 49, 13851394.
Demouy, S., Paquette, J.-L., Blanquat, M., de, S., Benoit, M., Belousova, E.A., O'Reilly,, S.Y.,
Garca, F., Tejada, L.C., Gallegos, R., Sempere, T., 2012. Spatial and temporal evolution
of Liassic to Paleocene arc activity in southern Peru unravelled by zircon U-Pb and Hf
in-situ data on plutonic rocks. Lithos 155, 183200.
Dickinson, W.R., Lawton, T.F., 2001. Carboniferous to Cretaceous assembly and fragmentation of Mexico. Geological Society of America Bulletin 113, 11421160.
Donnelly, T., Beets, D., Carr, M., Jackson, T., Klaver, G., Lewis, J., Maury, R., Schellekens, H.,
Smith, A., Wadge, G., Westercamp, D., 1990. History and tectonic setting of the Caribbean magmatism. In: Dengo, G., Case, J. (Eds.), The Caribbean Region: Boulder, Colorado, Geological Society of America, Geology of North America H, pp. 339374.
Ducea, M.N., Gehrels, G.E., Shoemaker, S., Ruiz, J., Valencia, V.A., 2004. Geologic evolution
of the Xolapa Complex, southern Mexico: evidence from U-Pb zircon geochronology.
Geological Society of America Bulletin 116, 10161025.
Eagles, G., 2007. New angles on South Atlantic opening. Geophysical Journal International
168, 353361.
Elas-Herrera, M., Ortega-Gutirrez, F., 2002. Caltepec fault zone: an Early Permian dextral
transpressional boundary between the Proterozoic Oaxacan and Paleozoic Acatln
complexes, southern Mexico, and regional implications. Tectonics 21. http://dx.doi.
org/10.1029/200TC001278.
Feininger, T., 1980. Eclogite and related high-pressure regional metamorphic rocks from
the Andes of Ecuador. Journal of Petrology 21, 107140.
Feininger, T., Bristow, C.R., 1980. Cretaceous and Paleogene geologic history of coastal
Ecuador. Geologische Rundschau 69, 849874.
Feininger, T., Silberman, M.L., 1982. K-Ar geochronology of basement rocks on the northern anks of the Huancabamba Deection, Ecuador. Open File Report. United States
Geological Survey, 82, p. 206.
Feininger, T., Barrero, D., Castro, N., 1972. Geologa de Antioquia et CaldasSub-zona II-B.
Boletin Geologa Bogota 20, 173.
Fortey, N.J., 1990. Petrographic data and course notes for the Cordillera Real Project,
Ecuador. British Geological Survey Technical Report WG/90/14/R (67 pp.).
Frost, C., Snoke, A., 1989. Tobago, West Indies, a fragment of a Mesozoic oceanic island arc:
petrochemical evidence. Journal of the Geological Society of London 146, 953964.
Frost, B.R., Barnes, C.G., Collins, W.J., Arculus, R.J., Ellis, D.J., Frost, C.D., 2001. A geochemical
classication of granitic rocks. Journal of Petrology 42, 20332048.
Funck, T., Jackson, H.R., Louden, K.E., Dehler, S.A., Wu, Y., 2004. Crustal structure of the
northern Nova Scotia rifted margin (eastern Canada). Journal of Geophysical Research Solid Earth 109. http://dx.doi.org/10.1029/2004JB003008.
Gabriele, P., 2002. HP terranes exhumation in an active margin setting: geology, petrology
and geochemistry of the Raspas Complex in SW Ecuador(PhD thesis) Universit de
Lausanne.
Gerbi, C.C., Johnson, S.E., Koons, P.O., 2006. Controls on low-pressure anatexis. Journal of
Metamorphic Petrology 24, 107118.
Gerdes, A., Zeh, A., 2009. Zircon formation versus zircon alteration new insights from
combined U-Pb and Lu-Hf in-situ LA-ICP-MS analyses, and consequences for the interpretation of Archean zircon from the Central Zone of the Limpopo Belt. Chemical
Geology 261, 230243.
Gerdes, A.G., Montero, P.M., Bea, F.B., Fershater, G.F., Borodina, N.B., Osipova, T.O.,
Shardakova, G.S., 2002. Peraluminous granites frequently with mantle-like isotope
compositions: the continental-type Murzinka and Dzhabyk batholiths of the eastern
Urals. International Journal of Earth Sciences 91, 319.
Gerya, T.V., Stckhert, B., Perchuk, A.L., 2002. Exhumation of high-pressure metamorphic rocks in a subduction channel: a numerical simulation. Tectonics 21,
1056.

Goldsmith, R., Marvin, R.F., Mehnert, H.H., 1971. Radiometric ages in the Santander Massif, Eastern Colombia, Colombian Andes. Professional paper United States Geological
Survey 750-D, pp. 4490.
Golonka, J., Bocharova, N.Y., 2000. Hot spot activity and the break-up of Pangea.
Palaeogeography Palaeoclimatology Palaeoecology 161, 4969.
Gmez, J., Nivia, A., Montes, N.E., Jimenez, D.M., Tejada, M.L., Sepulveda, J., Osorio, J.A.,
Gaona, T., Diederix, H., Uribe, H., Mora, M., 2007. Geological map of Colombia. Escala
1:1'000.000. Ingeominas, 2nd Edition, Bogota.
Gmez-Cruz, A.D.J., Moreno-Snchez, M., Pardo, A., 1995. Edad y origen del Complejo
metasedimentario Aranzazu-Manizales en los alrededores de Manizales
(Departamento de Caldas, Colombia). Geologa Colombiana 19, 8393.
Gonzlez, H., 1980. Geologa de las Planchas 167 (Sonson) e 187 (Salamina). Boletn
Geolgico de Ingeominas, Informe 1760.
Gradstein, F.M., Ogg, J.G., Smith, A.G., et al., 2004. A geological timescale 2004. Cambridge
University Press, Cambridge.
Grajales-Nishimura, J.M., Centeno-Garcia, E., Keppie, J.D., Dostal, J., 1999. Geochemistry of
Paleozoic basalts from the Juchatengo complex of southern Mexico: tectonic implications. Journal of South American Earth Sciences 12, 537544.
Hampel, A., 2002. The migration history of the Nazca Ridge along the Peruvian active margin: a re-evaluation. Earth and Planetary Science Letters 203, 665679.
Harris, C., Faure, K., Diamond, R.E., Scheepers, R., 1997. Oxygen and hydrogen isotope 966
geochemistry of S- and I-type granitoids: the Cape Granite suite, South Africa. Chemical Geology 143, 95114.
Harrison, T.M., Clrier, J., Aikman, A.B., Hermann, J., Heizler, M.T., 2009. Diffusion of 40Ar
in muscovite. Geochimica et Cosmochimica Acta 73, 10391051.
Hartmann, L.A., Santos, J.O.S., 2004. Predominance og high Th/U, magmatic zircon in
Brazilian Shield sandstones. Geology 32, 7376.
Hastie, A.R., Kerr, A.C., 2010. Mantle plume or slab window? Physical and geochemical
constraints on the origin of the Caribbean oceanic plateau. Earth-Science Reviews
98, 283293.
Hastie, A.R., Kerr, A.C., Pearce, J.A., Mitchell, S.F., 2007. Classication of altered volcanic island arc rocks using immobile trace elements: development of the Th-Co discrimination diagram. Journal of Petrology 48, 23412357.
Helbig, M., Keppie, J.D., Murphy, J.B., Solari, L.A., 2012. U-Pb geochronological constraints
on the TriassicJurassic Ayu Complex, southern Mexico: derivation from the western
margin of Pangea-A. Gondwana Research 22, 910927.
Herbert, H.J., Pichler, H., 1983. K-Ar ages of rocks from the Eastern Cordillera of Ecuador.
Deutschen Geologischen Gesellschaft 134, 483493.
Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M., Reyes-Harker, A., Stockli, D.F., 2010.
Linking sedimentation in the northern Andes to basement conguration, Mesozoic
extension, and Cenozoic shortening: evidence from detrital zircon U-Pb ages in the
Eastern Cordillera of Colombia. Geological Society of America Bulletin 122,
14231442.
Hughes, R.A., Pilatasig, L.F., 2002. Cretaceous and Tertiary terrane accretion in the Cordillera Occidental of the Ecuadorian Andes. Tectonophysics 345, 2948.
Hughes, R., Bermdez, R., Espinel, G., 1998. Mapa Geologico de la Cordillera Occidental del
Ecuador entre 01 S: Quito, Ecuador. Corporacin de Desarrollo e Investigacin
Geolgica, Minera y MetalrgicaMinisterio de Energia EcuadorBritish Geological
Survey, scale 1:200,000.
Jaillard, E., Soler, P., 1996. Cretaceous to early Paleogene tectonic evolution of the
northern Central Andes (01S) and its relations to geodynamics. Tectonophysics
259, 4153.
Jaillard, E., Soler, P., Carlier, G., Mourier, T., 1990. Geodynamic evolution of the northern
and central Andes during early to middle Mesozoic times: a Tethyan model. Journal
of the Geological Society 147, 10091022.
Jaillard, E., Laubacher, G., Bengston, P., Dhondt, A.V., Bulot, L.G., 1999. Stratigraphy and
evolution of the Cretaceous forearc Celica-Lancones basin of southwestern Ecuador.
Journal of South American Earth Sciences 12, 5168.
Jaillard, E., Ordoez, M., Surez, J., Toro, J., Iza, D., Lugo, W., 2004. Stratigraphy of the late
CretaceousPaleogene deposits of the cordillera occidental of central Ecuador:
geodynamic implications. Journal of South American Earth Sciences 17, 4958.
Jaimes, E., de Freitas, M., 2006. An AlbianCenomanian unconformity in the northern
Andes. Evidence and tectonic signicance. Journal of South American Earth Sciences
21, 466492.
John, T., Scherer, E.E., Schenk, V., Herms, P., Halama, R., Garbe-Schnberg, D., 2010.
Subducted seamounts in an eclogite-facies ophiolite sequence: the Andean
Raspas Complex, SW Ecuador. Contributions to Mineralogy and Petrology 159,
265284.
Jolly, W.T., Lidiak, E.G., Dickin, A.Ap., Wu, T.W., 2001. Secular geochemistry of central
Puerto Rican island arc lavas: constraints on Mesozoic tectonism in the eastern Greater Antilles. Journal of Petrology 42, 21972214.
Jones, P.R., 1981. Crustal structure of the Peru continental margin and adjacent Nazca
plate, 9S. In: Kulm, L.D., Dymond, E., Dasch, J., Hussong, D.M. (Eds.), Nazca Plate:
Crustal Formation and Andean Convergence. Memoir Geological Society of America,
154, pp. 423444.
Kennan, L., Pindell, J., 2009. Dextral shear, terrane accretion and basin formation in the
Northern Andes: best explained by interaction with a Pacic-derived Caribbean
Plate? Geological Society of London Special Publications 328, 487531.
Keppie, J.D., Gutierrez, F.O., 1999. Middle American Precambrian basement: a missing
piece of the reconstructed 1-Ga orogen. In: Ramos, V.A., Keppie, J.D. (Eds.),
Laurentia-Gondwana connections before Pangaea, Boulder, Colorado. Geological Society of America Special Paper, 336.
Keppie, J.D., Dostal, J., Miller, B.V., Ortega-Rivera, A., Roldan-Quintana, J., Lee, J.W.K., 2006.
Geochronology and geochemistry of the Francisco Gneiss: Triassic continental rift
tholeiites on the Mexican Margin of Pangea metamorphosed and exhumed in a tertiary core complex. International Geology Reviews 48, 116.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx


Kerr, A.C., Tarney, J., 2005. Tectonic evolution of the Caribbean and northwestern South
America: the case for accretion of two Late Cretaceous oceanic plateaus. Geology
33, 269272.
Kerr, A.C., Marriner, G.F., Tarney, J., Nivia, A., Saunders, A.D., Thirlwall, M.F., Sinton, C.
W., 1997. Cretaceous basaltic terranes in Western Colombia: elemental, chronological and Sr-Nd isotopic constraints on petrogenesis. Journal of Petrology 38,
677702.
Kerr, A.C., Aspden, J.A., Tarney, J., Pilatasig, L.F., 2002. The nature and provenance of accreted oceanic terranes in western Ecuador: geochemical and tectonic constraints. Journal of the Geological Society 159, 577594.
Kirsch, M., Keppie, J.D., Murphy, J.B., Solari, L.A., 2012. PermianCarboniferous arc
magmatism and basin evolution along the western margin of Pangaea: geochemical
and geochronological evidence from the eastern Acatln Complex, southern Mexico.
Geological Society of America Bulletin 124, 16071628.
Kirsch, M., Helbig, M., Keppie, J.D., Murphy, J.B., Lee, J.K.W., Solari, L.A., 2014. A Late Triassic tectonothermal event in the eastern Acatln Complex, southern Mexico, synchronous with a magmatic arc hiatus: the result of at-slab subduction? Lithosphere 6,
6379.
Lapierre, H., Bosch, D., Dupuis, V., Polv, M., Maury, R.C., Hernandez, J., Moni, P.,
Yeghicheyan, D., Jaillard, E., Tardy, M., de Lpinay, B.M., Mamberti, M., Desmet, A.,
Keller, F., Snebier, F., 2000. Mantle plume events in the genesis of the periCaribbean Cretaceous oceanic plateau province. Journal of Geophysical Research
105, 84038421.
Leal-Mejia, H., Draper, J.C.M.I., Shaw, R.P., 2011. Phanerozoic gold metallogeny in the
Colombian Andes. Proceedings let's talk ore deposits, SGA biennial meeting, Antofagasta, Chile.
Lebrat, M., Mgard, F., Dupuy, C., Dostal, J., 1987. Geochemistry and tectonic setting of precollision Cretaceous and Paleogene volcanic rocks of Ecuador. Geological Society of
America Bulletin 99, 569578.
Litherland, M., Aspden, J.A., Jemielita, R.A., 1994. The metamorphic belts of Ecuador. Overseas Memoir of the British Geological Survey, 11, p. 147 (Nottingham, England).
Luzieux, L.D.A., Heller, F., Spikings, R., Vallejo, C.F., Winkler, W., 2006. Origin and Cretaceous tectonic history of the coastal Ecuadorian forearc between 1N and 3S: paleomagnetic, radiometric and fossil evidence. Earth and Planetary Science Letters 249,
400414.
Mamberti, M., 2001. Origin and Evolution of Two Distinct Cretaceous Oceanic Plateaus Accreted in Western Ecuador (South America): Petrological, Geochemical and Isotopic
Evidence(Ph.D. thesis) Universit de Lausanne, p. 241.
Mamberti, M., Lapierre, H., Bosch, D., Jaillard, E., Ethien, R., Hernandez, J., Polv, M., 2003.
Accreted fragments of the Late Cretaceous Caribbean-Colombian Plateau in Ecuador.
Lithos 66, 173199.
Mamberti, M., Lapierre, H., Bosch, D., Jaillard, E., Hernandez, J., Polv, M., 2004. The Early
Cretaceous San Juan Plutonic Suite, Ecuador: a magma chamber in an oceanic plateau? Canadian Journal of Earth Sciences 41, 12371258.
Maniar, P.D., Piccoli, P.M., 1989. Tectonic discrimination of granitoids. Geological Society
of America Bulletin 101, 635643.
Maresch, W.V., Kluge, R., Baumann, A., Pindell, J.L., Krckhans-Lueder, G., Stanek, K., 2009.
The occurrence of high-pressure metamorphism on Margarita Island, Venezuela: a
constraint on Caribbean-South America interaction. Geological Society of London
Special Publications 328, 705741.
Martnez, A.M.C., 2007. Petrogenesis and evolution of Aburra Ophiolite, Colombian Andes,
Central Range(Ph.D. thesis) University of Brasilia, p. 178.
Martin-Gombojav, N., Winkler, W., 2008. Recycling of Proterozoic crust in the Andean
Amazon foreland of Ecuador: implications for orogenic development of the Northern
Andes. Terra Nova 20, 2231.
Maya, M., Gonzlez, H., 1995. Unidades Litodmicas en la Cordillera Central de Colombia.
Informe Unidad Operativa Medelln, Ingeominas, Colombia, pp. 4457.
McCourt, W.J., 1981. The geochemistry and petrography of the Coastal Batholith of Peru,
Lima Segment. Journal of the Geological Society 138, 407420.
McCourt, W.J., Aspden, J.A., Brook, M., 1984. New geological and geochemical data from
the Colombian Andes: continental growth by multiple accretion. Journal of the Geological Society 831845.
Mgard, F., 1984. The Andean orogenic period and its major structures in central and
northern Peru. Journal of the Geological Society of London 141, 893900.
Mikovi, A., Schaltegger, U., 2009. Crustal growth along a non-collisional cratonic margin: a Lu-Hf isotopic survey of the Eastern Cordilleran granitoids of Peru. Earth and
Planetary Science Letters 279, 303315.
Mikovi, A., Spikings, R.A., Chew, D.M., Koler, J., Ulianov, A., Schaltegger, U., 2009.
Tectonomagmatic evolution of Western Amazonia: geochemical characterization
and zircon U-Pb geochronologic constraints from the Peruvian Eastern Cordilleran
granitoids. Geological Society of America Bulletin 121, 12981324.
Montes, C., Guzman, G., Bayona, G., Cardona, A., Valencia, V., Jaramillo, C., 2010. Clockwise
rotation of the Santa Marta massif and simultaneous Paleogene to Neogene deformation of the Plato-San Jorge and Cesar-Ranchera basins. Journal of South American
Earth Sciences 29, 832848.
Moreno, M., Pardo, A., 2003. Stratigraphical and sedimentological constraints on Western
Colombia: implications on the evolution of the Caribbean plate. In: Bartolini, C.,
Bufer, R.T., Blickwede, J. (Eds.), The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habitats, Basin Formation, and Plate Tectonics. American Association of
Petroleum Geologists Memoir, 79, pp. 891924.
Mourier, T., Laj, C., Mgard, F., Roperch, P., Mitouard, P., Farfan Medrano, A., 1988. An accreted continental terrane in northwestern Peru. Earth and Planetary Science Letters
88, 182192.
Mukasa, S.B., 1986. Zircon U-Pb ages of super-units in the Coastal batholith, Peru: implications for magmatic and tectonic processes. Geological Society of America Bulletin
97, 241254.

43

Nivia, A., Marriner, G.F., Kerr, A.C., Tarney, J., 2006. The Quebradagrande Complex: a Lower
Cretaceous ensialic marginal basin in the Central Cordillera of the Colombian Andes.
Journal of South American Earth Sciences 21, 423436.
Noble, S.R., Aspden, J.A., Jemielita, R., 1997. Northern Andean crustal evolution: new U-Pb
geochronological constraints from Ecuador. Geological Society of America Bulletin
109, 789798.
Oliver, N.H.S., Zakowski, S., 1995. Timing and geometry of deformation, low-pressure
metamorphism and anatexis in the eastern Mt. Lofty Ranges: the possible role of extension. Australian Journal of Earth Sciences 42, 501507.
Oliveros, V., Feraud, G., Aguirre, L., Fornari, M., Morata, D., 2006. The early Andean magmatic province (EAMP): Ar-40/Ar-39 dating on mesozoic volcanic and plutonic
rocks from the Coastal Cordillera, Northern Chile. Journal of Volcanology and Geothermal Research 157, 311330.
Omarini, R.H., Sureda, R.J., Gotze, H.J., Seilacher, A., Puger, F., 1999. Puncoviscana folded
belt in northwestern Argentina: testimony of Late Proterozoic Rodinia fragmentation
and pre-Gondwana collisional episodes. International Journal of Earth Sciences 88,
7697.
Orrego, A., Restrepo, J.J., Toussaint, J.F., Linares, E., 1980. Datacin de un esquisto serictico
de Jambal, Cauca. , 25. Geologa Universidad Nacional, pp. 133134 (Special
Publications).
Ortega-Obregon, C., Solari, L., Gomez-Tuena, A., Elias-Herrera, M., Ortega-Gutierrez, F.,
Macias-Romo, C., 2013. PermianCarboniferous arc magmatism in southern
Mexico: U-Pb dating, trace element and Hf isotopic evidence on zircons of earliest
subduction beneath the western margin of Gondwana. International Journal of
Earth Sciences. http://dx.doi.org/10.1007/s00531-013-0933-1.
Pankhurst, R.J., Rapela, C.W., Fanning, C.M., 2000. Age and origin of coeval TTG, I- and Stype granites in the Famatinian belt of NW Argentina. Transactions of the Royal Society of Edinburgh: Earth Sciences 91, 151168.
Parson, L.M., Wright, I.C., 1996. The Lau-Havre-Taupo back-arc basin: a southwardpropagating, multi-stage evolution from rifting to spreading. Tectonophysics 263,
122.
Peccerillo, A., Taylor, S.R., 1976. Geochemistry of Eocene calc-alkaline volcanic rocks from
the Kastamonu area, Northern Turkey. Contributions to Mineralogy and Petrology 58,
6381.
Pindell, J.L., Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and
northern South America in the mantle reference frame: an update. Geological Society
of London Special Publications 328, 155.
Pindell, J.L., Kennan, L., Maresch, W.V., Stanek, K.P., Draper, G., Higgs, R., 2005. Plate kinematics and crustal dynamics of circum-Caribbean arccontinent interactions: tectonic controls on basin development in Proto-Caribbean margins. In: Lallemant, A.,
Sisson, V.B. (Eds.), CaribbeanSouth American Plate Interactions. Geological Society
of America Special Paper, 394, pp. 752.
Polliand, M., Schaltegger, U., Frank, M., Fontbot, L., 2005. Formation of intra-arc
volcanosedimentary basins in the western ank of the central Peruvian Andes during
Late Cretaceous oblique subduction: eld evidence and constraints from U-Pb ages
and Hf isotopes. International Journal of Earth Sciences 94, 231232.
Pratt, W.T., Duque, P., Ponce, M., 2005. An autochthonous geological model for the eastern
Andes of Ecuador. Tectonophysics 399, 251278.
Ramos, V.A., 2008. The basement of the Central Andes: the Arequipa and related terranes.
Annual Reviews of Earth and Planetary Sciences 36, 289324.
Ramos, V.A., 2009. Anatomy and global context of the Andes: main geologic features and
the Andean orogenic cycle. The Geological Society of America Memoir 204, 3165.
Ramos, V.A., Aleman, A., 2000. Tectonic evolution of the Andes. In: Cordani, U.G., Milani, E.
J., Thomaz Filha, A., Campos, D.A. (Eds.), Tectonic Evolution of South America. 31st International Geological Congress, Rio de Janerio, pp. 635685.
Rapela, C.W., Pankhurst, R.J., Fanning, C.M., Herve, F., 2005. Pacic subduction coeval with
the Karoo mantle plume: the early Jurassic subcordilleran belt of northwestern Patagonia. In: Vaughan, A.P.M., Leat, P.T., Pankhurst, R.J. (Eds.), Terrane processes at the
margins of Gondwana. Geological Society of London, 246, pp. 217239 (Geological
Society Special Publications).
Reitsma, M.J., 2012. Reconstructing the Late PaleozoicEarly Mesozoic Plutonic and Sedimentary Record of South-East Peru: Orphaned Back-Arcs Along the Western Margin
of Gondwana(PhD thesis) Terre & Environment, 111. University of Geneva,
Switzerland, p. 226.
Restrepo, J.J., Toussaint, J.F., 1976. Edades radiomtricas de algunas rocas de Antioquia,
Colombia. Publicacion Especial Geologa. , 12. Universidad Nacional de Medellin,
pp. 111.
Restrepo, J.J., Toussaint, J.F., 1982. Metamorsmos superpuestos em la Cordillera Central
de Colombia. V Congreso Latino-Americano de Geologa, Buenos Aires, Argentina,
pp. 505512.
Restrepo, J.J., Toussaint, J.F., 1988. Terranes and continental accretion in the Colombian
Andes. Episodes 7, 189193.
Restrepo, J.J., Toussaint, J.F., Gonzles, H., Cordani, U., Kawashita, K., Linares, E., Parica, C.,
1991. Precisiones geocronolgicas sobre el occidente Colombiano. Simposio sobre
magmatismo andino y su marco tectnico. Memorias, Tomo 1, Manizales,
Colombia, pp. 122.
Restrepo, J.J., Ordez-Carmona, O., Moreno-Snchez, M., 2009. A comment on The
Quebradagrande Complex: a Lower Cretaceous ensialic marginal basin in the Central
Cordillera of the Colombian Andes by Nivia et al. Journal of South American Earth
Sciences 28, 204205.
Restrepo, J.J., Ordoez-Carmona, O., Armstrong, R., Pimentel, M.M., 2011. Triassic metamorphism in the northern part of the Taham Terrane of the central cordillera of
Colombia. Journal of South American Earth Sciences 32, 497507.
Restrepo-Pace, P.A., Cediel, F., 2010. Northern South America basement tectonics and implications for paleocontinental reconstructions of the Americas. Journal of South
American Earth Sciences 29, 764771.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

44

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx

Restrepo-Pace, P.A., Ruiz, J., Gehrels, G., Cosca, M., 1997. Geochronology and Nd isotopic
data for Grenville-age rocks in the Colombian Andes: new constraints for Late
ProterozoicEarly Paleozoic paleocontinental reconstructions of the Americas. Earth
and Planetary Science Letters 150, 427441.
Reynaud, C., Jaillard, E., Lapierre, H., Mamberti, M., Mascle, G.H., 1999. Oceanic plateau and
island arcs of southwestern Ecuador: their place in the geodynamic evolution of
northwestern South America. Tectonophysics 307, 235254.
Riding, J.B., 1989. A palynological investigation of nine rock samples from Ecuador
(Maguazo Unit). British Geological Survey Technical Report WH/89/361/R (4 pp.).
Riel, N., Guillot, S., Jaillard, E., Martelat, J.-E., Paquette, J.-L., Schwartz, S., Gonclaves, P.,
Duclaux, G., Thebaud, N., Lanari, P., Janots, E., Yuquilema, J., 2013. Metamorphic and
geochronological study of the Triassic El Oro metamorphic complex, Ecuador: implications for high-temperature metamorphism in a forearc zone. Lithos 156159,
4168.
Rodriguez, G., Zapata, G., 2013. Comparative analysis of the Barroso Formation and
Quebradagrande Complex: a volcanic arc tholeiiticcalcoalcaline, segmented by
the fault system Romeral in Northern Andes. Boletin Ciencias de la Tierra 33,
3958.
Romero, D., Valencia, K., Alarcn, P., Pea, D., Ramos, V.A., 2013. The offshore basement of
Peru: evidence for different igneous and metamorphic domains in the forearc. Journal
of South American Earth Sciences 42, 4760.
Romeuf, N., Aguirre, L., Soler, P., Fraud, G., Jaillard, E., Ruffet, G., 1995. Middle Jurassic volcanism in the Northern and Central Andes. Revisita Geologica de Chile 22, 245259.
Rosas, S., Fontbot, L., Tankard, A., 2007. Tectonic evolution and paleogeography of the
Mesozoic Pucar Basin, central Peru. Journal of South American Earth Sciences 24,
124.
Rosenbaum, G., Giles, D., Saxon, M., Betts, P.G., Weinberg, R.F., Dubox, C., 2005. Subduction
of the Nazca Ridge and the Inca Plateau: insights into the formation of ore deposits in
Peru. Earth and Planetary Science Letters 239, 1832.
Rubatto, D., 2002. Zircon trace element geochemistry: partitioning with garnet and the
link between U-Pb ages and metamorphism. Chemical Geology 184, 123138.
Ruiz, J., Tosdal, R.M., Restrepo, P.A., Murillo-Muetn, G., 1999. Pb isotope evidence for
Columbiasouthern Mexico connections in the Proterozoic. In: Ramos, V.A., Keppie,
J.D. (Eds.), LaurentiaGondwana Connections Before Pangea. Geological Society of
America Special Paper, 336, pp. 183197.
Ruiz, G.M.H., Seward, D., Winkler, W., 2004. Detrital thermochronology; a new perspective on hinterland tectonics, an example from the Andean Amazon Basin, Ecuador.
Basin Research 16, 413430.
Ruiz, G., Seward, D., Winkler, W., Maria, A.M., David, T.W., 2007. Evolution of the Amazon
Basin in Ecuador with special reference to hinterland tectonics: data from zircon
ssion-track and heavy mineral analysis. Developments in Sedimentology 58,
907934.
Sarmiento, L.F., Rangel, A., 2004. Petroleum systems of the Upper Magdalena Valley,
Colombia. Marine and Petroleum Geology 21, 373391.
Sarmiento-Rojas, L.F., Van Wess, J.D., Cloetingh, S., 2006. Mesozoic transtensional basin
history of the Eastern Cordillera, Colombian Andes: inferences from tectonic models.
Journal of South American Earth Sciences 21, 383411.
Saur, W., 1950. Mapa geolgica del Ecuador 1:500 000. Quito, Universidad Central and
Direccin de Minera.
Schaaf, P., Weber, B., Weis, P., Gro, A., Ortega-Gutirrez, F., Khler, H., 2002. The Chiapas
Massif (Mxico) revised; new geologic and isotopic data for basement characteristics.
In: Miller, H. (Ed.), Contributions to Latin-American Geology: Neues Jahrbuch fr
Geologie und Palontologie. Abhandlungen, 225, pp. 123.
Scheuber, E., Gonzalez, G., 1999. Tectonics of the JurassicEarly Cretaceous magmatic arc
of the north Chilean Coastal Cordillera (2226S): a story of crustal deformation
along a convergent plate boundary. Tectonics 18, 895910.
Schlische, R.W., 2002. Progress in understanding the structural geology, basin evolution,
and tectonic history of the eastern North American rift system. Intorsvik In:
LeTourneau, P.M., Olsen, P.E. (Eds.), The Great Rift Valleys of Pangaea in Eastern
North America, 1. Columbia University Press, New York.
Schtte, P., Chiaradia, M., Beate, B., 2010. Geodynamic controls on Tertiary arc magmatism
in Ecuador: constraints from U-Pb zircon geochronology of OligoceneMiocene intrusions and regional age distribution trends. Tectonophysics 489, 159176.
Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto, V., Jacay, J., Arispe, O., Nraudeau, D.,
Crdenas, J., Rosas, S., Jimnez, N., 2002. Late PermianMiddle Jurassic lithospheric
thinning in Peru and Bolivia, and its bearing on Andean-age tectonics. Tectonophysics
345, 153181.
Sheppard, G., Bushnell, G.H.S., 1933. Metamorphic rocks of the eastern Andes near
Cuenca, Ecuador. Geological Magazine 70, 321330.
Sherlock, S.C., Kelley, S.P., 2002. Excess argon evolution in HP-LT rocks: a UVLAMP study
of phengite and K-free minerals, NW Turkey. Chemical Geology 182, 619636.
Shervais, J.W., 1982. Ti-V plots and the petrogenesis of modern and ophiolitic lavas. Earth
and Planetary Science Letters 59, 101118.
Sinton, C.W., Duncan, R.A., Storey, M., Lewis, J., Estrada, J.J., 1998. An oceanic ood basalt
province within the Caribbean plate. Earth and Planetary Science Letters 155,
221235.
Slma, J., Koler, J., Condon, D., Crowley, J.L., Gerdes, A., Hanchar, J.M., Horstwood, M.S.A.,
Morris, G.A., Nasdala, G.A., Norberg, L., Schaltegger, U., Schoene, B., Turbrett, M.N.,
Whitehouse, M.J., 2008. Pleovice zirconA new natural reference material for U-Pb
and Hf isotopic microanalysis. Chemical Geology 249, 135.
Solari, L.A., Dostal, J., Ortega-Gutierrez, F., Keppie, J.D., 2001. The 275 Ma arc-related La
Carbonera stock in the northern Oaxacan Complex of Southern Mexico: U-Pb geochronology and geochemistry. Revista Mexicana de Ciencias Geolgicas 18, 149161.
Solari, L.A., Gmez-Tuena, A., Ortega-Gutirrez, F., Ortega-Obregn, C., 2011. The Chaucs
Metamorphic Complex, central Guatemala: geochronological and geochemical constraints on its PaleozoicMesozoic evolution. Geological Acta 9, 329350.

Spadea, P., Espinosa, E., Orrego, A., 1989. High-Mg extrusive rocks from the Romeral
zone ophiolites in the south western Colombian Andes. Chemical Geology 77,
303321.
Spikings, R.A., Crowhurst, P.V., 2004. (U-Th)/He thermochronometric constraints on the
late Miocene-Pliocene tectonic development of the northern Cordillera Real and the
Interandean Depression, Ecuador. Journal of South American Earth Sciences 17,
239251.
Spikings, R.A., Seward, D., Winkler, W., Ruiz, G.M., 2000. Low-temperature
thermochronology of the northern Cordillera Real, Ecuador: tectonic insights
from zircon and apatite ssion track analysis. Tectonics 19, 649668.
Spikings, R.A., Winkler, W., Seward, D., Handler, R., 2001. Along-strike variations in
the thermal and tectonic response of the continental Ecuadorian Andes to the
collision with heterogeneous oceanic crust. Earth and Planetary Science Letters
186, 5773.
Spikings, R.A., Winkler, W., Hughes, R.A., Handler, R., 2005. Thermochronology of allochthonous terranes in Ecuador: unravelling the accretionary and post-accretionary history of the Northern Andes. Tectonophysics 399, 195220.
Spikings, R.A., Crowhurst, P.V., Winkler, W., Villagomez, D., 2010. Syn- and post accretionary cooling history of the Ecuadorian Andes constrained by their in-situ and detrital
thermochronometric record. Journal of South American Earth Sciences 30, 121133.
Sun, S.S., McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts:
implications for mantle composition and processes. Geological Society, London, Special Publications 42, 313345.
Taylor, S.R., McLennan, S.M., 1995. The geochemical evolution of the continental crust. Reviews of Geophysics 33, 241265.
Torres, R., Ruz, J., Patchett, P.J., Grajales-Nishimura, J.M., 1999. Permo-Triassic continental
arc in eastern Mexico; tectonic implications for reconstructions of southern North
America. In: Bartolini, C., Wilson, J.L., Lawton, T.F. (Eds.), Mesozoic Sedimentary and
Tectonic History of North-Central Mexico. Geological Society of America Special
Paper, 340, pp. 191196.
Toussaint, J.F., Restrepo, J.J., 1994. The Colombian Andes during Cretaceous times. In:
Salty, J.A. (Ed.), Cretaceous Tectonics of the Andes. Earth Evolution Series. Vieweg
and Teubner Verlag, pp. 61100.
Vallejo, C., Spikings, R.A., Luzieux, L., Winkler, W., Chew, D., Page, L., 2006. The early interaction between the Caribbean Plateau and the NW South American Plate. Terra Nova
18, 264269.
Vallejo, C., Winkler, W., Spikings, R.A., Luzieux, L., Heller, F., Bussy, F., 2009. Mode and
timing of terrane accretion in the forearc of the Andes of Ecuador. The Geological Society of America Memoir 204, 197216.
Van der Lelij, R., 2013. Reconstructing North-Western Gondwana With Implications
for the Evolution of the Iapetus and Rheic Oceans: A Geochronological,
Thermochronological and Geochemical study(PhD thesis) Terre & Environment,
121. University of Geneva, Switzerland, p. 221.
Van der Lelij, R., Spikings, R.A., Kerr, A.C., Kounov, A., Cosca, M., Chew, D., Villagomez, D.,
2010. Thermochronology and tectonics of the Leeward Antilles: evolution of the
southern Caribbean Plate boundary zone. Tectonics 29. http://dx.doi.org/10.1029/
2009tc002654.
Vsquez, M., Altenberger, U., Romer, R.L., Sudo, M., Moreno-Murillo, J.M., 2010. Magmatic
evolution of the Andean Eastern Cordillera of Colombia during the Cretaceous: Inuence of previous tectonic processes. Journal of South American Earth Sciences 29,
171186.
Vsquez, M., Altenberger, U., 2005. Mid-Cretaceous extension-related magmatism in
the eastern Colombian Andes. Journal of South American Earth Sciences 20,
193210.
Vergara, L., Prssl, K.F., 1994. Dating the Yav Formation (Aptian, Upper Magdalena Valley,
Colombia), palynological results. In: Etayo, F. (Ed.), Estudios Geolgicos del Valle Superior del Magdalena. Departamento de Geociencias, Universidad Nacional de
Colombia, Bogot, p. 14.
Villagmez, D., Spikings, R., 2013. Thermochronology and tectonics of the Central and
Western Cordilleras of Colombia: Early CretaceousTertiary evolution of the Northern Andes. Lithos 168, 228249.
Villagmez, D., Spikings, R., Magna, T., Kammer, A., Winkler, W., Beltrn, A., 2011. Geochronology, geochemistry and tectonic evolution of the Western and Central cordilleras of Colombia. Lithos 125, 875896.
Villamil, T., 1999. CampanianMiocene tectonostratigraphy, depocenter evolution and
basin development of Colombia and western Venezuela. Palaeogeography,
Palaeoclimatology, Palaeoecology 153, 239275.
Vinasco, C.J., Cordani, U.G., Gonzales, H., Weber, M., Pelaez, C., 2006. Geochronological,
isotopic, and geochemical data from Permo-Triassic granitic gneisses and granitoids
of the Colombian Central Andes. Journal of South American Earth Sciences 21,
355371.
Viscarret, P., Wright, J., Urbani, F., 2009. New U-Pb zircon ages of El Bal Massif, Cojedes
State, Venezuela. Revisita Tcnica de la Facultad de Ingenieria Universidad del Zulia
32, 210221.
Weber, B., Iriondo, A., Premo, W.R., Hecht, L., Schaaf, P., 2007. New insights into the history and origin of the southern Maya block, SE Mexico: U-Pb-SHRIMP zircon geochronology from metamorphic rocks of the Chiapas massif. International Journal of Earth
Sciences 96, 253269.
Weber, M., Cardona, A., Valencia, V., Garca-Casco, A., Tobn, M., Zapata, S., 2010. U/Pb detrital zircon provenance from late Cretaceous metamorphic units of the Guajira Peninsula, Colombia: tectonic implications on the collision between the Caribbean arc
and the South American margin. Journal of South American Earth Sciences 29,
805816.
White, R.V., Tarney, J., Kerr, A.C., Saunders, A.D., Kempton, P.D., Pringle, M.S., Klaver, G.T.,
1999. Modication of an oceanic plateau, Aruba, Dutch Caribbean: implications for
the generation of continental crust. Lithos 46, 4368.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

R. Spikings et al. / Gondwana Research xxx (2014) xxxxxx


Wiedenbeck, M., All, P., Corfu, F., Grifn, W.L., Meier, M., Oberli, F., Von Quadt, A.,
Roddick, J.C., Spiegel, W., 1995. Three natural zircon standards for U-Th-Pb, Lu-Hf,
trace element and REE analyses. Geostandards and Geoanalytical Research 19, 123.
Winkler, W., Villagomez, D., Spikings, R., Abegglen, P., Tobler, S., Egez, A., 2005. The
Chota basin and its signicance for the inception and tectonic setting of the
Interandean depression in Ecuador. Journal of South American Earth Sciences 19,
519.
Winter, Th., Laven, A., 1989. Morphological and microtectonic evidence for a major active right-lateral strikeslip fault across central Ecuador (South America). Annales
Tectonicae 3, 123139.
Wright, J.E., Wyld, S.J., 2004. Aruba and Curaao: remnants of a collided Pacic oceanic
plateau? Initial geologic results from the BOLIVAR project. Eos Transactions of the
American Geophysical Union 85, 47.
Xu, H., Ma, C., Ye, K., 2007. Early cretaceous granitoids and their implications for the collapse of the Dabie orogen, eastern China: SHRIMP zircon U-Pb dating and geochemistry. Chemical Geology 240, 238259.
Yanez, P., Ruiz, J., Patchett, P.J., Ortega-Gutierrez, F., Gehrels, G.E., 1991. Isotopic studies of
the Acatlan complex, southern Mexico: implications for Paleozoic North American
tectonics. Geological Society of America Bulletin 103, 817828.
Zapata, J.P., Restrepo, J.J., Martens, U., Cardona, A., Brito, R., 2011. Geochronology and geochemistry of the basic sequence of Altamira, Antioquia, Western Cordillera of
Colombia. 14th Congresso Colombiano de Geologa, Medellin, Colombia.
Zerfass, H., Chemale, F., Schultz, C.L., Lavina, E., 2004. Tectonics and sedimentation in
Southern South America during Triassic. Sedimentary Geology 166, 265292.

Dr. Richard Spikings graduated in geochemistry at the


University of St. Andrews in 1993. His research in
thermochronology earned a PhD in geology in 1998 from
La Trobe University, Melbourne. Since 1998, he has worked
as a postdoctoral fellow at the ETH-Zrich, and as tenured research staff at the University of Geneva where he manages
the 40Ar/39Ar laboratory. His research has focussed on
thermochronology and geochronology of the Andean cordilleras in Ecuador, Colombia, Venezuela, Peru and Chile. More
recently, Richard has focussed his research efforts on bulk
and in-situ U-Pb thermochronology of accessory phases.

Dr. Ryan Cochrane graduated in geology at the University of


Johannesburg in 2005, after which he worked as a gold exploration geologist in South Africa and Western China. He
completed a BSc(Hons) degree at the University of Cape in
2008 and went on to earn a PhD in tectonics, isotope geochemistry and thermochronology from the University of
Geneva in mid-2013. Ryan immediately started work at
Thomson Reuters GFMS in London as a Mine Economics Analyst with a focus on precious metals and mining research,
including the maintenance of mine economics products. Recently, Ryan was promoted to a Senior Analyst and focusses
on the economics of mining, and deriving corporate valuations for precious metal mining companies

45

Dr. Roelant van der Lelij completed his MSc in the University of Geneva in 2008 where studied the tectonic evolution of
the South Caribbean Plate Boundary. As part of his PhD project between 2008 and 2013 in the University of Geneva, he
worked on the Phanerozoic tectonic history of igneous and
metamorphic basement rocks exposed in the Santander
Massif of Colombia and the Merida Andes of Venezuela. Since
March 2014 he manages the new K/Ar geochronology laboratory in the Geological Survey of Norway in Trondheim,
which aims to constrain the history of brittle tectonics and
landscape evolution of onshore and offshore Norway.

Dr. Cristian Vallejo graduated at the Escuela Politcnica


Nacional-Quito, and obtained a PhD from ETH-Zrich, studying the Geodynamics of the Western Andes of Ecuador and
its relationship with the collision of the Caribbean Plateau.
After his PhD he worked as a Research Fellow on the Stratigraphic Development of Slope Systems Consortium Research
Project at the University of Aberdeen. Since 2009 he has been
working as a consultant on tectonics, sedimentology and
mineral exploration in South and Central America. Cristian
is also a part time lecturer at the Escuela Politcnica Nacional,
Quito.

Prof. Wilfried Winkler has held a position of senior lecturer


and professor for sedimentary petrology, basin analysis and
sedimentology at the ETH Zurich since 1988. He graduated
in 1977 at the University of Fribourg, Switzerland, where he
subsequently obtained a PhD in geology in 1981. From
1981 to 1988, he was a postdoctoral fellow and lecturer at
both Fribourg and Basel Universities. His main research interests are the bearings of tectonics and climate on basin
sedimentation, and provenance studies in different plate tectonics settings. He carried out research in the Alps,
Carpathians, Pyrenees, Northern Andes (Ecuador, Peru),
Sinai Peninsula (Egypt), Central Asian Orogenic Belt
(Mongolia) and Myanmar.

Prof. Bernado Beate is a Professor of Geology at the Escuela


Politcnica Nacional, Quito, Ecuador. His main research interests include the petrology of volcanic rocks, geothermal systems and ore geology, and he is a specialist in mapping
volcanic systems and mineral alteration assemblages. Prof.
Beate also operates as a consultant for the mineral resource
and geothermal industry, and has worked in Ecuador,
Colombia, Peru, Bolivia, Paraguay and New Zealand.

Dr. Diego Villagmez is qualied as an Engineer in Geology


(EPN, Quito, Ecuador) and a PhD in Geology from the University of Geneva (Switzerland). He currently works for Tectonic
Analysis, Ltd. (US). His job involves thermochronological
studies and regional tectonic interpretations for major oil
companies operating in different parts of Latin America from
Mexico to Ecuador. He worked as an external consultant for
the Ecuadorian government on three key national geothermal energy projects, dening the main targets for various
volcano-thermal systems and previously worked on mineral
exploration and engineering geology.

Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the
Caribbean Large Igneous Province (29075 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

Das könnte Ihnen auch gefallen