You are on page 1of 45

GR-01278; No of Pages 45

Gondwana Research xxx (2014) xxx–xxx

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 (290–75 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
b
Thomson Reuters, London, UK
c
Tectonic Analysis Ltd., Geneva, Switzerland
d
Norges Geologiske Undersøkelse, 7491 Trondheim, Norway
e
Geostrat S.A., Quito, Ecuador
f
ETH Zürich, Geological Institute, ETH Zentrum, NO E 59, Sonneggstrasse 5, CH-8092 Zürich, Switzerland
g
Facultad de Geología, Minas y Petróleos, Escuela Politécnica Nacional, A.P. 17-01-2759, Quito, Ecuador

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

Article history: Northwestern South America preserves a record of the assembly of western Pangaea, its disassembly and
Received 24 March 2014 initiation of the far western Tethys Wilson Cycle, subsequent Pacific margin magmatism and ocean plateau–
Received in revised form 4 June 2014 continent interaction since the Late Cretaceous. Numerous models have been presented for various time
Accepted 25 June 2014
slices although they are based on either spatially restricted datasets, or dates that are inaccurate estimates
Available online xxxx
of the time of crystallisation. Here we review a very large quantity of geochronological, geochemical,
Handling Editor: M. Santosh thermochronological, sedimentological and palaeomagnetic data that collectively provide tight constraints for
geological models. These data have been collected over a trench (Pacific)-parallel distance of N1500 km
Keywords: (Colombia and Ecuador), and reveal important temporal trends in rifting and subduction. The temporal frame-
Pangaea work for our model constraints are obtained from robust, concordant zircon U-Pb ages of magmatic rocks during
South America 290–75 Ma. The Late Cretaceous thermal history of the margin (b350 °C) is described by 40Ar/39Ar and fission
Tectonic reconstruction track data, and the higher temperature and thus older (pre-75 Ma) history are constrained by apatite U-Pb
Geochronology thermochronology. Variations in the isotopic compositions of Hf (zircon), Nd (whole) and O (quartz) with
Geochemistry
time have been used to track the evolution of the source of magmatism, and are used as proxies for crustal
Thermochronology
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 floor spreading between blocks of Central and South America by
216 Ma. Pacific active margin commenced at 209 Ma, and continued until 115 Ma above an east-dipping subduc-
tion 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 Pacific 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. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2. Geological framework of northwestern South America (Colombia and Ecuador) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3. Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4. Triassic: the disassembly of Pangaea and the formation of a passive margin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.1. Historical perspective and occurrence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.2. Geochronology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.2.1. Cordillera Real of Ecuador and Cordillera Central of Colombia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0

⁎ 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
2 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

4.2.2. Comparison with the ages of Permian and Triassic rocks in Venezuela and Peru . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.3. Geochemistry of the granites and migmatites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.3.1. Cordillera Real of Ecuador and Cordillera Central of Colombia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.3.2. Comparison with Permian and Triassic rocks in Venezuela and Peru . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.4. Geochemistry of the amphibolites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.5. Zircon Hf isotope geochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.5.1. Zircon Hf isotope geochemistry of the granites and migmatitic leucosomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.5.2. Zircon Hf isotope geochemistry of the amphibolites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.5.3. Comparison with Zircon Hf isotope compositions in Peru . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.6. Thermal histories during the Triassic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.7. Interpretation: Permian and Triassic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.7.1. Arc magmatism and metamorphism during 290–240 Ma along western Pangaea . . . . . . . . . . . . . . . . . . . . . . . . 0
4.7.2. Initiating the disassembly of western Pangaea during 240–200 Ma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.8. Conjugate margins to northwestern Gondwana . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4.9. Rifting between North and South America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5. Latest Triassic–Lower Cretaceous: arc magmatism and tectonic switching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.1. Historical perspective and occurrence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.1.1. Latest Triassic–Jurassic granitoid intrusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.1.2. Late Jurassic–Early Cretaceous rocks to the west of the Jurassic intrusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.2. Geochronology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.2.1. Latest Triassic and Jurassic intrusions: Cordillera Real, Cordillera Central and the Santander Massif . . . . . . . . . . . . . . . . 0
5.2.2. Early Cretaceous magmatic and sedimentary rocks: Cordillera Real and Cordillera Central . . . . . . . . . . . . . . . . . . . . 0
5.2.3. Comparison with Peru and the Merida Andes of Venezuela . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.3. Geochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.3.1. Latest Triassic–earliest Cretaceous granitoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.3.2. Early Cretaceous igneous rocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.3.3. Comparison with magmatic rocks from Peru . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.4. The tectonic setting during the latest Triassic–Jurassic (210–145 Ma) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.4.1. Why is there a gap in the Jurassic arc in Peru? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.5. The tectonic setting during the Early Cretaceous (145–115 Ma) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.6. Compression during the Early Cretaceous . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.7. The Chaucha Terrane and the Tahamí Terrane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5.8. Comparison with Peru (145–115 Ma) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6. The tectonic history of northwestern South America during 115–75 Ma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6.1. The formation of the Caribbean Large Igneous Province and its collision with South America. . . . . . . . . . . . . . . . . . . . . . . . 0
6.1.1. Geochemistry and geochronology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6.1.2. Time of initial accretion with South America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6.1.3. The nature of the CLIP–South America suture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
7. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
Appendix A. Supplementary data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0

1. Introduction Kennan, 2009; Villagómez and Spikings, 2013; Cochrane et al., 2014a).
Contrasting models partly exist because of the misinterpretation of
The northwestern South American plate hosts a Grenvillian base- K/Ar and Rb/Sr dates that were published in the 1980's and 1990's as
ment, which was modified during the amalgamation and disassembly accurate estimates of crystallisation age, ignoring the effects of daughter
of Pangaea, subsequent prolonged active margin magmatism and the isotope loss. We discard K/Ar and Rb/Sr dates in favour of recently pub-
collision of the voluminous Caribbean Large Igneous Province, which lished concordant zircon U-Pb dates, which are more accurate estimates
added new crust to South America. This manuscript is mainly a review of crystallisation age. The U-Pb dates are combined with geochemical
of a very large quantity of data, although some new U-Pb (apatite) and isotope data, sedimentological data and field relationships to
and 40Ar/39Ar dates are presented. These data are used to generate constrain the magmatic source regions and tectonic environment
robust constraints for any model that describes the disassembly within which the rocks formed. The tectonic histories are subsequently
and fragmentation of western Pangaea, the subsequent evolution of investigated using thermochronological and palaeomagnetic data.
the Pacific margin offshore northwestern South America during the We show that western Pangaea started to disassemble by rifting of
Jurassic–Early Cretaceous, and the early evolution of the Caribbean continental crust of Central America from South America at ~ 240 Ma,
region and its interaction with South America. The review is organised and that these had completely separated by ~216 Ma. The northwestern
into sections according to geological time, and compares the evolution margin of South America remained passive until ~ 209 Ma within
of northwestern South America (north of 5°S) with the margin of Peru Pangaea, and arc magmatism occurred during 209–114 Ma, accompa-
during 290–75 Ma. nying the separation of North and South America at ~ 180 Ma. The
Wide disagreements exist over the tectonic origin of voluminous Jurassic magmas formed in a continental arc, which questions previous
magmatic units, including Triassic anatectites, Jurassic–Early Cretaceous interpretations that place the Jurassic trench far from the location of the
arc rocks, obducted M-HP/LT rocks and allochthonous units that com- Jurassic arcs, due to the presence of suspect continental terranes. We
prise the western cordilleras and the forearc. These contrasting inter- draw a single east-dipping subduction zone during 209–114 Ma,
pretations result in significantly different interpretations for plate which retreated oceanward and extended the South American margin,
reconstructions during the Triassic–Late Cretaceous (e.g. Litherland culminating in compression that drove rock uplift and exhumation.
et al., 1994; Spikings et al., 2001; Pratt et al., 2005; Pindell and 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 3

formed on an intra-oceanic plateau prior to its collision and accretion autochthonous, which is similar to the model proposed by Villagómez
with South America at ~ 75 Ma, resulting in growth of the continent. and Spikings (2013) and Cochrane (2013) for Ecuador and Colombia.
Our interpretation differs from other models that rely on large-scale The retro-arc foreland basins of the Middle and Lower Magdallena
plate reconstructions to constrain the positions of continents and sub- Valley basins in Colombia, and the Oriente Basin of Ecuador preserve a
duction zones. record of the evolution of the margin during the Jurassic–Recent, and
these are utilised throughout the review to support or negate various
hypotheses.
2. Geological framework of northwestern South America (Colombia
and Ecuador) 3. Methodology

The South American Plate forms a relict part of western Gondwana, This review presents a very large quantity of data that was mainly
and formed during the opening of the Central Atlantic, South Atlantic previously peer reviewed and published, and the details of the method-
and the Inter-American Gap (Gulf of Mexico and the proto-Caribbean) ologies used by each study are provided in the respective publications. A
during 180–120 Ma (E.g. Eagles, 2007). The Atlantic margin remains summary of the geochronological data and pertinent geochemical and
passive whereas the western margin became active at ~ 500–480 Ma isotopic data is presented in Tables 1 and 2, and the complete geochem-
(e.g. Pankhurst et al., 2000; Van der Lelij, 2013), soon after the opening ical dataset used to plot all of the geochemical figures is provided as a
of the Iapetus Ocean during 570–535 Ma (Cawood et al., 2001). The supplementary file.
Northern plate margin formed during rifting from Yucatan in the Middle A majority of the geochronological, isotopic and geochemical
Jurassic (Pindell et al., 2005), forming the region of the proto-Caribbean. data for the period spanning 290–100 Ma (Mišković et al., 2009;
At the present time, the Northern Andes are separated from the Central Villagómez et al., 2011; Cochrane, 2013; Cochrane et al., 2014a,b) was
Andes by the Huancabamba Deflection (at ~6°S), which marks a distinct acquired at laboratories at the Universities of Geneva and Lausanne
change in the strike orientation of the Andean chain (see inset in Fig. 1). (Switzerland), and at the Goethe Universität Frankfurt. Accuracy and
The oldest rocks exposed in northwestern South America are external reproducibility of the methods were determined by analysing
Grenvillian aged gneisses, which are dispersed in inliers throughout i) Harvard 91500 (Wiedenbeck et al., 1995) and Plešovice zircon
the Eastern Cordillera of Colombia, Santander Massif and the Sierra (Sláma et al., 2008) during geochronology, ii) GJ-1 and Plešovice zircon
Nevada de Santa Marta (Fig. 1; e.g. Restrepo-Pace et al., 1997), where (Sláma et al., 2008) during Hf isotopic analyses, iii) standard JNdi-1 for
they are considered to form part of the Chibcha Terrane (e.g. Toussaint Nd isotopic analyses. Other geochronological analyses were performed
and Restrepo, 1994), although none crop out in Ecuador. The Phanerozoic at laboratories at Curtin University, the University of Grenoble (Riel
rocks, which form the focus of this review, can be separated into a et al., 2013), the Australian National University (Vinasco et al., 2006;
relatively undifferentiated oceanic Late Cretaceous sequence, which is Restrepo et al., 2011), the University of Arizona (Bustamante et al.,
faulted against older, differentiated continental crust. 2010; Cardona et al., 2010; Weber et al., 2010) and Westfälische
The Late Cretaceous oceanic rocks form the basement to the Western Wilhelms-Universität Munster (Lu-Hf; John et al., 2010). All zircon
Cordillera and the forearcs of Colombia (Calima Terrane; Fig. 1; e.g. Kerr U-Pb dates that are used in the interpretations are concordant, and
et al., 1997) and Ecuador (Pallatanga-Piñon Terrane; e.g. Vallejo et al., define a single age population with respect to their mean square
2009). Geochemical, isotopic and geochronological data suggest that weighted deviate statistic. The methodology used for each measure-
these ultramafic and mafic rocks formed in an oceanic hot-spot setting ment is highlighted in Tables 1 and 2.
during 99–87 Ma (e.g. Kerr et al., 1997; Vallejo et al., 2006; Villagómez Dense mineral assemblages and fission-track data for the period
et al., 2011), and that they are equivalent to the oceanic plateau rocks after 100 Ma (e.g. Spikings et al., 2000, 2010) were obtained at laborato-
that form the Caribbean Plate (e.g. Sinton et al., 1998). Field relation- ries at ETH-Zürich. A majority of geochemical and geochronological data
ships and U-Pb zircon dates show that the oceanic plateau was intruded from the Caribbean Large Igneous Province was obtained from laborato-
by an east-facing intra-oceanic arc prior to their collision with South ries at the University of Geneva (40Ar/39Ar; Luzieux et al., 2006), the
America in the Campanian (e.g. Feininger and Bristow, 1980; Vallejo Australian National University (U-Pb; Vallejo et al., 2006), and the
et al., 2006), although this is not consistent with the plate reconstruc- University of Lausanne (Mamberti et al., 2003).
tions 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 4. Triassic: the disassembly of Pangaea and the formation of a
continental margin of South America. The accretion of allochthons of passive margin
the Caribbean Large Igneous Province added at least 5x106–1x107 km3
of new crust to the South American Plate (Cochrane, 2013). Triassic rocks within the Cordillera Real, Amotape Terrane (Ecuador)
The Early Cretaceous continental margin hosts N-S trending linear and Cordillera Central (Colombia) are dominated by widely dispersed
belts that are exposed within the Cordillera Central of Colombia, outcrops of variably foliated granites, gneissic granites and migmatites,
and the Cordillera Real of Ecuador (Fig. 1). Traversing eastwards from and less abundant amphibolites, ultramafic rocks and meta-
the Campanian suture, these are M-HP/LT complexes of amphibolites, sedimentary rocks (Fig. 2). Several studies have shown that the
blueschists and eclogites, Early Cretaceous arc rocks of the magmatic and metamorphic rocks that formed during the Triassic are
Quebradagrande and Alao sequences, undifferentiated Palaeozoic geochemically distinct from younger magmatic rocks, and formed in a
rocks, which underwent anatexis in the Triassic, foliated Early different tectonic environment.
Cretaceous arc plutons, and large unfoliated Jurassic batholiths along
the eastern flank of the Cordillera Central in Colombia, and Cordillera 4.1. Historical perspective and occurrence
Real in Ecuador. Further east, the Eastern Cordillera of Colombia has
no equivalent topographic feature in Ecuador, and it includes the high Triassic rocks within the Cordillera Real and Amotape Terrane of
plains of the Santander Massif in the north (Fig. 1). The Santander Ecuador include granitoids of the Tres Lagunas and Moromoro units,
Massif hosts the oldest segment of the latest Triassic–Early Cretaceous migmatites of the Sabanilla unit, geographically scattered amphibolitic
continental arc sequence, which has no equivalent assemblage in dykes and sills (Piedras and Monte Olivo units), and sedimentary
Ecuador. Litherland et al. (1994) suggest that these belts are all in rocks of the Piuntza unit (Fig. 2). The granites were first described by
tectonic contact, and these allochthonous units were juxtaposed during Colony and Sinclair (1932). Mapping by the British Geological Survey
compression at 140–120 Ma. Alternatively, Pratt et al. (2005) suggest during 1986–1993 linked these occurrences into a semi-continuous
that the contacts are intrusive, and the rock units within Ecuador are 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
4 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

GP

Caribbean Plate
SNSM

SM
10°N SP

F
PSB

Venezuela

SM MA

Chocó-
Panamá

CAF
Terrane
OPF MMV

Tahami
GF Terrane
EC

F
Pa
5°N

IF
UMV
WC
CC LB
0m
300
CPV
Huancabamba
F

Calima
SJ

Deflection
Terrane
Nazca Plate Coastal
Batholith
Arequipa
Colombia dg a
Ri azc Terrane
e
N
F

South American Plate


Pu

0° 0m Pallatanga-Piñon
300 Terrane
IAD CR
Carnegie WC SZ
Ridge SLB PF
Ecuador
OB
IF
PB
Triassic Anatectite

RC
Latest Triassic -
Jurassic arc
Amotape
Terrane Early Cretaceous arc
CL
Peru Early Cret. M-HP/LT

80°W 75°W

Fig. 1. Digital elevation model for northwestern South America showing the cordilleras, suspect terranes, main faults and the exposure of Triassic–Early 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: Ibagué Fault, OPF: Otu-
Pericos 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-Patía 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: Piñon 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 Gómez et al. (2007).

1994) along with undifferentiated Palaeozoic metamorphic rocks of the monzogranites and diorites (Tres Lagunas unit), and medium- to high-
Chiguinda and Agoyan units. The Loja Terrane is bound to the east by the grade, sillimanite and kyanite bearing orthogneisses and migmatites
Llanganates Fault, and to the west by the Baños fault. Triassic igneous and (Sabanilla Unit). Litherland et al. (1994), Noble et al. (1997), Riel
metamorphic lithologies are dominated by cordierite and garnet bearing 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 5

geochronological and geochemical data from the igneous and meta- Ecuador yield concordant zircon and monazite U-Pb dates ranging
morphic rocks in Ecuador (Table 1). The Piuntza unit consists of meta- between 207.6 ± 9.2 Ma and 247.2 ± 4.3 Ma (Figs. 2 and 3a; Table 1;
morphosed and skarnified siliciclastic rocks, tuffs and limestones that Litherland et al., 1994; Aspden et al., 1995; Chew et al., 2008; Riel
host Triassic bivalves (Litherland et al., 1994). The unit is exposed et al., 2013; Cochrane et al., 2014a). These dates overlap with concor-
beyond the structural limits of the Loja Terrane of Litherland et al. dant zircon U-Pb dates obtained from twenty six gneissic granites and
(1994) along the eastern flank of the southern Cordillera Real (Fig. 2), pegmatites exposed in the Cordillera Central, Sierra Nevada de Santa
where it is surrounded by the Jurassic Zamora Batholith although the Marta and the Guajira Peninsula in Colombia, which range between
nature of the contact is either unknown or unreported. 222 ± 10 Ma and 288.1 ± 4.5 Ma (Vinasco et al., 2006; Cardona et al.,
Widely dispersed and variably deformed Permian and Triassic 2010; Montes et al., 2010; Weber et al., 2010; Villagómez et al., 2011;
meta-granitoids, ultra-mafic–mafic rocks and metasedimentary rocks Restrepo et al., 2011; Cochrane et al., 2014a; Figs. 2 and 3a). A majority
occur within the northern Cordillera Central of Colombia (Fig. 2). of these crystalline rocks are Triassic, although six rocks from northern
These rocks were initially described by Hall et al. (1972), Feininger Colombia yield Permian ages. Ordonez-Carmona et al. (2001) report
et al. (1972) and González (1980), who considered them to be Permo- a Sm-Nd whole rock–garnet isochron date of 226 ± 17 Ma from a
Triassic on the basis of K/Ar dates. Restrepo and Toussaint (1988) granulite in the Tahamí Terrane, which they interpret as a cooling age
suggested that the Permo-Triassic rocks defined the basement of the following peak metamorphism. Multi-phase, plateau 40Ar/39Ar dates
fault-bounded Tahamí Terrane, and placed them within the Cordillera (Table 1) from Triassic granites and migmatites in Colombia and
Central Polymetamorphic Complex (Restrepo and Toussaint, 1982). Ecuador (Spikings et al., 2001; Vinasco et al., 2006; Cochrane et al.,
The Tahamí Terrane of Restrepo and Toussaint (1988) is bound by the 2014a) are younger than the U-Pb dates obtained from the same rocks,
Otú-Pericos fault to the east, which separates it from Grenvillian aged and Triassic dates span between 213.7 ± 0.9 Ma and 243 ± 4 Ma
metamorphic basement of the Chibcha Terrane (e.g. Ordoñez-Carmona (Table 1). The 40Ar/39Ar dates reflect the time of cooling of each rock
et al., 1999), and the San Jerónimo Fault to the west, which separates it through mineral-specific argon partial retention zones (i.e. 550–300 °C;
from the Quebradagrande Arc (Fig. 1). Maya and González (1995) and hornblende, muscovite and biotite) subsequent to crystallisation and
Villagómez et al. (2011) group the Triassic metamorphosed igneous metamorphic retrogression.
and sedimentary rocks into the Cajamarca Complex, which will be Concordant zircon U-Pb dates of amphibolites and a plagiogranite
adopted in this manuscript. The geological map of Colombia (Gómez from the Cordillera Real and Cordillera Central range between
et al., 2007) reveals a paucity of Triassic lithologies within the Cordillera 216.6 ± 0.4 Ma and 243 ± 4 Ma (Fig. 3d; Noble et al., 1997; Vinasco
Central south of the Ibagué Fault (Fig. 2). Vinasco et al. (2006), Martínez et al., 2006; Martínez, 2007; Cochrane et al., 2014a). Within Ecuador,
(2007), Cardona et al. (2010), Montes et al. (2010), Weber et al. (2010), these are exposed as dykes and sills (e.g. the Piedras and Monte Olivo
Restrepo et al. (2011), Villagómez et al. (2011) and Cochrane et al. units; Fig. 2), whereas they are more massive in the Tahamí Terrane of
(2014a) present a large quantity of geochemical data and concordant the Cordillera Central. The youngest of these ages was obtained from a
zircon U-Pb dates (Table 1, Fig. 2) from the Cajamarca Unit, confirming plagiogranite that formed by hydrothermal metamorphism of the
the Permo-Triassic crystallisation ages of the igneous rocks. Similar to Aburrá Ophiolite in northern Colombia, and thus is a minimum age for
the Cordillera Real in Ecuador, the Permo-Triassic rocks of the Cordillera the ophiolite (Martínez, 2007).
Central intrude and are faulted against Palaeozoic metamorphic rocks A majority of U-Pb dates of zircons extracted from meta-granites
such as the La Miel Unit (e.g. Restrepo et al., 1991; Villagómez et al., and migmatites range between 240 and 230 Ma (Fig. 3a, d), and a com-
2011), although these will not be considered further in this review. parison with latitude (Fig. 3d) does not reveal any trends for the Triassic
Martínez (2007) report a series of metagabbros and amphibolites in period, with the exception of a possible increase in age in far northern
the northern Cordillera Central, which they attribute to a Triassic Colombia within the Guajira Peninsula. Permian ages are restricted to
ophiolitic sequence, referred to as the Aburrá Ophiolite (Fig. 2). exposures in the Sierra Nevada de Santa Marta, and faulted blocks in
Permo-Triassic igneous and metamorphic rocks have also been the region of the Ibagué Fault (Fig. 2). This may reflect exposure, or per-
recognised in the Guajira Peninsula, Sierra Nevada de Santa Marta and haps approximate the primary distribution of Permian magmatic
at the base of boreholes drilled though the Plato-San Jorge Basin located intrusions.
north of the Cordillera Central (Cardona et al., 2010; Montes et al., 2010; Some concordant 206Pb/238U dates determined by in-situ methods
Weber et al., 2010; Fig. 1). Granites from the basement of the Plato-San were also obtained from the cores of zircon grains that were identified
Jorge Basin are mildly deformed, while the intrusions from the Sierra using cathodoluminescence. A frequency analysis of the distribution of
Nevada de Santa Marta are mylonitised. these older dates from the granitoids and migmatitic leucosomes yields
broad peaks at 420–580 Ma and 950–1200 Ma (Fig. 3b), which are the
4.2. Geochronology ages of protolith rocks and inherited grains. These age peaks are typical
of the distribution of dates obtained from detrital zircons from most
4.2.1. Cordillera Real of Ecuador and Cordillera Central of Colombia Palaeozoic terranes along western South America (e.g. Chew et al.,
Early attempts to date the Permo-Triassic crystalline rocks utilised 2007). The younger age group broadly corresponds with the age of the
the K/Ar and Rb/Sr methods (e.g. Feininger et al., 1972; Hall et al., Famatinian Arc, the Braziliano Orogeny and the timing of rifting during
1972 McCourt et al., 1984; Restrepo et al., 1991; Litherland et al., the fragmentation of Rodinia. The Famatinian arc (~ 510–415 Ma;
1994; Ordoñez and Pimentel, 2002), resulting in a large scatter of ages E.g. see zircon U-Pb ages presented in Pankhurst et al., 2000; Cardona
spanning between the Permian–Tertiary due to variable degrees of et al., 2007; Chew et al., 2007; Bahlburg et al., 2009; Mišković et al.,
daughter isotope loss. This review of geochronological work is restricted 2009; Villagómez et al., 2011; Van der Lelij, 2013) formed during the
to more accurate and peer-reviewed measurements of the crystallisation subduction of Pacific lithosphere beneath western South America sub-
ages of granitoids and mafic intrusions, which have been provided by sequent to the fragmentation of Rodinia, and has been recorded in
numerous concordant zircon and few monazite U-Pb dates (Table 1), Venezuela, Colombia, Peru and Argentina (see previous citations).
obtained using TIMS, SHRIMP and LA-ICPMS. Unless otherwise stated, Inherited zircons with U-Pb dates spanning the 450–650 Ma range
the LA-ICPMS and SHRIMP dates that are reported here were obtained also occur in Cretaceous and Tertiary sedimentary rocks of the Amazon
from the rims of zircons, and are considered to date either the most Foreland Basin in Ecuador (Martin-Gombojav and Winkler;, 2008),
recent phase of magmatic crystallisation or the most recent metamor- although intrusions of the Famatinian arc have not been recorded in
phic event that crystallised zircon. Ecuador, and within Colombia they are only recorded within the north-
Fourteen metagranites and migmatites of the Tres Lagunas and ern Cordillera Central (440–470 Ma; La Miel orthogneiss; Villagómez
Sabanilla units in the Cordillera Real and Amotape Complex of 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
6 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

Table 1
Summary of data collected from Permo-Triassic rocks of Ecuador and Colombia.
206
Sample Unit Lithology Latitude N-S d°m's'' Longitude W d°m's'' Pb/238U age ± 2σ (Ma) MSWD 40
Ar/39Ar age ± 2σ (Ma)

S-type granites and migmatitic leucosomes


Ecuador
Eastern Cordillera
09RC25 Tr. Lagunas metagranite S 1° 23' 51" 78° 21' 15" 233.7 ± 0.8 1.1
09RC31 Tr. Lagunas metagranite S 0° 22' 33" 78° 08' 32" 234.4 ± 0.9 0.8
09RC42 Sabanilla metagranite S 4° 27' 43" 79° 08' 52" 247.2 ± 4.3 3.0
09RC53 Tr. Lagunas metagranite S 3° 9' 24" 78° 48' 45" 231.0 ± 1.9 2.1
09RC44 Sabanilla paragneiss S 4° 29' 02" 79° 08' 55"
09RC45 Sabanilla Paragneiss S 3° 58' 41" 79° 01' 15"
09RC56 Tr. Lagunas metagranite S 1° 23' 57" 78° 22' 08" 235.0 ± 1.5 3.0
11RC03 Agoyan fm. metagranite N 0° 23' 24" 77° 51' 44" 207.6 ± 9.2 1.9
Tres Lagunas Tr. Lagunas granite 227.3 ± 2.2
Amotape Complex
09RC40 Moromoro migmatite S 3° 42' 16" 79° 51' 07" 237.7 ± 5.2 4.6 214.6 ± 0.9m
VI-08-12 La Bocana migmatite 226.0 ± 1.3⁎
PU-08-10 La Bocana migmatite S 3° 42' 58" 80° 03' 18" 223.2 ± 2.2⁎
AV-08-31 La Bocana migmatite 229.3 ± 2.4
AV-08-28d La Bocana migmatite S 3° 40' 41" 79° 54' 14" 225.7 ± 6.5
Moromoro Moromoro granite 227.5 ± 0.8⁎
Colombia
Central Cordillera
10RC04 Cajamarca metagranite N 4° 19' 24" 75° 12' 07" 277.6 ± 1.6 1.2
10RC40 Cajamarca metagranite N 5° 53' 13" 75° 25' 28" 236.1 ± 3.3 3.7 221.8 ± 1.0m
10RC41 Cajamarca metagranite N 6° 01' 08" 75° 07' 28" 234.1 ± 1.2 1.2
10RC42 Cajamarca metagranite N 5° 59' 17" 74° 55' 37" 244.6 ± 2.4 2.3
10RC43 Cajamarca metagranite N 5° 58' 34" 74° 54' 02" 245.0 ± 2.0 0.6 213.7 ± 0.9m
10RC53 Cajamarca metagranite N 7° 00' 56" 75° 22' 28" 236.4 ± 1.8 3.0
10RC66 Cajamarca qtz-Schist N 5° 08' 20" 75° 09' 47"
10RC69 Cajamarca metagranite N 5° 09' 27" 75° 07' 57" 255.7 ± 1.5 1.2
10RC71 Cajamarca pegmatite N 5° 07' 34" 74° 54' 38" 236.0 ± 0.6 0.9
DV65 Cajamarca metagranite N 5° 59' 16" 74° 55' 34" 240.9 ± 1.5 0.6
DV82 Cajamarca metagranite N 4° 17' 16" 75° 13' 59" 275.8 ± 1.5 3.0
DV02 Cajamarca paragneiss N 4° 46' 42" 74° 57' 54" 238–582
DV18 Cajamarca gneiss N 4° 28' 19" 75° 33' 18" 236.2 ± 6.3 0.6
DV19 Cajamarca quartzite N 4° 28' 19" 75° 33' 18" 231–1163
Abejorral Abejorral gneiss 250 ± 10#
Palmitas Palmitas gneiss 240 ± 4#
Amaga Amaga granite 227.6 ± 4.5 1.4
La Honda La Honda granite 218.7 ± 0.3b
El Buey El Buey granite 219.3 ± 0.3m
Manizales Manizales granite 229.7 ± 0.5h
GSI1 Santa Isabel gneiss N 6° 57' 34" 74° 45' 13" 226.7 ± 1.6 1.2
GN1 Nechi gneiss N 8° 10' 13" 74° 46' 55" 236.4 ± 6.6 2.1
PALM-1 Palmas migmatite N 6° 09' 14" 75° 32' 36" 222 ± 10#
Sierra Nevada de Santa Marta
A14 St. M. mylonite granite 288.1 ± 4.5 1.0
A48 St. M. mylonite granite 276.5 ± 5.1 1.8
EAM-12-05 St. M. mylonite granite 264.9 ± 4.0 0.0
Plato-San Jorge Basin
Cicuco-2a unknown granite N 9° 16' 25" 74° 38' 53" 241.6 ± 3.9 3.9
Cicuco-3 unknown granite N 9° 17' 39" 74° 38' 52" 241.6 ± 3.9 6.0
Lobita 1 unknown granite N 9° 18' 30" 74° 41' 31" 239.6 ± 2.9 0.6
Guajira Peninsular
AVO-03 Uray Gneiss gneiss 247.6 ± 4.1 0.5
AVO-06 Uray Gneiss gneiss 245.6 ± 3.9 0.5
Amphibolites
Ecuador
10RC28 Chinchina amphibolite N 5° 03' 05" 75° 34' 25" 224.7 ± 1.9 0.8
11RC04 Monte Olivo amphibolite N 0° 23' 24" 77° 51' 44"
11RC10 Monte Olivo amphibolite S 1° 23' 56" 78° 22' 52" 231.9 ± 3.2 1.6
11RC14 Piedras amphibolite S 3° 39' 9" 79° 50' 35" 222.7 ± 6.3 1.9
JR148 Piedras amphibolite nr nr 221 ± 17.0
Colombia
10RC39 Santa Elena amphibolite N 5° 54' 06" 75° 24' 31"
10RC39A Santa Elena amphibolite N 5° 53' 52" 75° 24' 37" 239.7 ± 2.4 1.9
10RC50 Tr. Intrusive amphibolite N 6° 09' 26" 75° 44' 31"
AC32B El Picacho plagiogranite 216.6 ± 0.4 0.7
CMK040A El Picacho meta-gabbro
Padua Padua amphibolite 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 (Villagómez 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 7

εHf zircon ± 2σ εNd w.r ± 2σ (87Sr/86Sr)i wr ± 2σ (206Pb/204Pb)i wr ± 2σ δ18O (‰) ± 2σ Th/U zircon ± 2σ A/CNK wr (La/Yb)n wr Publication

10.5 to −3.2 15.3 ± 0.2 0.26 ± 0.1 1.99 13.24 Cochrane et al. (2014a)
−11.0 to +3.2 15.1 ± 0.2 0.04 ± 0.1 1.40 13.50 Cochrane et al. (2014a)
−5.3 to −0.5 16.8 ± 0.2 0.69 ± 0.5 1.23 10.58 Cochrane et al. (2014a)
−2.63 ± 0.43 0.24 ± 0.1 1.19 12.68 Cochrane et al. (2014a)
1.37 4.65 Cochrane et al. (2014a)
Cochrane et al. (2014a)
−6.0 to +1.7 12.1 ± 0.2 0.14 ± 0.1 2.24 6.92 Cochrane et al. (2014a)
−16.3 to −9.0 15.1 ± 0.2 0.01 ± 0.0 Cochrane et al. (2014a)
Litherland et al. (1994)

−7.5 to +0.8 0.42 ± 0.5 2.38 11.36 Cochrane (2013)


Riel et al. (2013)
1.85 Riel et al. (2013)
0.13 Riel et al. (2013)
0.10 1.50 Riel et al. (2013)
Aspden et al. (1995)

1.96 ± 0.31 13.6 ± 0.2 1.27 ± 0.6 1.18 16.23 Cochrane et al. (2014a)
−6.57 ± 0.66 17.4 ± 0.2 0.08 ± 0.1 1.73 8.19 Cochrane et al. (2014a)
−9.5 to −0.2 13.1 ± 0.2 0.23 ± 0.1 1.27 11.49 Cochrane et al. (2014a)
−8.2 to +1.4 13.1 ± 0.2 0.35 ± 0.1 1.33 12.00 Cochrane et al. (2014a)
−11.7 to −3.1 0.42 ± 0.4 1.36 15.70 Cochrane et al. (2014a)
−5.9 to +3.1 15.9 ± 0.2 0.30 ± 0.2 1.56 14.27 Cochrane et al. (2014a)
1.84 12.63 Cochrane et al. (2014a)
−3.16 ± 0.7 15.6 ± 0.2 1.10 ± 0.2 1.70 12.81 Cochrane et al. (2014a)
−6.0 to +0.4 0.31 ± 0.1 Cochrane et al. (2014a)
−5.9 to +0.7 0.26 ± 0.2 Cochrane et al. (2014a)
−3.7 to +0.3 0.66 ± 0.1 Cochrane et al. (2014a)
Villagómez et al. (2011)
Villagómez et al. (2011)
Villagómez et al. (2011)
0.82 Vinasco et al. (2006)
0.25 Vinasco et al. (2006)
0.30 Vinasco et al. (2006)
Vinasco et al. (2006)
Vinasco et al. (2006)
Vinasco et al. (2006)
0.19 Restrepo et al. (2011)
0.23 Restrepo et al. (2011)
0.24 Restrepo et al. (2011)

0.73 Cardona et al. (2010)


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

nr Montes et al. (2010)


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

0.20 Weber et al. (2010)


0.59 Weber et al. (2010)

13.31 ± 0.25 9.83 0.70354 17.520938 0.20 ± 0.1 0.66 1.41 Cochrane et al. (2014a)
0.61 2.59 Cochrane et al. (2014a)
−6.3 to +11.2 5.03 0.71470 18.707878 0.19 ± 0.1 0.63 1.71 Cochrane et al. (2014a)
15.00 ± 0.29 9.79 0.70271 17.754038 0.32 ± 0.2 0.61 0.81 Cochrane et al. (2014a)
0.52 Noble et al. (1997)

8.98 0.70430 18.119529 0.82 2.34 Cochrane et al. (2014a)


−4.8 to +10.0 4.13 0.70535 18.298843 0.62 2.02 Cochrane et al. (2014a)
10.18 0.70243 16.607997 0.50 0.49 Cochrane et al. (2014a)
3.4 0.70448 0.97 8.00 Martínez (2007)
8.4 0.61 0.64 Martínez (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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
8
Caribbean Large Igneous Province (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the

Table 2
Summary of geochronological data collected from Jurassic and Early Cretaceous rocks of the cordilleras of Ecuador and Colombia.
206
Sample Unit Lithology Latitude N-S Longitude W Pb/238U MSWD 40
Ar/39Ar Lu-Hf εHf zircon ± 2σ εNdi w.r ± 2σ (87Sr/86Sr)i Th/U A/ (La/Yb) Publication
d°m's'' d°m's'' age ± 2σ (Ma) (U/Pb) age ± 2σ (Ma) age ± 2σ (MSWD) wr ± 2σ zircon CNK n wr
(Ma), wr
(MSWD)

Jurassic intrusions and volcanic rocks


Ecuador
Eastern Cordillera
09RC57 Azafrán Granodiorite S01° 14' 37" 78° 09' 59" 143.5 ± 1.3 2.2 9.25 ± 0.54(3.6) 5.04 1.07 18.05 Cochrane (2013)
09RC59 Azafrán Monzogranite S01° 14' 37" 78° 09' 27" 140.7 ± 0.7 1.6 8.70 ± 0.37(2.1) 4.61 1.08 22.21 Cochrane (2013)
09RC65 Rosa Florida Quartz diorite N00° 15' 4.0" 77° 19' 23" 182.4 ± 0.6 0.7 2.84 ± 0.24(1.5) 0.44 0.9 14.79 Cochrane (2013)
09RC22 Abitagua Granite S01° 14' 47" 78° 06' 48" 174.0 ± 1.2 1.8 4.74 ± 0.26(0.9) 1.78 1.09 18.93 Cochrane (2013)
09RC60 Abitagua Granodiorite S00° 24' 33" 77° 28' 51" 173.0 ± 1.3 2.8 4.02 ± 0.60(4.1) 1.27 1.07 32.17 Cochrane (2013)
09RC61 Abitagua Monzogranite S00° 22' 19" 77° 29' 45" 169.8 ± 1.1 2 4.89 ± 0.23(1.6) 1.9 1.11 19.59 Cochrane (2013)
09RC43 Zamora Granodiorite S04° 24' 0.0" 79° 04' 49" 131.6 ± 1.1 0.4 6.25 ± 0.88(10) 3.28 1.75 16.82 Cochrane (2013)

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


09RC46 Zamora Granodiorite S03° 34' 43" 78° 30' 59" 178.1 ± 1.4 3 6.23 ± 0.23(0.8) 2.84 1.06 19.57 Cochrane (2013)
101E Zamora Granodiorite S03° 12' 15" 78° 24' 05" 160.5 ± 1.7h Chiaradia
et al. (2009)
138 Zamora Granodiorite S03° 11' 18" 78° 25' 08" 153.8 ± 1.5h Chiaradia
et al. (2009)
DTR64/ Porphyry Qtz diorite S04° 01' 28" 78° 47' 25" 145.4 ± 0.21 0.46 Chiaradia
69/184 et al. (2009)
MI90 Misahualli Andesite 172.3 ± 2.1h 0.88 16.99 Romeuf et al.
(1995)
Colombia
Central Cordillera
DV04 Ibagué Diorite N04° 47' 00.2" 74° 58' 31.4" 159.2 ± 5.2h Villagómez et al.
(2011)
h
DV05 Ibagué Granodiorite N04° 24' 27.7" 75° 16' 05.3" 166.0 ± 10.0 0.29 153.1 ± 2.0 Villagómez et al.
(2011)
DV06 Ibagué Granite N04° 24' 08.9" 75° 17' 40.3" 182.6 ± 2.4h Villagómez et al.
(2011)
DV07 Ibagué Granite N0°4 24' 25.4" 75° 18' 04.5" 148.9 ± 3.3h Villagómez et al.
(2011)
DV09 Ibagué Granite N04° 24' 29.7" 75° 18' 11.8" 159.6 ± 2.4 0.63 151.8 ± 0.9b Villagómez et al.
(2011)
DV129 Unnamed Granodiorite N01° 10' 02.5" 76° 51' 32.7" 175.8 ± 1.7 1.3 −1.04 ± 0.20(1.6) Cochrane (2013)
DV132 Saldana Agglomerate N01° 07' 55.5" 76° 50' 54.4" 179.0 ± 2.0 1.6 −2.90 ± 1.70(48) Cochrane (2013)
DV137 Unnamed Granite N01° 04' 55.4" 76° 48' 34.1" 173.6 ± 1.5 2.5 −1.86 ± 0.51(3.5) Cochrane (2013)
DV138 Saldana Rhyolite N01° 06' 45.0" 76° 50' 18.6" 181.5 ± 1.6 3 −0.01 ± 0.43(2.3) Cochrane (2013)
10RC02 Ibagué Granite N04° 10' 39" 75° 07' 19" 164.4 ± 1.1 0.8 4.47 ± 0.26(1.1) 1.60 1.07 23.28 Cochrane (2013)
10RC06 Ibagué Granite N04° 28' 12" 74° 35' 07" 168.8 ± 0.7 2.5 2.68 ± 0.19(1.5) 0.32 1.07 Cochrane (2013)
10RC08 Ibagué Granodiorite N04° 14' 48" 75° 09' 47" 156.5 ± 1.1 0.5 7.65 ± 0.24(1.0) 3.86 1.04 22.47 Cochrane (2013)
10RC10 Ibagué Granite N04° 14' 31" 75° 10' 09" 155.7 ± 2.2 3.4 7.44 ± 0.23(1.4) 3.71 1.03 28.45 Cochrane (2013)
10RC78 Segovia Monzogranite N06° 17' 53" 74° 18' 0" 188.9 ± 2.0 1.5 −2.94 ± 0.31(1.5) −3.69 1.00 8.77 Cochrane (2013)
10RC03 Saldana Rhyodacite N04° 10' 39" 75° 07' 19" 158.5 ± 1.0 1.9 3.76 ± 0.45(2.9) 1.12 1.04 18.35 Cochrane (2013)
10RC07 Saldana Qtz porphyry N04° 28' 12" 74° 35' 06" 146.8 ± 1.5 0.8 8.19 ± 0.24(1.3) 4.24 1.11 10.88 Cochrane (2013)
CB0007A Las Minas Monzodiorite N02° 14' 37.5" 75° 48' 22.2" 187.4 ± 2.3 0.62 3.32 0.89 12.66 Bustamante et al.
(2010)
CB0010 Ibagué Granite N02° 24' 04.2" 75° 53' 46.8" 189.1 ± 2.9 0.59 4.32 0.87 13.35 Bustamante et al.
(2010)
Garzón Massif
CB0001 Garzón Granite N02° 11' 28.8" 75° 35' 19.2" 173.9 ± 1.6 1.5 5.32 Bustamante et al.
(2010)
CB0005 Altamira Monzogranite N02° 02' 24.9" 75° 45' 34.8" 179.0 ± 2.2 0.99 2.32 0.91 30.11 Bustamante et al.
(2010)
Santander Massif
Caribbean Large Igneous Province (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the

10VDL05 Corcova 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 Van der Lelij
(2013)
10VDL22 Pescadero Granodiorite N06° 49' 48" 72° 59' 27" 199.1 ± 1.3 2.5 −4.72 ± 0.21 −4.44 ± 0.16 0.7082 0.88 1.06 27.25 Van der Lelij
(2013)
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 Van der Lelij
(2013)
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 Van der Lelij
(2013)
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 Van der Lelij
(2013)
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 Van der Lelij
(2013)
10VDL39 Pegmatite N07° 10' 46" 72° 59' 48" 208.8 ± 1.2 2.2 −6.55 ± 0.85 −13.06 ± 3.55 0.7216 0.03 1.05 22.62 Van der Lelij
(2013)
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 Van der Lelij
(2013)
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 Van der Lelij
(2013)
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 Van der Lelij
(2013)
10VDL61 Rio Surata Granodiorite N07° 09' 59" 73° 05' 17" 200.0 ± 1.5 1.2 −5.48 ± 0.55 −7.20 ± 0.09 0.7167 1.13 1.13 15.15 Van der Lelij
(2013)

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


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 Cochrane (2013)
11RC13 Upano Greenschist S03° 14' 54" 78° 41' 39" 121.0 ± 0.8 1.4 0.84 Cochrane (2013)
09RC63 Chinguál Monzogranite N 00° 22' 51" 77° 17' 39" 125.3 ± 0.9 2.8 8.15 ± 0.45(3.0) 4.24 1.05 11.2 Cochrane (2013)
SEC16-1 Raspas Blueschist S03° 35' 40.6" 79° 55' 30.5" 126.4 ± 4.0 0.81 7.85 John et al. (2010)
(4.1)
SEC47-4 Raspas Metapelite 129.9 ± 5.6 John et al. (2010)
(2.0)
P
Raspas Metapelite 132 ± 5 Feininger (1980)
Raspas Metapelite 123.9 ± 1.4P Gabriele
(2002)
Raspas Metapelite 129.3 ± 1.3P Gabriele (2002)
09PR47 Peltetec Metabasalt 134.3 ± 12.8Pl 1.14 0.48 2.87 This study
04PR48 Peltetec Gabbro 134.7 ± 0.9Pl 0.56 3.12 This study
Colombia
Central Cordillera
DV176 Quebradagrande Diorite N05° 27 '16.0'' 75° 28' 28.2'' 112.9 ± 0.8 2.3 12.11 ± 0.17(1.4) 1.13 32.81 Cochrane (2013)
DV20 Quebradagrande Tuff N04° 29' 27.8" 75° 34' 02.0" 114.3 ± 3.8 2.00 Villagómez et al.
(2011)
10RC27 Quebradagrande Sandstone N05° 03' 13.3" 75° 34' 04.6" 149.2 ± 6.1max Cochrane (2013)
DV89b Arquía Amphibolite N04° 16' 24.7" 75° 47' 22.2" 112.0 ± 3.7h Villagómez et al.
(2011)
190B Barragán Schist 120.7 ± 0.3m Bustamante et al.
(2012)
Eastern Cordillera
Pa5 Pacho Gabbro N05° 32' 34.4" 74° 09' 10.7" 136.0 ± 0.4h 3 0.70621 Vásques et al.
(2010)
h
Pj7 Pajarito Gabbro N05° 18' 29.3" 72° 42' 7.7" 120.5 ± 0.6 5.4 0.70413 Vásques 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 (Villagómez 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.

9
10 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

2010; Van der Lelij, 2013). Field relationships (Litherland et al., 1994) within the high-K calc-alkaline and calc-alkaline fields when comparing
suggest that the mesosomal rocks of the Triassic migmatites and SiO2 with K2O (Fig. 5a). The Triassic anatectites have strongly pera-
S-type granites within Ecuador are considered to be sedimentary rocks luminous Aluminium Saturation Indices (ASI 0.97–2.38; calculated
of the Palaeozoic, fossil bearing Chiguinda and Isimanchi units of the using Maniar and Piccoli, 1989; Fig. 4b), while the Permian granitoids
Cordillera Real (Fig. 2). These sparsely studied sequences yield a detrital tend to cluster at slightly lower peraluminous and mildly metaluminous
zircon U-Pb age spectrum that has the same age peaks (Fig. 3c; Chew values (ASI 0.92–1.73, with a majority b 1.1). The Triassic anatectites
et al., 2008), although the tectonic setting within which these sequences yield elevated δ18O quartz (Fig. 4c), which along with their high ASI in-
were deposited is undetermined. The Brasiliano metamorphic belts dices places these granites and leucosomes within the “S-type granite”
(Cordani et al., 2003) formed during the late Neoproterozoic amalgam- field of Chappell and White (1974) and Harris et al. (1997). N-MORB
ation of Gondwana, and may have supplied some detritus to western normalized trace element abundances of Triassic granitoids from
South America. However, these belts are located in eastern South Ecuador and Colombia are identical (Fig. 4d), suggesting that there are
America, and a lack of evidence for detritus being sourced from the no significant along-strike changes in the fractionation and assimilation
intervening Amazonia Craton suggests that the Brasiliano Orogenic history of these high-SiO2 melts. The trace elements are enriched in
belts were not a major source region (Chew et al., 2008). Finally, Light Ion Lithophile Elements (LILE), and negative Nb and Ta anomalies
all magmatism associated with Neoproterozoic extension is mafic are present in both the Permian and Triassic granitoids (Fig. 4e, f), sug-
(e.g. the Puncoviscana fold belt in northwestern Argentina; Omarini gesting that a subduction-derived component was incorporated into
et al., 1999), which led Chew et al. (2008) to suggest that it is unlikely these rocks. The Triassic granitoids yield slight negative Ba, Eu, Sr and
that these rocks were a major contributor of zircons to Palaeozoic Ti anomalies, which suggest that plagioclase and Fe-Ti oxides have frac-
sequences along western South America. tionated, and a positive Pb anomaly that may be derived from a
protolith within the continental crust. In contrast, the Permian granites
4.2.2. Comparison with the ages of Permian and Triassic rocks in Venezuela do not yield Ba, Eu and Sr anomalies, although they do have negative Ti
and Peru anomalies, suggesting that they evolved via a different fractionation
Van der Lelij (2013) report concordant zircon U-Pb dates (LA- scheme. Trace element concentrations normalized to the composition
ICPMS) from four granitoid intrusions and a dacitic lava from the Merida of average upper continental crust (Taylor and McLennan, 1995) plot
Andes of Venezuela (Fig. 1). These dates range between 202.0 ± 1.8 Ma close to unity, corroborating the S-type character of these rocks
(La Quinta Fm.) and 243.5 ± 3.4 Ma, and overlap with dates obtained (Fig. 4f). REE abundances in the Triassic granites and leucosomes nor-
from Colombia and Ecuador (Fig. 3d). No Permian concordant zircon malized to N-MORB reveal light-REE enrichment with (La/Yb)n ranging
U-Pb dates have been reported from the Merida Andes. Rhyolites and between 2.3 and 19.8, with a mildly positive correlation with 206Pb-238U
granites of the El Baul massif in Venezuela yield zircon U-Pb dates that crystallisation age (Fig. 4h). (La/Yb)n ratios from Permian granites yield
span between 283.3 ± 2.5 Ma and 291.1 ± 3.1 Ma (Viscarret et al., a larger range of 8.8–55.1, and the REE concentrations have a larger
2009), and a zircon U-Pb age of 272.2 ± 2.6 was obtained from a granitic range relative to N-MORB, compared to the Triassic rocks (Fig. 4g).
intrusion in the Paragauana Peninsula (Van der Lelij, 2013).
Voluminous, partly migmatised Late Permian–Triassic magmatic
intrusions are exposed throughout the southern and central Eastern 4.3.2. Comparison with Permian and Triassic rocks in Venezuela and Peru
Cordillera of Peru (Mišković et al., 2009). Zircon U-Pb dates range Migmatised granitoids within the Eastern Cordillera of Peru that
between 223 and 285 Ma (Fig. 3d; Mišković et al., 2009; Reitsma, 2012), crystallised during 285–223 Ma are high-SiO2 granites, mildly meta-
with a peak at 240–260 Ma. The crystallisation ages show a southward luminous to peraluminous (ASI 0.9–1.1; Fig. 4b), and yield K2O/Na2O
younging trend, and the oldest plutons south of 11.5°S are younger than ratios that mainly range between 0.8 and 1.2 (Mišković et al., 2009).
245 Ma. The Mitu Group of the central and southern Eastern Cordillera Mišković et al. (2009) report U-Pb zircon ages from monzogranites
of Peru hosts abundant Triassic sedimentary and volcanic sequences. in southern Peru (Cordillera de Carabaya) which range between 190
Volcanic tuffs in the south yield concordant zircon U-Pb (LA-ICPMS) and 216 Ma. These intrusions are geochemically distinct from the
dates ranging between 234.3 ± 0.3 Ma and 238.7 ± 1.8 Ma (Mišković older Permo-Triassic group (223–285 Ma) because they are strongly
et al., 2009; Reitsma, 2012), and Chew et al. (2005) report a zircon peraluminous (ASI 0.98–1.42; Fig. 4b), and yield anomalously high
U-Pb age of 219.7 ± 1.8 Ma from a rhyolite of the Mitu Group in central whole rock K2O/Na2O ratios (0.55–2.22, with a majority N 1.20). Geo-
Peru. Detrital zircons extracted from oxidised terrigeneous sedimentary chemically, the Permian and Triassic migmatites and granites of the
rocks of the Mitu Group yield minimum crystallisation dates ranging Cordillera Real of Ecuador and the Cordillera Central of Colombia
between 217.2 ± 4.1 Ma and 250.7 ± 4.9 Ma, which constrain their (K2O/Na2O 0.77–2.93; ASI 0.92–2.38) resemble the Late Triassic granites
maximum stratigraphic ages (Reitsma, 2012). Several authors propose of the Eastern Cordillera of Peru. Mišković et al. (2009) combined these
that the Mitu Group was deposited within a rift (Mégard, 1978; major element characteristics with iron oxide number, SiO2 (e.g. Frost
Laubacher et al., 1988; Reitsma, 2012), and Reitsma (2012) suggests et al., 2001) and trace element abundances, to classify the Permo-
that the rift formed in a back-arc basin setting. Triassic (285–223 Ma) and Late Triassic plutons as late- to post-
Romero et al. (2013) recently published a concordant zircon U-Pb orogenic.
age of 243 ± 0.1 Ma from a basalt exposed in Macabí Island offshore The shift from a well characterised, calc-alkaline Carboniferous arc in
northern Peru (~8°S). Peru to Permian post-tectonic alkali feldspar granites (Mišković et al.,
2009) and ultimately alkaline bimodal volcanic rocks of the Mitu
4.3. Geochemistry of the granites and migmatites Group is characteristic of lithospheric thinning (e.g. Xu et al., 2007).
Mišković et al. (2009) suggest that the Permo-Triassic granitoids
4.3.1. Cordillera Real of Ecuador and Cordillera Central of Colombia (223–285 Ma) formed by dehydration melting of the lower crust during
Major oxide, trace element and Rare Earth Element (REE) abun- basaltic underplating (e.g. Sisson et al., 2005), which was driven by de-
dances and oxygen isotope compositions (Table 1) have been obtained compression subsequent to break-off of the Carboniferous slab. Granit-
from Permian and Triassic granites and migmatitic leucosomes from oid intrusions and rift-related magmatism (Dalmayrac et al., 1980)
Colombia (Vinasco et al., 2006; Martínez, 2007; Cardona et al., 2010) initiated in the Middle Triassic, forming the Mitu Rift and bimodal volca-
and Ecuador (Litherland et al., 1994; Cochrane et al., 2014a). These nic rocks of the Mitu Group (Reitsma, 2012). Finally, the highly
rocks span the boundaries of calcic and alkali-calcic differentiation peraluminous Late Triassic granites formed by melting of the fertile up-
trends on the modified alkali-lime index of Peacock (1931; Fig. 4a), permost crust, consisting of an igneous protolith and a substantial sed-
with a compositional range of 62–78 wt% SiO2. The same rocks plot imentary 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 11

4.4. Geochemistry of the amphibolites from multiple crustal sources, while some variation found in the
xenocrystic bearing zircons may be due to the fractionation of Hf iso-
A comparison of the abundance of K2O and SiO2 (Fig. 5a) in all of the topes between cores and overgrowths during melting (e.g. Gerdes and
magmatic rocks that yield Triassic ages reveals the bimodal nature of Zeh, 2009). In contrast, a small proportion of granites yield εHfi (zircon)
magmatism within northwestern South America as part of Pangaea be- values that statistically define a single population, suggesting that they
tween 216.6 ± 0.4 Ma and 243 ± 4 Ma (Fig. 3d). Amphibolitic dykes and were derived from a distinct, homogeneous source. No correlation is
massive metagabbros from the Cordillera Real and Amotape Complex of found between the εHfi values obtained from the rims of xenocryst-
Ecuador (Chinchina, Monte Olivo and Piedras units) and the Cordillera bearing and xenocryst- free zircons, and crystallisation age, which is
Central of Colombia (Santa Elena, Padua and Aburrá) yield low K2O not surprising given the heterogeneous nature of the source rocks with-
(b 0.5 wt%) relative to SiO2 (46–55 wt%), placing them within the tho- in the crust. εHfi values obtained from xenocrystic zircon cores (Fig. 6a)
leiitic field of Peccerillo and Taylor (1976; Fig. 5a). However, a compar- span a large range and are representative of the sedimentary protoliths
ison of the immobile elements Th and Co (Fig. 5b) suggests that that melted to form the anatectites. The large heterogeneity in εHfi cor-
amphibolitic dykes straddle the tholeiite and calc-alkaline fields, while roborates the large range in 238U/206Pb dates of the protoliths.
the massive metagabbros of the Aburrá Ophiolite plot in the tholeiite
field. These discrepancies suggest that the amphibolitic dykes may be 4.5.2. Zircon Hf isotope geochemistry of the amphibolites
partially altered, although this was not visible in hand specimen or in Cochrane et al. (2014a) report εHfi (zircon) values from four am-
thin sections. The amphibolites and metagabbros have values of Zr/ phibolites which show a negative correlation with crystallisation age
TiO2 that are lower than 0.01 (Fig. 5c), placing them within the sub- (Fig. 6b). The two older amphibolitic dykes (240–232 Ma) yield a
alkaline basalt field, implying that the tholeiitic nature is primary. The large range in εHfi, with juvenile values (7.4 to 10) obtained from
amphibolitic dykes are enriched in Ti relative to V (Fig. 5d), and plot patchy or unzoned (cathodoluminescence) zircons, and crustal compo-
in the MORB or back arc basin basalt (BABB) field of Shervais (1982). sitions (−3.6 to −4.8) from zircons that exhibit oscillatory zoning, sim-
However, the massive metagabbros of the Aburrá Ophiolite (El Picacho ilar to the zircons extracted from the anatectites (Fig. 6c). The two
metagabbros) that yield very low K2O abundances of 0.02–0.12 wt% youngest amphibolites (225–223 Ma) define single populations of εHfi
(Fig. 5a), plot closer to the arc field. LILE abundances within the (13.3–15.0) from unzoned zircons, which approach the depleted mantle
amphibolitic dykes from Colombia and Ecuador (Fig. 5e) are enriched array. Crustal contamination of the mafic melts during emplacement
(up to ~100 times) relative to N-MORB and lack significant Nb and Ta was an important process in the petrogenesis of the older amphibolites
anomalies. The HFSE elements plot close to parity with N-MORB, prior to ~225 Ma. However, there is no evidence for the assimilation of
which is consistent with the tectonic discrimination plots. Similarly, significant continental crust after ~225 Ma. This interpretation is sup-
N-MORB normalised REE plots (Fig. 5f) for the amphibolitic dykes plot ported by a negative correlation between εHfi (zircon) and (La/Yb)n,
close to MORB compositions, although the LREE are slightly enriched (Fig. 6b) suggesting that the mafic dykes trend towards MORB compo-
with (La/Yb)n ratios varying between 0.59 and 3.16, while the massive sitions with a depleted mantle source, during 240–223 Ma.
metagabbros of the Aburrá Ophiolite yield approximately flat REE pat-
terns that are slightly depleted relative to N-MORB. N-MORB normal-
ised La/Yb ratios from all of these rock sequences show a progressive 4.5.3. Comparison with Zircon Hf isotope compositions in Peru
reduction with crystallisation age from 243 ± 4 Ma to 216.6 ± 0.4 Ma Hf isotopic compositions obtained from Permian and Triassic
(Fig. 5g). Finally, whole rock εNdi values for the amphibolitic dykes peraluminous granitoids of the Eastern Cordillera of Peru yield no
and the metagabbros range between 3.40 and 10.18, and become clear trends with time, and εHft values range between 6 and − 8
more juvenile with younger crystallisation ages (Fig. 5h). The most ju- (Fig. 6d; Mišković and Schaltegger, 2009). This range overlaps with
venile rocks are characteristic of MORB and BABB isotopic compositions. that obtained from Permian and Triassic granitoids of Ecuador and
A single amphibolite from the Monte Olivo unit in the Cordillera Real Colombia, and individual, intra-sample spot analyses show a large
of Ecuador yields a whole-rock 87Sr/86Sri of 0.7147 (Fig. 5i), which is ex- spread, reflecting the heterogeneity of the source rocks. Mean zircon
tremely high relative to its 143Nd/144Ndi of 0.5126 and low La/Yb ratio of εHfi values from the Early and Late Triassic of 0.02 ± 1.56 and
1.71 (Cochrane et al., 2014a). This is consistent with low temperature −1.96 ± 1.56 suggest that this period was characterised by the addi-
alteration, which has preferentially mobilized the LILE but had a mini- tion of isotopically juvenile, mantle derived magmas, which were un-
mal effect on the REE. derplating previously attenuated continental crust (Mišković and
Schaltegger, 2009). Calculated crustal Hf model ages for the granitoids
4.5. Zircon Hf isotope geochemistry within the Eastern Cordillera of Peru range between 1.4 and 1.0 Ga
(Mišković and Schaltegger, 2009), suggesting that the source rocks
Cochrane et al. (2014a) report Hf isotopic compositions from zircons were basement that formed within the Sunsas Orogeny during the
extracted from eighteen migmatitic leucosomes, more massive granit- amalgamation of Rodinia.
oids and amphibolitic dykes throughout the Cordillera Real of Ecuador
and the Cordillera Central of Colombia (Table 1). The zircons, which 4.6. Thermal histories during the Triassic
have been dated by LA-ICP-MS (U-Pb), yield a large range of weighted
mean εHfi values of +15 and −20 (Fig. 6), which are consistent with Numerous K/Ar dates have been obtained from Palaeozoic and
crustal recycling and the addition of new continental crust (Collins Triassic magmatic rocks in the Cordillera Real of Ecuador (Litherland
et al., 2011; Cochrane et al., 2014a). 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).
4.5.1. Zircon Hf isotope geochemistry of the granites and However, these dates cannot be unambiguously interpreted as
migmatitic leucosomes crystallisation ages given the potential for isotopic disturbance by
Single leucosomes of migmatites and peraluminous granites gener- thermally activated diffusive loss and fluid assisted loss of the daughter
ally yield high, intra-sample variations (e.g. εHfi + 3 to − 11; Fig. 6a) isotopes. Within Ecuador, muscovite, biotite and whole rock K/Ar dates
within coeval magmatic rims that surround variably aged xenocrystic of the Tres Lagunas Granite and Sabanilla unit (migmatites) range
cores (Cochrane et al., 2014a), and within samples that lack older between 100 and 50 Ma (Litherland et al., 1994), revealing a clear
cores. These variations are too large for magmatic zircons that disturbance to the isotopic system. Unfortunately, we cannot extract
crystallised from a single, well-mixed source (e.g. Gerdes et al., 2002), useful time–Temperature (t–T) information from these data because
and Cochrane et al. (2014a) conclude that they are mainly derived the degree of daughter isotope loss cannot be quantified. 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
12 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

76° 75° 74°


Jurassic
Metalluminous granitoids 236.4±6.6
(mainly granodiorite).
Continental arc intrusions 8°N
Triassic OPF
Peraluminous granites and migmatites
236.4±1.8
(e.g. Tres Lagunas Granite, Sabanilla Aburrá Ophiolite
Migmatite) 216.6±0.4 226.7±1.9
Amphibolites and ultramafic rocks 234.1±1.2
222±10*
(e.g. Piedras unit, Monte Olivo unit) Santa Elena
240±4* Amphibolites
244.6±2.4
Palaeozoic - Triassic 245.0±2.0*
227.6±4.5
Undifferentiated para- and 218.7±0.3 213.7±0.9
M
ortho-, schists and gneisses 219.3±0.3 240.9±1.5
(Ecuador: Agoyán, Chiguinda, Piuntza
239.7±2.4 227.6±4.5
units.
236.1±3.3 243±4

SJF
Colombia: Cajamara Unit)
221.8±1.0 255.7±1.5
U-Pb zircon, monazite 250±10*
244.6±2.4 236.0±0.6
LA-ICPMS, intrusions 238 - 2800
U-Pb zircon 224.7±1.9 I 220 - 600
277.3±3.0 IF
SHRIMP, intrusions Chinchina Stock Cajamarca Amphibolites
U-Pb zircon, LA-ICPMS, 236.2±6.3 229.7±0.5
233.7±4.8 - 2.6 Ga detrital zircons,
4°N
220 - 1200
metasedimentary rocks CAF
40Ar/39Ar plateau date
213.7±0.9 hornblende, muscovite,
biotite, intrusions 277.6±1.6 275.8±1.5
Also dated by apatite
247.2±4.3*
U-Pb for t-T analysis
78°W
City
2°N
Colombia
0 100 km
Monte Olivo
Amphibolites
1°N P ed
lat
re
ue
ag
Ib

207.6±9.2

Tres Lagunas
Granite PF 0°
Ecuador
234.4±0.9 Q
235.0±1.5

79°W
233.7±0.8 227.3±2.2

231.9±3.2
LF

2°S
Moromoro Migmatite BF
& Piedras Amphibolite

80°W 231.0±1.9
222.7±6.3 PF
3°S
221±18
223-229
237.7±5.2 Piuntza
214.6±0.9 Unit
227.5±0.8 4°S
L
247.2±4.3* Sabanilla
Migmatite
234±5 - 2600 Zumba
Ophiolite
5°S

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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 13

McCourt et al. (1984) and Litherland et al. (1994) interpret these data to Plate reconstructions of western Pangaea during the Permian to
reflect continental collision at ~120 and 65–55 Ma. Early Triassic (Elías-Herrera and Ortega-Gutiérrez, 2002; Weber et al.,
Spikings et al. (2000, 2001) and Villagómez and Spikings (2013) 2007) juxtapose Acatlán, Oaxaquia and the Chortis Block against
present 40Ar/39Ar (white mica, biotite, alkali feldspar) and fission track north-western South America (Fig. 9). At the present time, Oaxaquia is
(zircon, apatite) data from Triassic magmatic rocks in Ecuador and considered to underlie central and southern Mexico, including the
Colombia, which were used to construct t–T paths. However, the colli- Maya Block. The Maya Block hosts extensive, undeformed Permian
sion of the Caribbean Large Igneous Province with South America at granites, and deformed and foliated Permian granitic gneisses and
~75 Ma drove more than 350 °C of cooling, and Triassic thermal histo- migmatites (Chiapas Massif; e.g. Solari et al., 2008), which underwent
ries were not preserved in the isotopic systems. rock uplift and erosion during the early Triassic (Schaaf et al., 2002).
Recently, Cochrane et al. (2014b) published apatite U-Pb data from a The undeformed and deformed Permian granites yield concordant U-
Triassic leucosome of the Sabanilla Unit (migmatite 09RC42; Table 1) of Pb (zircon) dates ranging between 289 and 255 Ma (Yanez et al.,
southern Ecuador. The authors demonstrate that Pb was lost from the 1991; Solari et al., 2001; Elías-Herrera and Ortega-Gutiérrez, 2002;
apatite grains by thermally activated diffusion, and thus the dates can Ducea et al., 2004; Weber et al., 2007; Kirsch et al., 2012;
be combined with grain sizes and the diffusion properties of Pb in apa- Ortega-Obregon et al., 2013; Kirsch et al., 2014) and are considered to
tite to generate a series of plausible t–T paths (Fig. 7a) using a computed form part of a continental arc (e.g. Torres et al., 1999). We suggest
Monte Carlo algorithm, at temperatures N350 °C. Those paths reveal that the Permian, peraluminous granitoids exposed within north-
rapid cooling subsequent to anatexis at ~250 Ma. The leucosome subse- western South America formed within the same tectonic regime, and
quently remained at temperatures lower than the Pb Partial Retention are a continuation of the Permian belt that is exposed in southern
Zone throughout the Triassic. The same method has been applied to a Mexico (e.g. Centeno-Garcia and Keppie, 1999; Dickinson and Lawton,
peraluminous Triassic granite (10RC43; Table 1; Fig. 7b) from the Caja- 2001; Kirsch et al., 2014). Remnants of Permian magmatism have also
marca Complex of central Colombia. Similarly, the computed t–T paths been found within the Sierra de Perijá (Dasch, 1982), Paraguana Penin-
also reveal rapid cooling subsequent to Triassic anatexis, after which sula (Van der Lelij, 2013) and the El Baul Massif in Venezuela (Viscarret
the rock was colder than the apatite Pb Partial Retention Zone. Data et al., 2009). The Permian granites and migmatites exposed in Colombia
from the latter sample are new, although the methodology is identical formed in a different tectonic regime to similar lithologies that formed
to that presented in Cochrane et al. (2014b), and the data are presented during the Triassic, which was dominated by extension (see next
in a supplementary file. Rapid cooling during the Triassic corroborates section).
the indistinguishable U-Pb dates from apatites with a large range in Weber et al. (2007) report concordant zircon U-Pb dates from
grain size (Fig. 7b). migmatites of the Maya Block (Chiapas Massif) of 251.8 ± 3.8–254.0
Rapid cooling from anatectic temperatures to less than ~ 380 °C is ± 2.3 Ma, which they interpret to be a result of MP-HT metamorphism
probably mainly a consequence of thermal relaxation subsequent to during compression, and stacking within an orogenic wedge. The colli-
the removal of the heat source at a local geographic scale. Some compo- sion event post-dates the amalgamation of Pangaea, which is recorded
nent of cooling may also be a consequence of exhumation, and given along the diachronous Ouachita-Marathon suture (Fig. 9) that had
that the samples remained colder than ~380 °C throughout the remain- formed by the Early Permian. Weber et al. (2007) attribute the compres-
der of the Triassic and the Jurassic, it is likely that these particular sam- sional event to closure of a marginal basin, which had previously formed
ples were at depths of ≤15 km within the crust after ~ 220 Ma. Rapid during extension that accompanied Permian arc magmatism. Similarly,
cooling during ~250–220 Ma temporally coincides with a significant in- Cardona et al. (2010) attribute anatexis at ~250 Ma in the Sierra Nevada
crease in zircon εHfi obtained from the amphibolitic dykes (Fig. 7a, b). de Santa Marta (Fig. 1) to a compressional event. The cause of compres-
sion has been attributed to either terrane accretion, subduction of thick-
4.7. Interpretation: Permian and Triassic ened, topographically prominent and buoyant oceanic lithosphere (e.g.
Weber et al., 2007), or increased plate coupling during the waning
4.7.1. Arc magmatism and metamorphism during 290–240 Ma along stages of the amalgamation of Pangaea (Cardona et al., 2010).
western Pangaea The high-SiO2, Permo-Early Triassic granites and alkali feldspar
The exposure of Permian and early Triassic (290–240 Ma) granitoids granites of the Eastern Cordillera of Peru are interpreted to have formed
within the Northern Andes is restricted to the Sierra Nevada de Santa within a continental arc setting during extensive lithospheric thinning
Marta, Guajira Peninsula and the central Cordillera Central within (Sempere et al., 2002; Mišković et al., 2009). Roll-back is considered to
Colombia, while almost all intrusions within the Cordillera Real and have driven anti-clockwise rotation of the forearc, dextral displacement
Amotape Terrane of Ecuador are ≤240 Ma. The Aluminium Saturation and southward younging of the onset of extension along the Peruvian
Index of the alkali-calcic to calcic granites that crystallised during arc, all of which commenced at ~280 Ma (Mišković et al., 2009). Perm-
290–240 Ma straddles the peraluminous and metaluminous fields, ian arc magmatism in Peru is consistent with arc magmatism that is
while the d18O values of 14%–17‰ suggest that these rocks formed by sporadically preserved within the Northern Andes, and more abundant-
partial melting of sedimentary rocks. Zircons yield magmatic Th/U ratios ly within the conjugate margin that is now exposed in southern Mexico.
of 0.26–1.27, suggesting that they have not recrystallized during subse- Weber et al. (2007) propose that the Permian arc intrusions within the
quent metamorphic events (Table 1; Fig. 8). The whole rock trace ele- Maya Block were also emplaced during lithospheric thinning. However,
ment abundances are characteristic of subduction related magmatism. their interpretation is based on the tectonic-switching model of Collins
Magmatism during this period was not accompanied by mafic dyke em- (2002), who state that a thinned, hot lithosphere is a prerequisite for
placement, and all rocks yield N58 wt% SiO2. The 290–240 Ma granites subsequent crustal thickening and anatexis within a period of ~20 My.
within the Sierra Nevada de Santa Marta (Cardona et al., 2010) and Geochemical and isotopic evidence from Permo-Triassic intrusions in
the Cordillera Central of Colombia (Villagómez et al., 2011) are Peru provides no evidence for compression at ~250 Ma (Mišković et al.,
interpreted to have formed above an east dipping Pacific subduction 2009). Coeval compression and extension along different regions of the
zone beneath Pangaea (Cochrane et al., 2014a; Fig. 8). 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 (±2σ) 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: Ibagué, L: Loja, M: Medellin, P: Pasto, Q: Quito. Faults. Faults: BF: Baños Fault, CAF: Cauca-Almaguer Faults, LF: Llanganates Fault, OPF: Otú-Pericos Fault, PF: Peltetec Fault.
Map compiled from Litherland et al. (1994) and Gómez 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
14 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

A Frequency
9

8 Triassic Permian

7
6
Anatectites
5 Colombia (n = 27)
4 Ecuador (n = 14)
3
2
1

0
200 225 250 275 300
206Pb/238U age (Ma)
B C
Frequency Palaeozoic metasedimentary
Frequency units (melanosome)
100 9 480 - 630 Ma
Triassic anatectite 960 - 1200 Ma
Colombia (n = 360) 8
N = 99
Ecuador (n = 288) 7

Permian anatectite 6
Colombia (n = 104) 5
10
4
3
2
1

0 0
220
280
340
400
460
520
580
640
700
760
820
880
940
1000
1060
1120
1180

360
420
480
540
600
660
720
780
840
900
960

1140
1020
1080

1200
1260
1320
1380
1440
206Pb/238U age (Ma) 206Pb/238U age (Ma)
D Zircon U-Pb
age (Ma)±2σ
Ecuador Colombia Venez’ Peru
300
SNSM EBM
290
Permian

280
PP
270

260
GP
250 MA
A
240
Triassic

230

220
Metagranite,
210 migmatite
Amphibolite,
200 plagiogranite
190
6 4 2 0 2 4 6 8 10 12
South North
Latitude (degrees)

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 (Villagómez 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 (Mišković 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
A B C
Caribbean Large Igneous Province (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the

10 Peru 4.0
Colombia - Triassic Perm.- Mid.
Peraluminous
Na2O + K2O - CaO (wt%)
8 Colombia - Permian Triassic 3.5
Ecuador - Triassic 2.0 I-type S-type

Al/(Ca+Na+K)
3.0
6 Peru

Metalluminous
Al/(Na+K)
Late Triassic 2.5
4 1.5
Alkalic 2.0

2 1.5 Peru
l cic Late Triassic Peraluminous
i- ca l i ne 1.0 1.0
l
0 lka a lka S-type
Peru Metaluminous
A c- Perm.- Mid. Triassic
l Lachlan FB 0.5
Ca Calcic 0.5 1.0 1.5 2.0 2.5 3.0
-2 0.5
Al/(Ca+Na+K)
50 55 60 65 70 75 80 10 11 12 13 14 15 16 17 18
D SiO2 (wt%) E F 18O (quartz)
1000 10

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


1000 Ecuador - Triassic Ecuador - Triassic
Ecuador - Triassic Colombia - Permian Colombia - Permian
Colombia - Triassic
100
100

10
1
10 1

0.10 N-MORB
1 UCC
Cs RbBa Th U Nb Ta La Ce Pb Pr Sr Nd Zr Hf SmEu Ti Tb Y TmYb 0.1
Cs Rb Ba Th U Nb Ta La Ce Pr Sr Nd Zr Hf SmEu Ti Tb Y TmYb
H
N-MORB 280
0.1 Colombia
Rb Ba Th Nb La Ce Pb Sr Zr Ti Y
G 270 Ecuador Peru
Perm.- Mid. Triassic
age±2

10 0
Ecuador - Triassic 260
Colombia - Permian
10 250
206Pb/238U

240
1
230

0.1 220
Peru
Late Triassic
N-MORB 210
0.01 8 10 12 14 16 18 20
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Yb Lu
(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 Permo-
Triassic intrusions within the Eastern Cordillera of Peru are from Mišković 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), Martínez (2007), Cardona et al. (2010), Cochrane et al. (2014a).

15
16
A B C D
Caribbean Large Igneous Province (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
Please cite this article as: Spikings, R., et al., The geological history of northwestern South America: from Pangaea to the early collision of the

K2 O Th ppm SiO2 wt% V (ppm)


5 10 Ecuador (dykes, sills) 60 500
High-K ARC
Shoshonitic Colombia (dykes, sills) calc-alkaline
4 Colombia (Aburrá Ophiolite)

10
400

=
MORB

Ti/V
3 High-K Andesite
1 BABB
calc-alkaline Calc-alkaline 55

0
=2
Phonolite 300

V
2

Ti/
Rhylite, granodiorite
granite,
Calc-alkaline Sub-alkali =50
Ti/V

Basaltic , diorite

alt
andesite
basalt

li bas
1 200
Tholeiite

dacite,
Tholeiitic 0.1 50 OIB

Alka
andesite
0 00

Basalt
Ti/V=1

Gabbro
45 50 55 60 65 70 75 Bas/Trach/ 100
SiO2 Neph

,
0.01 45 0
0 10 20 30 40 50 60 70 80 0.001 0.01 0.1 0 5 10
Co ppm Zr/TiO2 Ti/1000 (ppm)

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


E F
100 3.5
Ecuador (dykes, sills) N-MORB Ecuador (dykes, sills)
Colombia (dykes, sills) 3.0
Colombia (dykes, sills)
2.5 Colombia (Aburrá Ophiolite)
10
2.0
1.5
1 1.0
0.5
N-MORB 0.0
0.1 La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Yb Lu
Cs Rb Ba Th U Nb Ta La Ce Pb Pr Sr Nd Zr Hf Sm Eu Ti Tb Y Tm Yb
G H I
12
N-MORB

age (Ma)
age (Ma)

240 240

8 Altered

Nd
230 230 amphibolite
206Pb/238U
206Pb/238U

4
220 220

210 210 0
0 1 2 3 0 2 4 6 8 10 12 0.700 0.704 0.708 0.712 0.716
87Sr/86Sri
(La/Yb)n Nd

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 Aburrá 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: Martínez (2007), Cochrane et al. (2014a).
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 17

the location of terranes that were outboard of South America (Fig. 9). Ophiolite; Fig. 2) that formed within a back-arc basin. U-Pb dating of
This supports a hypothesis that compression at ~ 250 Ma was a con- magmatic zircons from the plagiogranite yields an age of ~ 216 Ma
sequence of the accretion of terranes that now form the basement of (Table 1) suggesting that seafloor spreading started between ~ 223
Central America. Alternatively, varying stress regimes may be a conse- and ~ 216 Ma. The youngest plateau 40Ar/39Ar date of 213.7 ± 0.9 Ma
quence of varied displacement at the ocean–continent plate interface (Fig. 2; Table 1) obtained from an anatectite in Colombia records cooling
to the north and south of the Huancabamba deflection. below ~300 °C as the margin drifted further from the source of elevated
heat flow, and possibly exhumation during continued extension. Some
4.7.2. Initiating the disassembly of western Pangaea during 240–200 Ma regions of the continental margin remained at temperatures that
The crustal anatectites of the Cordillera Central of Colombia and the were sufficiently high to form metamorphic zircon rims until ~207 Ma
Cordillera Real of Ecuador that formed during 240–225 Ma yield lower (Fig. 3d; Table 1; Cochrane et al., 2014a). The rift and transition to
whole rock (La/Yb)n (Fig. 8), lower zircon Th/U, and are significantly drift occurred over a period of ~20–25 My (Fig. 8), which is comparable
more peraluminous than the Permian granitoids in Colombia, the East- with the duration of other examples of rifting and the transition to a
ern Cordillera of Peru (Mišković et al., 2009), and the Maya Block of drift phase, including the Lau-Havre-Taupo system (south Pacific;
southern Mexico (Weber et al., 2007). These data are indicative of in- Parson and Wright, 1996), and the west Iberian–Newfoundland con-
creased partial melting of pelitic rocks after 240 Ma, forming metamor- jugate margins (Russell and Whitmarsh, 2003). The combination of
phic zircon (zircon Th/U values of b ~0.1; Rubatto, 2002; Hartmann and bimodal magmatism, an increase in the depleted mantle component
Santos, 2004; Fig. 8) within the anatectites, and perhaps an overall of mafic magmas, increased partial melting of the crust, geochemical
greater volume of magmatism compared to prior to 240 Ma, which is compositions of a back-arc basin or mid-ocean ridge setting, and the for-
reflected by a peak in the quantity of dated samples during 240– mation of oceanic lithosphere at ~216 Ma (Aburrá Ophiolite) is entirely
225 Ma (Fig. 3a). Cochrane et al. (2014a) suggest that mafic under- inconsistent with collisional orogenesis.
plating elevated the geothermal gradient, driving fluid expulsion from Rifting starting at ~240 Ma is consistent with the geochronology and
the pelitic protoliths and lowering their solidus. geochemistry of highly peraluminous granites and volcanic tuffs in the
The trace element content of the amphibolitised tholeiitic dykes is Eastern Cordillera of Peru. Mišković et al. (2009) suggest that decom-
characteristic of a back-arc basin or MORB setting, and the progressive pression melting generated basaltic magmas at the base of the lower
trend in isotopic compositions towards the depleted mantle, combined crust, resulting in crustal melting during 260–200 Ma, ultimately giving
with progressive depletions in incompatible elements during 240– rise to the bimodal Mitu Group within a continental rift. The range
225 Ma suggests that they were emplaced within a lithosphere that in dates of magmatic rocks of the Mitu Group (217–238 Ma; see
was thinning. Hafnium isotopic compositions in zircons shows that Section 4.2.2) closely overlaps with the timing of rifting within
mantle derived tholeiites emplaced during 240–232 Ma assimilated iso- Ecuador and Colombia. Reitsma (2012) interpret the rift as a back-arc
topically evolved continental crust (Figs. 6 and 8), whereas there is little basin on the assumption that the Triassic arc has been removed by tec-
evidence of the assimilation of significant crust after 225 Ma, when εHfi tonic erosion. There is no evidence for the formation of Triassic oceanic
approaches depleted mantle compositions. lithosphere within the Eastern Cordillera of Peru, although the rift foun-
The geographically widespread occurrence of coeval tholeiitic, dered during ~220–190 Ma, resulting in the deposition of limestones of
amphibolitised dykes and crustal anatectites supports an extensional the Pucará Group during thermal sag (Rosas et al., 2007). Extensive
setting, which probably formed within a region of increased heat flow tracts of continental Triassic–Precambrian continental crust, including
that is characteristic of an attenuated back-arc basin (e.g. Collins and the Arequipa Terrane, lie outboard of the Mitu Group, and it is likely
Richards, 2008). Cochrane et al. (2014a) suggest that the period be- that the rift failed to advance to a drift phase, and remains as an
tween 240 and 225 Ma was dominated by progressive thinning of the aulocagen (Fig. 9). In contrast, no Triassic continental crust is found out-
continental lithosphere during rifting and the disassembly of western board of the Triassic anatectites within Colombia and Ecuador (Fig. 2),
Pangaea (Fig. 8). Mafic underplating and anatexis occurred because of and this region advanced to the drift phase. The recent discovery of Tri-
doming and decompression of the asthenosphere during extension, assic basalt offshore northern Peru (8°S; Romero et al., 2013) suggests
and heat convection during the intrusion of mafic magmas into the that this may also be a product of extension of the continental crust.
crust (Fig. 9). This interpretation is supported by i) the tentatively These rocks are exposed along the Outer Shelf High (Romero et al.,
mapped Triassic Zumba Ophiolite in the southernmost Cordillera Real 2013), and no continental crust has been found to the west, suggesting
of Ecuador, although its Triassic age is assumed based on its structural that the Central American conjugate margin may have extended into
position (Fig. 1; Litherland et al., 1994), and ii) late Middle to Upper Peru, within Pangaea.
Triassic continental and marine volcano-sedimentary rocks of the Triassic U-Pb dates obtained from the amphibolites and crustal
Piuntza Unit that were deposited in rift grabens (Litherland et al., anatectites from northern Colombia to southernmost Ecuador reveal
1994; Fig. 2). Anatexis during low pressure metamorphism and exten- no trend with latitude (Fig. 3d), suggesting that extension and rifting
sion has been recorded in several locations, including the eastern Mt. was not diachronous within the northern Andes. Similarly, Reitsma
Lofty Ranges (Oliver and Zakowski, 1995) and the southern Menderes (2012) compare U-Pb dates of detrital zircons within the Mitu
Massif, Turkey (Bozkurt and Park, 1994). Gerbi et al. (2006) utilize ther- Group along the strike of the Eastern Cordillera in central and south-
mal modelling to show that crustal scale detachment faulting is re- ern Peru, and conclude that the basal strata were coeval everywhere,
quired to provide sufficient heat for low-pressure anatexis. A suspect, although Triassic plutons young from central towards southern Peru
Triassic detachment fault has not been identified within the Eastern (Mišković et al., 2009). Rifting along western Pangaea commenced at
(Ecuador) and Central Cordillera (Colombia), and it is likely that is ~ 240 Ma and affected more than 2500 km of the Gondwanan margin.
was reactivated during subsequent tectonism (see below). Evidence includes Middle to Late Triassic extensional basins that are
No products of melting continental crust have been found younger dispersed throughout large regions of South America, including
than 225 Ma, and the basaltic amphibolites that formed during and southern Chile, western Argentina (230–246 Ma from the Cuyana
after 225 Ma yield N-MORB isotopic and geochemical signatures Basin; Franzese and Spalletti, 2001; Barredo et al., 2012), southern
(Figs. 6 and 8). Cochrane et al. (2014a) suggest that the continental Brazil (Zerfass et al., 2004) and Bolivia (Sempere et al., 2002), sug-
crust was either extremely thin, or not present, after 225 Ma. Martínez gesting that western Gondwana was placed under tension (e.g.
(2007) document a series of metagabbros, amphibolites and plagio- Ramos, 2009) at ~ 240 Ma, culminating in the fragmentation of
granites from the northern Cordillera Central, and utilize petrological western Pangaea by ~ 180 Ma.
observations, isotopic and geochemical analyses and field mapping We tentatively suggest that rifting along north western South
to conclude that they are part of an ophiolitic sequence (the Aburrá 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
18 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

A 20
DM
Extension,
10
crustal growth
CHUR
εHfi (zircon) 0

-10

-20 Crustal
Recycling Metagranite zircon cores
-30 Metagranite zircon rims
Amphibolite zircons
-40
150 250 350 550 750 950 1150
206Pb/238U age (Ma)
B
DM
15

10
εHfi (zircon)

εHfi (zircon)
5

N-MORB
0

-5
Range of anatectites
-10
MS
MS
MSWD 0.78 M
MS
MSWD 1.60 MS
S
MSWD 1.90 2 1 0
MSWD
WD 1.90
whole rock
11RC14 10RC28 11RC10 10RC39A (La/Yb)n
220 225 230 235 240
206Pb/238U age (Ma)
C

100 um
100 um

10RC39A 238.3 ± 8.7


εHf = + 8.6 240.0 ± 3.5
234.0 ± 5.8 εHf = - 4.1
εHf = + 8.4

237.5 ± 7.1
241.0 ± 4.0 εHf = + 7.4
εHf = - 3.6 236.9 ± 7.1
εHf = - 4.8

100 μm
235.3 ± 6.1 232.7 ± 6.3
εHf = + 9.1 εHf = +10.0

D Granitoids, zircons, Eastern Cordillera, Peru


6

4
εHfi (zircon)

-2

-4

-6

-8 Eastern Cordillera, Peru


Merida Andes, Venezuela
-10
200 220 240 260 280 300
206Pb/238U 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 19

rocks are preserved along the back-bone to the southern coast of Regardless of the true position of Oaxaquia, Triassic roll-back of
Mexico, and within the Chaucús complex of central Guatemala (Solari the east-dipping Pacific slab drove widespread extension along
et al., 2011). These granites yield zircon U-Pb dates spanning from the N2500 km of western Gondwana, resulting in complete rifting of
Early Permian until ~ 218 Ma (Torres et al., 1999; Keppie et al., 2006; continental crust in the region of northwestern South America.
Solari et al., 2011), and thus they were coincident with rifting. Further- Tensile forces failed to form oceanic lithosphere further to the
more, Helbig et al. (2012) describe transitional arc to MORB like meta- south (e.g. the Mitu Aulocagen; Fig. 9), perhaps because the ortho-
volcanic rocks within the Acatlan Complex (southern Mexico), which gonal component of roll-back was less severe, or the continental
these authors assigned to a back arc setting in the late Triassic. This in- crust was stronger.
terpretation is consistent with Martínez (2007) who suggests that the
Aburrá ophiolite formed in a back-arc. 4.9. Rifting between North and South America
We consider rifting between 240 and 216 Ma within northwest-
ern South America to represent the initial stage of the disassembly The relationship between the timing of rifting that separated North
of Pangaea because i) it was geographically extensive along western and South America, and back-arc rifting along western Pangaea is un-
Gondwana, ii) it resulted in the formation of oceanic lithosphere to clear. Rifting commenced in south-eastern North America at ~230 Ma
the west of the future site of the Gulf of Mexico and the locus of sub- (Schlische, 2002), and possibly also within far southern North America
sequent disassembly within Pangaea, giving rise to the Central (e.g. northern Florida; Beutel, 2009). Magmatic injection in the form of
Atlantic Magmatic Province at 202 Ma (e.g. Beutel, 2009), and iii) it giant dyke swarms (Central Atlantic Magmatic Province) accompanied
overlapped in time with rifting in south eastern North America rifting within North America at 200 ± 2 Ma (Beutel, 2009 and refer-
(~ 230 Ma; Schlische, 2002). However, it is unclear what caused the ences therein), which coincides with ~ 202 Ma tuffs at the base of the
locus of extension to displace eastwards, leading to fragmentation rift-related, la Quinta Fm. of the Merida Andes, Venezuela (Fig. 1; Van
of Pangaea. der Lelij, 2013), and oceanic crust started forming between North and
The Pacific margin of northwestern Gondwana remained passive South America at ~180 Ma. Beutel (2009) utilise finite element model-
until ~ 209 Ma, after which a prolonged period of latest Triassic–Early ling of stress fields obtained from dyke and rift orientations to suggest
Cretaceous magmatism (see Section 5) modified the margin of that the separation of North and South America was initiated by the mi-
Colombia and Ecuador, and southernmost Peru. gration of North America towards the northwest at ~230 Ma, giving rise
to amagmatic rifting within North America, followed by the displace-
ment of South America towards the southwest at ~ 200 Ma, forming
4.8. Conjugate margins to northwestern Gondwana coeval magmatism, rifting, and seafloor formation by ~180 Ma. The re-
lationship between slab-retreat tensional forces that drove extension
Several lines of evidence suggest that the basement to Central along western Gondwana starting at 240 Ma and forces that separated
American terranes (e.g. Oaxaquia, which is exposed within the Maya North and South America is unclear.
Block) may have been the conjugate margin to northwestern Gondwa-
na (Fig. 9). These include i) Grenville-aged granulite belts with similar 5. Latest Triassic–Lower Cretaceous: arc magmatism and
Pb isotope signatures (Restrepo-Pace et al., 1997; Ruiz et al., 1999; Cam- tectonic switching
eron et al., 2004), ii) anorthosite complexes (Keppie and Gutierrez,
1999; Restrepo-Pace and Cediel, 2010), iii) similar Cambrian fauna Jurassic rocks within the Cordillera Real and Cordillera Central are
(Cocks and Torsvik, 2002; Restrepo-Pace and Cediel, 2010), and iv) Tri- dominated by extensive granitoid batholiths that form the eastern
assic bimodal magmatism and rift-related sedimentary rocks (Keppie flanks of the cordilleras (Fig. 1). A majority of these intrusions are not
et al., 2006). Late Triassic (216–197 Ma) tholeiitic basalts in the Guerre- metamorphosed or foliated, with the exception of parts of the aerially
ro composite terrane of Mexico are interpreted to have formed in a con- extensive Ibagué Batholith in central Colombia, proximal to major
tinental rift (Keppie et al., 2006), and the same authors attribute Late faults, and the Azafrán Batholith in Ecuador. The Jurassic intrusions
Triassic rocks of the Oaxaquia Terrane to a back-arc setting. Similarly, are accompanied by coeval volcanic rocks that are mainly exposed
Helbig et al. (2012) suggest that late Triassic volcanic rocks with E- in the foreland basin to the east, and extend from northern Chile to
MORB to MORB characteristics within the Acatlán Complex formed Colombia (Romeuf et al., 1995). Some of the protoliths of meta-
within a back-arc. Weber et al. (2007) and Cochrane et al. (2014a) use sedimentary and meta-volcanic rocks located to the west of the batho-
these similarities to suggest that Oaxaquia and other Mexican terranes liths may also be Jurassic.
were juxtaposed against northwest Gondwana. A potential obstacle to
this reconstruction is the lack of magmatic rocks in southern Mexico 5.1. Historical perspective and occurrence
with crystallisation ages spanning between 230 and 216 Ma. However,
an asymmetric distribution of magmatic rocks across continental rifts 5.1.1. Latest Triassic–Jurassic granitoid intrusions
has also been documented across the Nova Scotia–Morocco Rift (latest Jurassic plutons within the eastern flank of the Cordillera Real in
Triassic–Jurassic; e.g. Dehler, 2012), where reduced magma production Ecuador form, from north to south, the undeformed and unmetamor-
is related to the proximity of the transform margin (Funck et al., 2004). phosed Rosa Florida, Abitagua and Zamora batholiths (Fig. 10). The
Alternatively, Ramos and Aleman (2000) and Ramos (2008) propose granitoids were first described by Colony and Sinclair (1932) and Saur
that continental crust of Oaxaquia was located offshore central and (1950), and the British Geological Survey mapped these intrusions
northern Peru, based on correlations between trilobite assemblages. In (Litherland et al., 1994). These rocks are generally coarse, biotite quartz
this case, the reconstruction of Central American and Mexican basement monzonites, monzogranites and hornblende-biotite granites, granodio-
relative to South America within Pangaea will need to be revised. rites 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. εHfi 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 εHfi (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 εHfi 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) εHfi (zircon) data from Permo-Triassic granitoids of the Eastern Cordillera of Peru (rims and cores; Mišković 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
20 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

A 09RC42, migmatitic leucosome, Sabanilla Unit, zircon U-Pb 247.2±4.3 Ma


Burial and elevated Collision of
Anatexis heat flow Caribbean Plateau
0 16

εHfi (zircon)
8
200 constrained by
0
Temperature (°C)

(U-Th)/He, FT and
40Ar/39Ar data.
-8
APbPRZ

400
40Ar/39Ar
muscovite

600 Quebradagrande Complex


Azafrán and Chinguál batholiths
Zamora, Abitagua batholiths
Ibague, Segovia batholiths
800 Triassic amphibolitic dykes
Permian Triassic Jurassic Cretaceous Cenozoic

300 250 200 150 100 50 0


Time (Ma)
B 10RC43, granite, Cajamarca Complex, zircon U-Pb 245.0±2.0 Ma
Burial and elevated Collision of
Anatexis heat flow Caribbean Plateau
0 16

εHfi (zircon)
8
Temperature (°C)

200 constrained by
(U-Th)/He, FT and 0
40Ar/39Ar data.

206Pb/238U date (Ma)


-8
400
APbPRZ

40Ar/39Ar 240
muscovite 235
230
600
225
50 100 150 200
Effective diffusion radius (μm)
Triassic Jurassic Cretaceous Cenozoic

250 200 150 100 50 0


Time (Ma)
C Cordillera Central Cordillera Real
20
40
AHePRZ 60
80 AHePRZ
APAZ
Temperature (°C)

APAZ
100
120
160 140
ArPRZ (kspar) ?
200 ?
180
240 ?
ZPAZ 220
280
260 ? ZPAZ
320
Cretaceous Cenozoic
300 Cretaceous Cenozoic

140 120 100 80 60 40 20 0 120 100 80 60 40 20 0


Time (Ma) Time (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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 21

with contemporaneous volcanic rocks of the Misahuallí Fm. to the east, (Cochrane, 2013). Litherland et al. (1994) consider the Salado Terrane
and they may intrude their volcanic pile in a few locations. Late Jurassic– to have formed within a marginal basin setting. Traversing westwards,
Early Cretaceous schists of the Upano Fm. (the Salado Terrane; the Upano Unit is faulted against (Litherland et al., 1994; Llanganates
Litherland et al., 1994) are juxtaposed against the western margin of Fault; Figs. 10 and 11) poorly differentiated Palaeozoic metasedi-
the Jurassic batholiths via the Cosanga Fault (Fig. 10). The foliated and mentary rocks and Triassic anatectites (see Section 4) of the Loja
occasionally gneissic Azafrán Batholith and Chinguál pluton are also Terrane of Litherland et al. (1994). Pratt et al. (2005) interpret this con-
located to the west of the Cosanga Fault, within the Salado Terrane of tact as a metamorphosed intrusive contact. Further west, the Triassic
Litherland et al. (1994; Fig. 10). These plutons are dominated by diorites anatectites are faulted against greenschist grade para-graphitic schists,
and granodiorites, with minor granite, quartz monzonite and gabbro. and submarine meta-andesites and meta-agglomerates across the
Litherland et al. (1994) present Rb/Sr isochron and K/Ar dates from Baños Fault (Fig. 11), which were first described by Sheppard and
all of these intrusive phases, although more accurate estimates of the Bushnell (1933), and subsequently by numerous workers. There have
crystallisation ages are provided by U-Pb zircon dates (Litherland et al., been few geochronological determinations, which include K/Ar dates
1994; Noble et al., 1997; Chiaradia et al., 2009; Cochrane, 2013). Geo- (Herbert and Pichler, 1983; Litherland et al., 1994) and Jurassic pollen,
chemical analyses of these rocks have been presented by Litherland spores and acritarchs (Riding, 1989). Bristow (1973) considered these
et al. (1994) and Cochrane (2013). lithologies to be equivalent to sedimentary rocks that are now exposed
Jurassic batholiths extend northwards from Ecuador, and define within the Western Cordillera of Ecuador (Yunguilla Fm.), although
the eastern flank of the Cordillera Central in Colombia (Fig. 10). These Litherland et al. (1994) interpret these to represent a distinct island
include the geographically extensive Ibagué Batholith, and Segovia arc terrane, and recognise forearc, arc and backarc sequences, which
batholith, where the lithologies are dominated by hornblende tonalities they refer to as the Alao Arc (Figs. 10 and 11). The western margin of
and granodiorites (Alvarez, 1983; Aspden et al., 1987). Jurassic intru- these sequences is defined by the 1–2 km wide Peltetec fault zone,
sions also crop out in the Sierra Nevada de Santa Marta where tonalities which hosts inliers of steeply dipping, anastomosed blocks of peridotite,
are also common (Alvarez, 1983), and the Santander Massif (Figs. 1 and olivine gabbros, spilitised dolerites, basalt and volcanoclastic rocks
10), where the intrusions are mainly biotite monzonites, granodiorites (Fig. 11). These rocks were initially documented by Litherland et al.
and granites (Irving, 1975). These intrusions are generally unfoliated, (1987), and are described in detail in Fortey (1990) and Litherland
and they crop out to the east of the Triassic anatectites. Numerous et al. (1994) who suggest that they define an ophiolitic assemblage,
authors presented a large quantity of K/Ar and Rb/Sr dates from these which they refer to as the Peltetec Unit. Pratt et al. (2005) describe a
intrusions in the 1980's (e.g. Goldsmith et al., 1971; see additional refer- stratigraphic transition from the greenschists of the Alao Arc into
ences in Aspden et al., 1987), although more accurate zircon U-Pb dates metasedimentary rocks located west of the Peltetec Unit, and thus
have been published by Leal-Mejia et al. (2011), Villagómez et al. they rule out the Peltetec Fault as a terrane boundary. This publication
(2011), Van der Lelij (2013) and Cochrane (2013). presents new 40Ar/39Ar dates from plagioclase extracted from basalts
of the Peltetec Unit, and we are not aware of any other geochronological
5.1.2. Late Jurassic–Early Cretaceous rocks to the west of the data. Finally, the Peltetec Fault defines the eastern boundary of a series
Jurassic intrusions of gently dipping slates and quartzites, which are exposed in isolated in-
A series of poorly dated, Jurassic–Early Cretaceous meta-sedimentary liers. These marine rocks were first reported by Litherland et al. (1994),
and meta-volcanic rocks, along with the Azafrán Batholith, is exposed to and host Jurassic–Early Cretaceous ammonites. Litherland et al. (1994)
the west of the Cosanga Fault in the Cordillera Real of Ecuador, and named these the ‘Guamote Sequence’ and suggest that they form part
similar sequences crop out within the Cordillera Central (Fig. 10). Within of allochthonous continental crust of the Chaucha Terrane (Figs. 10
Ecuador, Aspden and Litherland (1992) and Litherland et al. (1994) and 11), whereas Pratt et al. (2005) suggest that they are a stratigraphic
utilise structural, metamorphic and geochronological relationships to continuation of the Alao volcanic sequence.
group these into several north-south trending, fault bounded terranes, The same sequence of rocks is also exposed with the Amotape
which accreted via dextral strike-slip transpression. However, Pratt Terrane of southern Ecuador (Fig. 10), which is rotated ~65° clockwise
et al. (2005) re-interpret several of these faults as intrusive contacts, relative to the Cordillera Real. Here, the Triassic anatectites are faulted
and suggest that the Cordillera is an eroded and uplifted Palaeozoic against a complicated mélange of deformed and metamorphosed igne-
core, flanked by autochthonous Jurassic–Cretaceous sequences. ous rocks that are considered to represent an accretionary prism. The
Meta-turbidites and meta-andesites of the Upano Unit, which were mélange hosts eclogites and blueschists of the Raspas unit (Litherland
originally described by Saur (1965), are located west of the Jurassic et al., 1994; Arculus et al., 1999; Bosch et al., 2002; John et al., 2010),
batholiths (Fig. 11). Baldock (1982) and Pratt et al. (2005) suggest which may be a tectonic equivalent of the Peltetec Unit.
that the metasedimentary rocks are equivalent to Cretaceous sedimen- Metamorphosed igneous and sedimentary rocks of the Quebrada-
tary rocks (Hollin and Napo fms.) that are exposed within the region of grande Unit are juxtaposed, via the San Jerónimo Fault, against the Tri-
the Amazon Foreland Basin (Fig. 1). However, Litherland et al. (1994) assic Cajamarca Unit in the Cordillera Central of Colombia, and thus they
regard the Upano Unit as having a distinctly different history based on are located in a similar structural position as the Alao Terrane in Ecuador
their metamorphic grade, and assign them, along with the Azafrán Bath- (Figs. 1 and 10). The Quebradagrande Unit was first described by Botero
olith, to the fault bounded “Salado” Terrane. Sparse geochronological (1963) and is exposed in highly deformed inliers along the entire length
analyses include pollen analyses (Riding, 1989), K/Ar dates (Feininger of the Cordillera Central. Lithologies are dominated by low-grade meta-
and Silberman, 1982), and U-Pb zircon dates of the Azafrán Batholith morphosed 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 t–T 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 t–T paths colder than 350 °C and younger than 75 Ma are constrained by multi-phase 40Ar/39Ar, fission track and (U-Th)/He data (Spikings et al., 2000, 2001, 2005, 2010). The blue line
is the best-fit 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. εHfi (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) t–T 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, fission track and (U-Th)/He data (Spikings et al., 2000, 2001, 2010; Villagómez 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
22 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

CRUSTAL ANATECTITES AMPHIBOLITES


Tectonic
Th/Uzircon (La/Yb)n εHfizircon εNdi (La/Yb)n stage
0.001 0.1 10 8 10 12 14 16 18 20 20 15 10 5 12 10 8 6 4 0 2

180

Jurassic
190

N-MORB
Active
margin
200

DM
210
DRIFT/
Aburrá Passive
ophiolite margin
220
Triassic

230
RIFT
Time (Ma)

240

250

Ecuador
Zircon
260
Whole rock Early
Colombia anatexis
Zircon Continental
Permian

270 Whole rock Arc?

280

290

Fig. 8. Geochronological and geochemical summary of data from Permian and Triassic crustal anatectites, tholeiitic basaltic sills and dykes, and the Aburrá Ophiolite. These data define
i) Permian to earliest Triassic anatexis (290–240 Ma) during arc magmatism, ii) bimodal magmatism during lithospheric thinning and continental rifting (240–223 Ma), and iii) a passive
margin stage during which oceanic lithosphere was forming (223 Ma–209 Ma). Data are from Vinasco et al. (2006), Martínez (2007) and Cochrane et al. (2014a).

2006; Villagómez et al., 2011; Rodriguez and Zapata, 2013), and marine Barragán (e.g. Bustamante et al., 2012) complexes, amongst others
metasedimentary rocks that are covered by marine and terrestrial (Fig. 10). These authors describe eclogite and blueschist facies meta-
meta-sedimentary rocks of the Abejorral Fm., which hosts Hauterivian morphic rocks (epidote glaucophane schists and chlorite-lawsonite
to lower Albian fossils (González, 1980). The highly deformed nature schists), associated with amphibolites, meta-ultramafic rocks, serpenti-
of the Quebradagrande Unit lead Maya and Gonzales (1996) to assign nites and protocataclasites. Rocks exposed in the Jambaló region are
to term Quebradagrande Complex, which will be used in this review. faulted against the eastern margin of the Quebradagrande Complex
Fossil evidence suggests that the sedimentary component was depo- (Bustamante et al., 2008), although the Arquía and Barragán sequences
sited during the Berriasian to Aptian (145–112 Ma using the timescale crop out to the west of the Quebradagrande Complex, across the Silvia-
of Gradstein et al., 2004; see Nivia et al., 2006 and references therein). Pijao Fault, and their western margin is limited by the Cauca-Almaguer
Other geochronological analyses include K/Ar dates (Toussaint and Fault (Fig. 10). Early radiometric dates of these rocks were K/Ar analyses
Restrepo, 1978), and concordant U-Pb (zircon) dates (Nivia et al., 2006; (Feininger and Silberman, 1982), and numerous 40Ar/39Ar dates have
Villagómez et al., 2011; Cochrane, 2013). Geochemical analyses of the been reported (Bustamante et al., 2008; Villagómez et al., 2011;
meta-volcanic rocks are presented by Nivia et al. (2006), Villagómez Bustamante et al., 2012). The Arquía and Barragán complexes occupy
et al. (2011), Rodriguez and Zapata (2013) and Cochrane (2013). Tec- an equivalent tectonic position as the Peltetec and the Raspas
tonic interpretations vary from mid-ocean-ridge (e.g. González, 1980; complexes in the Cordillera Real and Amotape Terrane of Ecuador,
Bourgois et al., 1987), oceanic arc (Villagómez et al., 2011), continental respectively.
arc (Cochrane, 2013) and an intra-cratonic marginal basin (Nivia et al.,
2006). Vásquez and Altenberger (2005) document small, Early Cre- 5.2. Geochronology
taceous intrusions within the Colombian Eastern Cordillera (Fig. 1),
which they relate to rifting. 5.2.1. Latest Triassic and Jurassic intrusions: Cordillera Real, Cordillera
The Quebradagrande Complex is in contact with faulted slices of Central and the Santander Massif
mafic igneous rocks that have been metamorphosed under medium- Goldsmith et al. (1971), McCourt et al. (1984), Aspden et al. (1987)
to high-pressure and low temperature conditions. These rocks are ex- and Litherland et al. (1994) present Rb/Sr isochron and K/Ar dates from
posed in discontinuous, fault-bounded lenses along the strike of the all of these intrusive phases, although they are not considered to be ac-
Cordillera Central, and include the Arquía (e.g. Restrepo and Toussaint, curate crystallisation ages given the high potential for daughter isotope
1976), Jambaló (Orrego et al., 1980; Bustamante et al., 2011) and 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 23

A A’ Permian - 240 Ma
240 Ma, extension begins Pangaea Ouachita-Marathon-Suture
Back-arc Laurentia
Permian basin formation
Slab arc Mexico
rollback
Oceanic crust O
Lithosphere A
A
Pacific Ch Gondwana
Asthenosphere Ocean A’

240 - 225 Ma Pangaea Peru

Attenuated and underplated Arc


Slab crust
rollback 240- 210(?) Ma

Laurentia
Mexico

A A
O

back
Roll
225 - 216 Ma Central American NW Gondwana
Basement blocks Ch
(Maya?) Continental break-up A’
and ophiolite formation
Mit Gondwana
u(
au
loc
ge
n)

Fig. 9. Schematic reconstruction and cross sections for northwestern South America within western Pangaea during 240–216 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), Elías-Herrera and Ortega-Gutiérrez (2002), Weber et al. (2007) and Cochrane et al. (2014a).

concordant U-Pb zircon dates (Fig. 10; Table 2), which range between Ibagué, Rosa Florida, Abitagua and Zamora batholiths (145–189 Ma).
194 and 209 Ma in the Santander Massif (Van der Lelij, 2013), 147– The range in dates of the unfoliated Jurassic intrusions closely overlaps
189 Ma in the Cordillera Central (Bustamante et al., 2010; Villagómez with the age of Jurassic arc magmatism recorded in southern Peru
et al., 2011; Cochrane, 2013) and 141–182 Ma in the Cordillera Real (Fig. 12), Patagonia and northern Chile (195–147 Ma; zircon U-Pb and
(Chiaradia et al., 2009; Cochrane, 2013). Leal-Mejia et al. (2011) report K/Ar dates; Scheuber and Gonzalez, 1999; Rapela et al., 2005; Castro
a range of zircon U-Pb dates from the Santander Massif and Cordillera et al., 2011).
Central (210–149 Ma), although they do not present their data, and
thus they cannot be evaluated. 40Ar/39Ar plateau dates from retentive 5.2.2. Early Cretaceous magmatic and sedimentary rocks: Cordillera Real
phases (hornblende; closure temperature 500–550 °C) are reported and Cordillera Central
here because they may be reasonable estimates of the time of crystal- A majority of dates obtained from the Upano, Alao Arc, Quebrada-
lisation, although they are frequently younger than the U-Pb zircon grande Complex and the mafic and ultramafic rocks of the Peltetec,
dates (Table 2). A comparison of zircon U-Pb age and latitude (Fig. 12) Arquia, Barragán and Jambaló complexes are K/Ar and Rb/Sr dates
suggests that the timing of the onset of magmatism may become youn- (E.g. Feininger and Silberman, 1982). A meta-andesite of the Upano
ger towards southern Ecuador. However, a better defined trend is seen Unit yields a zircon U-Pb crystallisation age of 121.0 ± 0.8 Ma
when the dates are compared with their distance from the equivalent (Cochrane, 2013). Cochrane (2013) present a U-Pb zircon age for the
Silvia-Pijao and Peltetec Faults (Fig. 11b), indicating that Jurassic Chinguál Batholith of 125.3 ± 0.9 Ma (Fig. 10), which is considered to
magmatism becomes younger as it approaches the approximate reside in the same “Salado Terrane” (Litherland et al., 1994) as the
location of the contemporaneous plate margin. Clearly, latest Triassic Upano Unit, and the Azafrán Batholith (141–144 Ma; Table 2).
magmatism initiated far from the trench at ~ 210 Ma, and these rocks Cochrane (2013) presents 238U-206Pb dates from the rims and cores of
are currently exposed within the Santander Massif (209–194 Ma; detrital zircons extracted from a quartzite of the Upano Fm. (Fig. 13),
Fig. 10; Van der Lelij, 2013) and in the region of Mocoa in the far south- which reveals a large spread in ages from rims and cores, and the youn-
ern Cordillera Central (Leal-Mejia et al., 2011). Magmatism migrated gest age of 143.3 ± 9.9 Ma constrains the maximum stratigraphic age to
westwards at ~ 195 Ma and stabilised within the region that is now the latest Jurassic. These dates suggest that the contacts between the
exposed within the Central Cordillera, throughout the Jurassic. The Azafrán Batholith and parts of the Upano Unit are unconformable,
older magmatic belt (N189 Ma) is not exposed, or did not form in supporting the interpretation of Pratt et al. (2005). Age peaks occur at
Ecuador. The earliest Cretaceous (Gradstein et al., 2012) intrusions of 500–600 Ma, 900–1200 Ma, 1500 Ma and the oldest is ~ 2 Ga. The
the foliated Azafrán Batholith (141–144 Ma; Table 2) occur to the maximum stratigraphic age is consistent with pollen assemblages that
west of the mainly unfoliated Jurassic intrusions (Fig. 11) of the Segovia, have a poorly resolved Early Jurassic–Cretaceous 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
24 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

No zircon U-Pb dates have been obtained from the meta-volcanic 5.2.3. Comparison with Peru and the Merida Andes of Venezuela
rocks of the Alao Arc, located west of the Triassic anatectites (Loja Ter- Jurassic magmatic rocks in Peru are limited to the coast of the
rane) and the Baños Fault (Fig. 10) within Ecuador. Litherland et al. Arequipa Terrane (Fig. 1), and in the southernmost Cordillera de
(1994) report imprecise K/Ar (hornblende) dates of 115 ± 12 Ma and Carabaya (southern Peru, east of the city of Cusco). There is a gap in
142 ± 36 Ma, and pollen assemblages suggest that the volcanoclastic exposure of Jurassic magmatic rocks between southernmost Ecuador
sedimentary rocks are Middle Jurassic (Riding, 1989). If the pollen oc- at 5°S, and the Paracas Peninsula at ~ 17°S (e.g. Sempere et al., 2002;
currence is accurately calibrated against absolute time, then the K/Ar Boekhout et al., 2012). The Jurassic arc along coastal southern Peru is
hornblende dates are partially reset. Cochrane (2013) presents 238U- the northern termination of a continuous belt that extends to central
206
Pb dates from the rims and cores of detrital zircons extracted Chile (28°S; Oliveros et al., 2006). The Ilo Batholith intrudes the
from a quartzite of the Alao Arc (Fig. 13), yielding a minimum age of Grenvillian basement of the Arequipa Terrane, and the uppermost levels
163.7 ± 1.6 Ma (Table 2), constraining its maximum stratigraphic age. intrude volcanic rocks of the Chocolate Fm. The batholith yields concor-
Similar detrital age populations are found compared to the detrital dant zircon U-Pb (LA-ICPMS) dates ranging between 152 and 173 Ma
dates from the Upano Unit, suggesting that they were supplied by the (Boekhout et al., 2012; Fig. 12b), while the Chocolate Fm. yields zircon
same source regions. Furthermore, Early Ordovician acritarchs suggest U-Pb ages ranging between 216 and b135 Ma (Boekhout et al.,
that they have been reworked into the Alao arc sequence (Litherland 2013a). Demouy et al. (2012) obtained concordant zircon U-Pb dates
et al., 1994), corroborating the detrital zircon U-Pb age spectrum ranging between 161 and 200 Ma from diorites and gabbros from the
(Fig. 12). We conclude that no reliable age dates exist for the Alao arc, western Arequipa Terrane. Mukasa (1986) report zircon U-Pb dates
although it is Middle Jurassic, or younger. Villagómez et al. (2011)) between 184 and 188 Ma from a gabbro- tonalite unit located in south-
and Cochrane (2013) report concordant zircon U-Pb dates of magmatic ern coastal Arequipa, and Mišković et al. (2009) present zircon U-Pb
rocks of the Quebradagrande Fm. of 114.3 ± 3.8 Ma (tuff) and 112.9 ± dates from four syenites within the Eastern Cordillera, inboard of the
0.8 Ma (diorite) (Fig. 10; Table 2). Toussaint and Restrepo (1994) report Arequipa Terrane (Cordillera de Carabaya; Allincapac Complex), which
a K/Ar (whole rock basalt) date of 105 ± 0.8 Ma, although it is likely to range between 173 and 195 Ma. These dates suggest that Jurassic arc
be at least partially reset. The U-Pb dates are consistent with Berriasian magmatism may have been active in southern Peru since ~ 216 Ma,
to Aptian (145–112 Ma; Nivia et al., 2006 and references therein) fossil and throughout the Jurassic, which overlaps with the entire span of Ju-
dates, and a maximum stratigraphic age of ~149 Ma obtained from de- rassic dates found in the northern Andes (Fig. 12). Demouy et al. (2012)
trital zircons (Fig. 13). Vásquez et al. (2010) report hornblende 40Ar/ use the distribution of U-Pb dates from Jurassic plutons to suggest that
39
Ar (plateau) dates of 120.5 ± 0.6 Ma (tuff) and 136.0 ± 0.4 Ma (dio- the arc axis within Arequipa migrated oceanward at ~ 175 Ma. The
rite) from small gabbroic intrusions located in the Eastern Cordillera of Jurassic intrusions host few xenocrystic cores within the zircons, with
Colombia (Figs. 1 and 11; Table 1). ages clustering at approximately ~ 500–600, ~ 900–1050, and older
Dating the crystallisation ages of the mafic and ultramafic protoliths (Fig. 12b; Demouy et al., 2012).
of the MP-HP/LT rocks of the Arquía, Barragán, Jambaló, Peltetec and Early Cretaceous granitoid intrusions in Peru are concentrated along
Raspas complexes is problematic due to i) metamorphic overprinting, the coastline, forming the Coastal Batholith within the Western
and ii) the lack of high-U, low-(common)Pb minerals, which are amena- Cordillera of Peru, and along coastal Arequipa, although none have
ble to U-Pb dating. No U-Pb dates have been published, although this been found in the Eastern Cordillera of Peru (e.g. Mišković et al.,
could be addressed by recent developments in dating baddeleyite (e.g. 2009). Early Cretaceous zircon U-Pb crystallisation ages from the Coast-
Chamberlain et al., 2010), which is stable in mafic rocks. Bustamante al Batholith located north of the Arequipa Terrane are ~ 100 Ma
et al. (2012) present a weighted mean 40Ar/39Ar date from three plateau (Mukasa, 1986) and 115–117 Ma (de Haller et al., 2006), which are sim-
white mica dates of 120.7 ± 0.6 Ma (Fig. 10; Table 2) from a single ilar to the ages obtained from the Quebradagrande Complex in
muscovite schist of the Barragán Complex, which they suggest records Colombia (Table 2; Fig. 12). Boekhout et al. (2012) present concordant
retrogression during exhumation. This Early Cretaceous date is similar zircon U-Pb dates ranging between 106 and 110 Ma from coastal Are-
to a plateau hornblende 40Ar/39Ar date of 112.0 ± 3.7 Ma (Villagómez quipa, which were originally mapped as the Ilo Batholith, although
et al., 2011) obtained from the Arquía Complex, which has been they form a distinctly younger age group than the Jurassic granitoids.
interpreted as the time of retrogression. The only other interpretable A compilation of these dates suggests that there is a magmatic gap in
dates from Colombia are plateau 40Ar/39Ar white mica (paragonite and southern Peru between ~110 and ~132 Ma.
phengite) dates from blueschists of the Jambaló Complex, which range No Cretaceous magmatic rocks have been documented in the Merida
between 68 and 62 Ma (Bustamante et al., 2011), which were inter- Andes, and this period is characterised by the deposition of sandstones
preted to record the mylonitic event that was responsible for exhuming and marine mudstones (Van der Lelij, 2013).
the rocks. Numerous dates have been obtained from eclogites and
blueschists of the Raspas Complex in the Amotape Complex (Ecuador). 5.3. Geochemistry
Phengite K/Ar (Feininger, 1980) and 40Ar/39Ar dates (Gabriele, 2002) of
132 ± 5 Ma and 123–129 Ma (Fig. 10) were interpreted to record cooling 5.3.1. Latest Triassic–earliest Cretaceous granitoids
during retrogression of the eclogites. However, these dates must be Latest Triassic to Jurassic (209–141 Ma) granitoids from the Cordillera
interpreted with caution, given the propensity for phengite in high- Real, Cordillera Central and the Santander Massif, are metaluminous
pressure rocks to incorporate excess 40Ar, which is not resolvable by in- (Fig. 15a) and plot in the calc-alkaline to alkali-calcic field of Frost et al.
verse isochron analysis (e.g. Sherlock and Kelley, 2002). John et al. (2001; Fig. 15b), and in the high-K calc-alkaline and calc-alkaline fields
(2010) report Lu-Hf isochron (garnet, whole rock, amphibole, pyroxene) when comparing immobile elements (Fig. 16a). Negative Nb, Ta and Ti
ages of 126.4 ± 4.0 and 129.9 ± 5.6 Ma, while the older estimate is con- anomalies, combined with enriched LILE relative to HFSE, and LREE
sidered to be more accurate, given its MSWD of 2.0 (Table 2). These relative to HREE suggest that these rocks formed in a subduction related
dates are interpreted to record garnet growth (John et al., 2010), setting and they are interpreted as continental arc intrusions (Fig. 15c, d).
and overlap with the phengite 40Ar/39Ar dates of Gabriele (2002). Granitoids from the Santander Massif, which were probably located
Mafic and ultramafic assemblages of the Peltetec sequence have furthest from the plate margin, are slightly more enriched in LILE and
only experienced greenschist grade metamorphism, and we present are more peraluminous that the younger granitoids which intruded clos-
a plagioclase 40Ar/39Ar plateau dates of 134.7 ± 0.9 Ma and 134 ± er to the palaeo-trench. The magmas within the Santander Massif may
13 Ma (Table 2; details in the supplementary material; Fig. 14), have assimilated more continental crust, perhaps because they intruded
which we interpret as the minimum crystallisation age of the basaltic through thicker crust. Finally, the trace and major elements reveal no
protolith. 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 25

the Northern Andes (Fig. 15). εNdi (whole rock) values for the latest blueschists and amphibolites of the Barragán Complex were normal
Triassic–Jurassic granitoids range between −7.2 and 5.3, and the intru- mid-ocean ridge basalts. Similarly, Villagómez et al. (2011) suggest
sions within the Santander Massif host distinctly less radiogenic Nd that the protolith to the amphibolites of the Arquía Complex may
isotopic compositions than younger intrusions located further west have formed at a mid-ocean ridge.
(Fig. 16b). εNdi (whole rock) values from the unfoliated intrusions of Metamorphosed ultramafic–mafic rocks of the Raspas and Peltetec
the cordilleras Real and Central (red and black symbols in Fig. 16b) de- complexes in Ecuador are geochemically similar to the Barragán and
fine no particular trend. εHfi (zircon) has a large range of −6 to 9.25 Arquía units in Colombia. Eclogites of the Raspas Complex yield flat
(Fig. 16c), and a well-defined trend indicates that the isotopic composi- REE profiles when normalised against N-MORB ((La/Yb)n 0.69–2.20;
tion becomes more juvenile as the rocks become younger, although a Arculus et al., 2002; Bosch et al., 2002; John et al., 2010), and their
large range in εHfi (zircon) can be found in the unfoliated intrusions of LILE contents are not significantly enriched relative to the HFSE, while
the cordilleras Real and Central, at any given time. The youngest, earliest negative Nb, Ta and Ti anomalies are missing (Fig. 17h, i). The eclogites
Cretaceous foliated intrusions (Chinguál and Azafrán) are located to the plot in the N-MORB field when comparing Nb/La with (La/Sm)n
west of the older granitoids, and yield the most juvenile isotopic compo- (Fig. 17c), and in the ocean plateau tholeiite field when comparing La/
sitions. These trends suggest that the latest Triassic–earliest Cretaceous Yb and Zr/Th (Fig. 17b). εNdi (whole rock) values for the eclogites are
granitoid intrusions become more juvenile as the intrusions migrate juvenile, and range between 6.9 and 10.8 (Bosch et al., 2002; Fig. 16d),
towards the palaeo-plate margin. The arc intrusions were intruding corroborating their depleted chemical compositions. In contrast, the
through continental crust that was becoming thinner between blueschists from the Raspas Complex have enriched LILE relative to
~194 Ma and ~ 189 Ma, and during the earliest Cretaceous, starting at their HFSE, and elevated LREE relative to their HREE ((La/Yb)n 6.0–
~143 Ma (Azafrán and Chinguál intrusions). This could be interpreted 7.9; John et al., 2010; Fig. 17h, i). The blueschists plot in the seamount
as extension along the plate margin, or the migration of the arc axis field when comparing Nb/La with (La/Sm)n (Fig. 17c), and in island
towards the trench, or a combination of both. arc tholeiite field when comparing La/Yb and Zr/Th, although they lack
Nb, Ta and Ti anomalies when normalised against N-MORB. John et al.
5.3.2. Early Cretaceous igneous rocks (2010); Fig. 17h interpret the eclogites as subducted oceanic litho-
Dacites, andesites, basalts and gabbros of the Quebradagrande sphere that was typical of N-MORB, whereas the protoliths to the
(Colombia) and Alao (Ecuador) sequences are metaluminous and span blueschists were considered to be seamounts. Bosch et al. (2002) sug-
a larger range in Aluminium Saturation Index than the Jurassic granit- gest that the mafic and ultramafic rocks metamorphosed under high-
oids (Fig. 17a). Tectonic discrimination diagrams suggest that these pressure conditions and originally formed part of an oceanic plateau,
rocks formed in a variety of tectonic environments, spanning from while Arculus et al. (2002) report that the protoliths originated as
calk-alkaline arc to island arc tholeiite, ocean plateau tholeiite and both N-MORB, and within an oceanic plateau, and equilibrated with
MORB compositions (Fig. 17b, c). These observations corroborate i) peak conditions at 1.3–2.0 GPa and ≤ 600 °C. Finally, altered gabbros
the N-MORB normalised REE plot (Fig. 17e) which reveals both N- and basalts of the Peltetec unit, which is exposed within the Peltetec
MORB-like compositions, and rocks that are enriched in LREE, which is Fault Zone (Ecuador; Fig. 11) have received less attention, and the
more characteristic of subduction-related rocks, and ii) N-MORB nor- only available geochemical data are from Litherland et al. (1994), and
malised trace element plot (Fig. 17d), which shows that some samples new data that are published in this review (supplementary material).
have characteristic negative Nb, Ta and Ti anomalies, while these are These rocks are generally depleted compared to the other mafic rocks
missing in the rocks that yield almost flat REE patterns. These rock se- (45–50 wt% SiO2; Cochrane, 2013; Fig. 17h) that lie in a similar structur-
quences are exposed within faulted blocks, and distinguishing between al position in Ecuador and Colombia, with lower trace element and REE
rock sequences that form the Quebradagrande Complex and the juxta- abundances, and they yield slightly enriched LREE relative to HREE
posing Arquía Complex is difficult in the field. Therefore, we suggest ((La/Yb)n 2.3–4.6; Fig. 17i), while there is a general absence of distinc-
that some of the basalts that yield N-MORB signatures may be a tive Nb, Ta and Ti anomalies when normalised against N-MORB. The
structurally detached component of the Arquía or Peltetec com- trace element abundances plot close to unity when normalised against
plexes, which yield MORB and E-MORB signatures (see later), and N-MORB, and it is very likely that the LILE trends have been disturbed
are now intercalated within arc rocks of the Quebradagrande and by alteration. The gabbros and basalts plot within the calc-alkaline to
Alao sequences. εNdi (whole rock; Fig. 16d) from the volcanic rocks tholeiitic field when comparing immobile elements (Fig. 16a), and a
of the Quebradagrande and Alao sequences ranges between − 0.64 majority of the rocks plot within the MORB–seamount field when com-
and 7.63 (Cochrane, 2013), and there is a general reduction in (La/ paring Nb/La against (La/Sm)n (Fig. 17c). εNdi (whole rock) values for
Yb)n as εNdi (whole rock) becomes more radiogenic (Fig. 16d). The the Peltetec unit range between 1.1 and 1.2 (this study; Fig. 16b), and
least radiogenic basalts within these volcanic sequences are more ju- they are less radiogenic than those obtained from other mafic rocks
venile than the most radiogenic Nd isotopic compositions obtained (45–50 wt% SiO2; Cochrane, 2013) that lie in a similar structural posi-
from the Jurassic granitoids. Similarly, the single εHfi (zircon) mea- tion in Ecuador and Colombia. We interpret these data to suggest that
surement from the Quebradagrande Complex (Cochrane, 2013) is the protolith to these greenschist facies rocks formed either within an
more radiogenic that the same measurements from the Jurassic ocean plateau, or perhaps within a back-arc basin setting, as transitional
granitoids (Fig. 16c). crust.
Blueschists and amphibolites of the Barragán and Arquía complexes
yield flat REE (N-MORB normalised) multi-element plots ((La/Yb)n 5.3.3. Comparison with magmatic rocks from Peru
0.74–4.68), and their trace element abundances lack strongly negative Granitoids of the Jurassic Ilo Batholith (U-Pb zircon 152–173 Ma; see
Nb, Ta and Ti anomalies (Fig. 17f, g), contrasting with arc-related andes- above) within the Arequipa Terrane of southern Peru are metaluminous
ites and basalts of the Quebradagrande Unit, which is faulted against to slightly peraluminous (Fig. 15a), and show a strong enrichment in
their eastern margin. These features are consistent with a tholeiitic frac- LILE with negative Nb, Ta and Ti anomalies (Boekhout et al., 2012).
tionation trend (Fig. 16a), and juvenile εNdi (whole rock) values of 3.2 εNdi (whole rock) values range between − 5.1 and − 2.4 (Fig. 16),
to 9.6 (Arquía Complex only; Fig. 16d), which are more radiogenic and εHfi (zircon) range between −9.5 and 7, suggesting that mantle de-
than the Early Cretaceous volcanic rocks of the Quebradagrande unit. rived melts assimilated a significant amounts of heterogeneous Sunsas-
The MP-HP/LT metamorphosed rocks plot in the MORB to E-MORB aged (~ 1 Ga) continental crust (Demouy et al., 2012; Boekhout et al.,
field when comparing Nb/La with (La/Sm)n (Fig. 17c), and island arc 2013b), with perhaps variable depths of melting. The few inherited
to ocean-plateau tholeiite field when comparing La/Yb and Zr/Th xenocrystic cores in Jurassic intrusions corroborate the assimilation of
(Fig. 17b). Bustamante et al. (2012) suggest that the protoliths to the 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
26 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

76° 75° 74° 73°


195.8±1.5
Segovia
Cretaceous (Early) Batholith 202.2±1.0
8°N Santander
Ultramafic - mafic H-MP/LT Massif
metabasalts, metagabbros. 196.0±1.1
Arquía, Barragán complexes (Colombia) 201.1±1.4
Raspas, Peltetec complexes (Ecuador) 199.8±1.2
B 208.8±1.2
Metalluminous volcanic rocks
Alao Arc (Ecuador) 200.0±1.5
Quebradagande Complex (Colombia) A 155.6±6.2 198.3±0.8
Metalluminous granitoids 199.1±1.3
M 188.9±2.0
Continental arc intrusions
Azafrán Batholith, Chinguál Pluton 198.0±0.8
Upano Unit (basalts and turbidites) 112.9±0.8 SPF
201.0±0.9
159.2±5.2 200.4±0.7
Jurassic - latest Triassic 149.2±6.1

SJF
168.8±0.7
Slates, quartzites 5°N 146.8±1.5
Guamote Sequence (Ecuador) 114.3±3.8
I 148.9±3.3
Metalluminous granitoids Barragán
120.7±0.3 IF 166.0±10
Continental arc intrusions
Arquía 182.6±2.4
112.0±3.7 159.6±2.4
164.4±1.1
City 156.5±1.1
Ibagué
Batholith 155.7±2.2
0 100 km 3°N 158.5±1.0

CAF
189.1±2.9
78°W Ibagué
Batholith 187.4±2.3
2°N

175.8±1.7
125.3±0.9 179.0±2.0
P 173.6±1.5
Chinguál
Pluton 181.5±1.6
157.9±7.3 Rosa Florida
LF Batholith
PF 0°
Q 182.4±0.6
Azafrán CF
169.8±1.1
Batholith U-Pb zircon
173.0±1.3 244.6±2.4
79°W LA-ICPMS
140.7±1.7 Abitagua
143.5±1.3 Batholith
277.3±3.0 U-Pb zircon
Y TIMS
168.8±2.2 Y’ 174.0±1.2
155.0±6.1 U-Pb zircon
Z (youngest detrital zircon)
Z’ 233.7±4.8 LA-ICPMS
Peltetec 163.7±1.6
134.7±0.9 metasedimentary rocks
134.3±13 40Ar/39Ar plateau date
80°W 143.3±9.9 213.7±0.9
CF hornblende, muscovite,
3°S phengite, plagioclase
Raspas PF
126.4±4.0 121.0±0.8 244.6±2.4 Lu-Hf isochron
129.9±5.6 BF
178.1±1.4
Raspas 145.4±0.2
123.9±1.4 L
T 153.8±1.5
129.3±1.3
132±5 160.5±1.7
Zamora
131.6±1.1 Batholith
5°S

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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 27

A Y Y’
Passive E. Cret. arc Pz - Triassic Jurassic - earliest Cretaceous arc
Margin? basement
Guamote Alao Terrane Loja Salado Terrane Abitagua Batholith
Sequence Terrane
1°21’ S
PF V V V
N V V BF
V LF
V
V CF
3 km

Upano Fm.
Tungurahua
volcano
1°59’ S
E. Cret. arc
Passive
Z Margin? Z’
Guamote Sequence Alao Terrane
V
1°53’ S
V
V
PF V V
V
V
V V
V V
1°59’ S

Peltetec Unit

B
Age (Ma)
Cretaceous ultramafic - mafic complexes
250
Peltetec (Ecuador)
Raspas (Ecuador)
200
Arquía Unit (Colombia)
Gz Barragán Unit (Colombia)
150
Cretaceous Volcanic Rocks
Alao Arc (Ecuador)
100
Quebradagrande Complex (Colombia)
Cretaceous Intrusions
50
0 100 200 300 400 Eastern Cordillera (Colombia)
Distance from suture (km) Jurassic Intrusions
Azafrán and Chinguál (Ecuador)
Zamora, Abitagua (Ecuador)
Ibagué, 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: Garzón Massif.

environment within continental crust, and in this sense they appear 150 km inland from present-day position of the Jurassic palaeo-margin
to be similar to the Jurassic granitoid intrusions exposed within the (Fig. 11b). Middle Jurassic nepheline syenites of the Allincapac igneous
Northern Andes, which yield the same crystallisation ages, although complex in the Cordillera de Carabaya are more alkali than all Jurassic
the Nd isotopic compositions of the Peruvian rocks along the coast of igneous rocks found in the northern Andes (Fig. 15a, b), and formed
Arequipa (Fig. 1, inset) are less radiogenic (Fig. 16b). The Ilo Batholith from SiO2 under-saturated magmas (Mišković et al., 2009), which are
is currently exposed along the coastline of southern Arequipa, whereas considered to have intruded through thinned continental crust in the
coeval granitoid intrusions within the Northern Andes are exposed 30– Jurassic back-arc (Mišković 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 Jurassic–Early Cretaceous
arcs, obducted mafic–ultramafic 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
(±2σ; see references in Table 2) are shown. Sections Y-Y' and Z-Z' are shown (see Fig. 11). Cities, B: Bucaramanga, I: Ibagué, L: Loja, M: Medellin, P: Pasto, Q: Quito. Faults: BF: Baños
Fault, CF: Cosanga Fault; CAF: Cauca-Almaguer Fault, IF: Ibagué Fault, LF: Llanganates Fault, OPF: Otú-Pericos Fault, PF: Peltetec Fault, SJF: San Jeronimo Fault, SPF: Silvia-Pijao Fault.
Map compiled from Litherland et al. (1994) and Gómez 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
28 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

The Coastal Batholith of Peru (~115–37 Ma) is bimodal, and is dom- extensional system. This interpretation is consistent with i) Sempere
inated by metaluminous tonalites (N 70 wt% SiO2) with enriched LILE et al. (2002), who utilize sedimentary facies to demonstrate that
that are typical of calc-alkaline rocks. εHfi values from the Coastal Bath- the margin of Arequipa was extending throughout the Jurassic, and
olith range between 5.2 and ~10.0 from rocks with U-Pb zircon ages of Boekhout et al. (2012) report subsidence rates of ~3.5 km/My, ii) Juras-
~115 Ma (Polliand et al., 2005; de Haller et al., 2006; Schaltegger unpub- sic back arc magmatism through thin continental crust (173–195 Ma;
lished data) along the coastline in Central Peru, and between − 5 and Mišković et al., 2009) in the Cordillera de Carabaya, southern Eastern
7.5 (100–110 Ma) within Arequipa (Demouy et al., 2012; Boekhout Cordillera of Peru, and iii) reconstructions of Pindell and Kennan
et al., 2013b), which are less juvenile than a diorite of similar age (2009), who draw rift zones extending from Central Peru towards
(~ 113 Ma; Cochrane, 2013) from the Quebradagrande Complex northern Colombia throughout the Jurassic period (Fig. 18b). Jurassic
(Table 2; Fig. 16c). εNdi (whole rock) from Early Cretaceous granitoids extensional basins are recognised to the east of the Jurassic intrusions,
along coastal Arequipa range between −4.4 and 0.1 (Boekhout et al., which are characterised by grabens with varying rates of subsidence,
2012), which is also less radiogenic than magmatic rocks of a similar along with marine environments (e.g. Bogota Basin) in Colombia
age in the Northern Andes (Fig. 16b). (Fabre, 1984; Toussaint and Restrepo, 1994). Subsidence gave rise to a
2000 m thick sequence of Sinamurian limestones and sandstones of
5.4. The tectonic setting during the latest Triassic–Jurassic (210–145 Ma) 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).
A combination of field studies, concordant zircon U-Pb dates, geo- The oldest crystallisation ages at any given location (Fig. 12a) show a
chemistry and isotopic data clearly shows that an I-type, metaluminous, general younging trend from northern Colombia to southern Ecuador,
high-K to calc-alkaline continental arc formed within northwestern suggesting that the onset of subduction may have been diachronous
South America at ~ 209 Ma, due to subduction of Pacific oceanic litho- north of the Huancabamba Deflection. However, arc magmatism in
sphere beneath western South America. This time period marks the for- Peru may have started as early as 216 Ma, and occurred until b 135 Ma
mation of a new subduction zone inboard of the Central American (Fig. 18a), as recorded by the Chocolate Fm. and the Ilo Batholith.
terranes that were the Permo-Triassic conjugate margin to northern Calc-alkaline, metaluminous arc intrusions along coastal Arequipa
South America within Pangaea. Arc magmatism during 194–209 Ma were accompanied by coeval, alkali back-arc magmatism in the
was focussed within the rocks of the Santander Massif, currently located Cordillera de Carabaya during 173–195 Ma (Fig. 18a). This time period
280–350 km inboard of the Silvia-Pijao Fault (Fig. 18a). The Peltetec– broadly corresponds with a westward jump of the arc axis in the Northern
Silvia Pijao Fault is the western boundary of the Quebradagrande Com- Andes during 194–189 Ma, although coeval arc-back arc relationships
plex, and is considered here to represent the Jurassic–Early Cretaceous have not been found in the Northern Andes. Clearly, the potentially
continental margin (Fig. 19). These early arc magmas assimilated large diachronous onset of arc magmatism in the Northern Andes cannot be
quantities of continental crust, and are highly enriched in LILE, LREE extended across the Huancabamba Deflection.
and non-radiogenic Nd and Hf isotopes. These rocks were coeval with The t–T paths generated from apatite U-Pb dates (Fig. 7) shows
arc magmatism within southern Peru. The lack of arc magmatism within that the current surface of the Northern Andes was cooler than 380 °C
Venezuela during 194–209 Ma suggests that it was located more than throughout the Jurassic, and thus very little information can be obtained
400 km from the Pacific active margin, while the inter-American gap from the U-Pb data. One sample reveals reheating to temperatures N 380
had not yet opened. °C commencing at a poorly constrained interval time of ~ 150 Ma,
The arc axis migrated westwards after ~194 Ma, giving rise to a long- whereas the other sample reveals no reheating into the apatite lead
lived calc-alkaline, metaluminous arc (141–189 Ma) which is defined partial retention zone (APbPRZ; ~ 380–550 °C). We interpret this as
by the Segovia, Ibagué, Rosa Florida, Abitagua, Azafrán, Chinguál and Za- burial of fault blocks in southern Ecuador at ~ 150 Ma, whereas other
mora batholiths throughout Colombia and Ecuador, and several volcanic fault blocks were not buried, depending on their structural setting
formations (e.g. the Misahuallí Fm. in Ecuador; Figs. 18a and 19). These within the extensional system.
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 5.4.1. Why is there a gap in the Jurassic arc in Peru?
source, or have assimilated less continental crust. A comparison of age The Jurassic arc is recorded along the entire length of the Northern
and longitude suggests that these intrusions are also young to the Andes, although it is missing between ~ 6°S and ~ 12°S (northern
west over a (present day) distance of ~100 km (Figs. 11b and 18a), tra- Peru). This may be because an arc never formed in northern Peru
versing the cordilleras Real and Central, with the oldest intrusions oc- (Fig. 18a), implying there was insufficient subduction to create
curring in the Segovia Batholith (Colombia), and the youngest in the magma due the angle of the South American Plate margin relative to
foliated Azafrán Batholith (Ecuador). This trend suggests that the arc the plate convergence direction. Alternatively, there was a Jurassic arc
axis migrated to the west during 189–141 Ma, albeit at a much slower along the entire Peruvian margin, but it was removed from northern
rate than during 194–189 Ma. Westward migration of the Middle– Peru by either i) tectonic erosion, or ii) lateral displacement. The Jurassic
Late Jurassic arc axis by 100 km is similar to the quantity measured by arc axis in Arequipa crops out along the coastline, ~ 50 km from the
Boekhout et al. (2012) and Demouy et al. (2012) in the region of Arequi- present-day trench, and thus we hypothesise that the Jurassic forearc
pa (from the city of Arequipa towards the coast), although migration at has been removed, and did originally form along the entire margin.
that location is considered to have started at ~175 Ma. Migration of the The Peruvian and northern Chilean margin has undergone rapid tecton-
arc axes may be due to either slab steepening and migration relative to ic erosion since at least the Eocene (Clift et al., 2003; Clift and Hartley,
the trench, or migration of the trench and subducted slab to the west, 2007). The highest rates are found north of the present day intersection
extending the margin. Given that the younger intrusions yield more of the Nazca Ridge with South America, which coincides with the
juvenile Nd and Hf isotopic compositions, and are located further to northern extent of the Arequipa Terrane. Clift et al. (2003) estimate
the west, we suggest that the crust was extending, resulting in progres- that flat subduction of the Nazca Ride beneath Peru has removed ~110
sively less continental contamination of magmas derived from the man- km of continental crust since the ridge started subducting at ~ 16 Ma
tle wedge above a subduction zone that was retreating. Unlike the other (Rosenbaum et al., 2005) or ~ 11 Ma (Hampel, 2002), and that 38 km
Jurassic intrusions, which are generally unfoliated, the younger and was eroded during the preceding ~35 Ma. The Nazca Ridge has not yet
earliest Cretaceous Azafrán and Chinguál intrusions are foliated, and subducted beneath the Arequipa Terrane, which may account for the
are located to the west of the Cosanga Fault in Ecuador (Fig. 10). We preservation of the Jurassic arc, and the spatial coincidence supports
hypothesise that the Cosanga Fault may have originated within an 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 29

A
Zircon U-Pb
age (Ma)±2
Ecuador Colombia Venez’ Peru
210

200
I b a g u é
Segovia
190

no igneous ativity
Rosa
Florida
Zamora
180 Abitagua
Jurassic

170

160
Cordillera Real,
Cordillera Central
150
Santander
Massif
140
Garzón
Massif
130
4 2 0 2 4 6 8
South North
B
Frequency Frequency
80
Cordillera Real
4
70 Cordillera Central
60 Santander Massif
Peru (Arequipa) 3
50

40 2
30
20 1

10
0
0
380
500
620
740
860
980
1100
1220
1340
1460
1580
1700
1820
1940
2060
2180
100
110
120
130
140
150
160
170
180
190
200
210
220

206Pb/238U age (Ma) 206Pb/238U age (Ma)

Fig. 12. A) A comparison of latest Triassic–Jurassic concordant zircon U-Pb dates with latitude along the Cordillera Real, Cordillera Central, and the Santander and Garzón massifs of
Colombia. The ranges of concordant zircon U-Pb dates obtained from granitoid intrusions and volcano-sedimentary rocks from the Eastern Cordillera (Mišković 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 Triassic–Early Cretaceous. C) Age frequency histogram for LA-ICPMS in-situ data (single spots) for the Devonian–Precambrian, revealing a paucity of xenocrystic zircon
cores in the Jurassic arc.

An alternative explanation for the Jurassic gap may be provided by require a pivot point and a dextral shear force, which may have been
palaeomagnetic data obtained from Jurassic tuffs (K/Ar dates only; provided by the northward transcurrent migration of crust from Peru.
Bayona et al., 2010) from northern Colombia (Sierra Nevada de Santa Within this context, Jurassic quartzites and slates of the Guamote
Marta and the Santander Massif; Fig. 1). These data suggest that the Sequence, located to the west of the Peltetec Fault in Ecuador (Fig. 11)
tuffs acquired their permanent remnant magnetization at southern may be a remnant of a dismembered terrane (the Chaucha Terrane of
palaeo-latitudes that corresponded with northern Peru and Ecuador Litherland et al., 1994) that was displacing towards to the north. This
during the Early Jurassic, and the fault blocks were at least within the hypothesis is consistent with the reconstructions of Pindell and
vicinity of Colombia by the Late Jurassic–Early Cretaceous. Bayona Kennan (2009), who invoke a lateral translation of the Tahamí Terrane
et al. (2010) conclude that there has been significant along-margin of the Cordillera Central of Colombia, during the Jurassic–Early Creta-
northward translation of terranes relative to the Amazonian Craton. ceous (Fig. 18b). Lateral translation of terranes along active margins
Therefore, the hypothesis that the Jurassic arc of northern Peru may has been documented in Nova Scotia, within the Rheic Ocean Wilson
have displaced northwards, and is currently partly exposed within Cycle (Gibbons et al., 1996). Factors that oppose this hypothesis are i)
Colombia should also be tested. Northward migration of terranes during the timing of northward displacement proposed by Bayona et al.
the late Jurassic could also account for ~63° of clockwise rotation of the (2010) is based around K/Ar dates, which rarely record crystallisation
Amotape Terrane (southern Ecuador; Fig. 1), which started at ~115 Ma ages and probably do not relate to the time that the remnant mag-
(Mourier et al., 1988; Jaillard et al., 1999). Clockwise rotation would netisation 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
30 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

Guamote Sequence (Ecuador) Quebradagrande Complex (Colombia) Upano Unit (Ecuador)


30 1 sandstone 1 sandstone
3 sandstones 7

Silvia-Pijao Fault
Youngest single zircon: 6 Youngest single zircon: Youngest single zircon:
149.2 ± 6.1 Ma, N = 50 6 143.3 ± 9.9 Ma
155 ± 6.1 Ma, N = 262 5
20 5 N = 59
4
N N N4
3
3
10 2
2
1 1
0 0
0
1200
1500
1800

1200
1500
1800
0
300
600
900

2100
2400
2700
0
300
600
900

2100
2400
2700

Llanganates Fault
Alao Arc (Ecuador)
Age (Ma) 14 Age (Ma)
1 sandstone
12 Youngest single zircon:

Peltetec Fault
163.7 ± 1.6 Ma, N = 58
10
N 8
6
4
2

1200
1500
1800
0
300
600
900

2100
2400
2700
3000
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 5.5. The tectonic setting during the Early Cretaceous (145–115 Ma)
young to the west, and the Jurassic arc is not duplicated in Colombia
or Ecuador (Fig. 11b). Consequently, we conclude that the Jurassic arc The compiled zircon U-Pb dates, geochemistry and isotopic data
in Northern Peru either never formed, or has been removed by subduc- clearly show that I-type, metaluminous, high-SiO2 (N 75%; Fig. 15) arc
tion erosion. rocks formed the Azafrán and Chinguál Batholiths, and parts of what is

40Ar/39Ar date
36Ar/40Ar
±2 (Ma)
1400 Inverse isochron age 115.35 ± 43.66 Ma
Metabasalt 09PR47 0.04
40Ar/36Ar intercept 319.7 ± 57.2
1200
MSWD 2.47 step used to calculate age
1000 0.03
step not used
800
0.02
600
400 134.26±12.84 (4 steps)
0.01
200

0 0.00
0 20 40 60 80 100 0.00 0.02 0.04 0.06
%39Ar released 39Ar/40Ar

40Ar/39Ar date
±2 (Ma) 36Ar/40Ar

Inverse isochron age 134.12 ± 8.65 Ma


Metagabbro 09PR48 0.04 40Ar/36Ar intercept 298.4 ± 42.8
400
MSWD 2.25
0.03
300

134.71±0.89 (5 steps) 0.02


200

0.01
100
0 20 40 60 80 100
%39Ar released 0.00
0.00 0.02 0.04 0.06
39Ar/40Ar

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 defined by
gas that is contaminated with excess 40Ar (step heating data are presented in the supplementary files).

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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 31

mapped as the Zamora Batholith (granodiorite 09RC43; Table 2; Fig. 19) intrusions (Fig. 16). Scant geochronological data suggest that the
during 144–132 Ma. These units are exposed to the west of the older arc volcanic rocks within the Quebradagrande Unit erupted at ~ 114 Ma,
intrusions in Ecuador, are bound against the Cosanga Fault (Fig. 11) while the volcanoclastic rocks were deposited after ~149 Ma (Fig. 13).
to their east, and exhibit a weak to strong foliation (Pratt et al., 2005). No accurate age estimates of the Alao arc are available although sedi-
The Hf and Nd isotopic compositions (Fig. 16) indicate that they mentation could have occurred after ~ 164 Ma (Fig. 13). K/Ar dates of
crystallised from melts that were more juvenile than the older, un- 115 ± 12 Ma and 142 ± 36 Ma suggest that it could be coeval with
foliated Jurassic intrusions and they are considered by Cochrane the Quebradagrande Complex. This tectonic correlation is supported
(2013) to have erupted through thinned crust (Fig. 19). This interpreta- by its outboard position relative to the Palaeozoic basement, Triassic
tion is consistent with Pratt et al. (2005) who suggest that the foliations anatectites and Jurassic intrusions, and we tentatively assign an Early
formed during pure shear, and Litherland et al. (1994) who suggest that Cretaceous age to the Alao arc.
magmatism within the “Salado Terrane” occurred within a marginal The Alao Arc hosts large volumes of quartz-zircon-tourmaline
basin (Fig. 19). U-Pb dates of detrital zircons sandstones of the Upano rich arenites (Cochrane, 2013), and the U-Pb ages of detrital zircons
Unit suggest that this marginal basin was forming after 143 Ma (Fig. 13) reveal a derivation from cratonic South America. Therefore,
(Fig. 13), which is consistent with the crystallisation ages of the Azafrán the Alao Arc is interpreted to have formed above an east-dipping sub-
and Chinguál plutons. Extension during 144–132 Ma coincides with a duction zone along the thinned fringe of a continental margin, giving
region-wide angular unconformity in the Upper Magdallena Valley rise to isotopically juvenile, mafic volcanic rocks (Cochrane, 2013;
Basin (Fig. 1; Jaimes and de Freitas, 2006), the Oriente Basin in Fig. 19). Extension was sufficient to form MORB-like rocks with tholeiit-
Ecuador (Balkwill et al., 1995) and in northern Peru (Jaillard et al., ic signatures above the subduction zone. Alternatively, the tholeiitic ba-
1990). The older Jurassic intrusions to the east (N 145 Ma; e.g. the salts may form part of the Peltetec Unit (Fig. 11), and were structurally
Abitagua Batholith) were not affected by the extensional system, and emplaced against the calk-alkaline rocks during subsequent compres-
remain unfoliated. sion. This interpretation differs from that of Litherland et al. (1994),
Mafic, calk-alkaline to tholeiitic volcanic rocks of the Alao and who draw the Alao Arc as a Middle Jurassic Island Arc, although this is
Quebradagrande Complexes are preserved to the west of the Azafrán inconsistent with the presence of abundant Precambrian zircons. Simi-
and Chinguál Batholiths, and west of an uplifted core (Pratt et al., larly, the volcanic rocks of the Quebradagrande Complex are associated
2005) of Palaeozoic basement and Triassic migmatites, via the Baños with sedimentary rocks (Abejorral Fm.; E.g. Gómez-Cruz et al., 1995)
Fault (Fig. 11). Major, trace element and REE compositions suggest that host abundant quartz, and Nivia et al. (2006) interpreted this
that these rocks formed within arcs (Fig. 17), although some samples sequence as a thinned marginal basin along a continental margin
yield an affinity with MORB. The volcanic rocks are located outboard (Fig. 19). Villagómez et al. (2011) suggest that the arc rocks erupted
of the Jurassic intrusions (Fig. 11b), and yield juvenile Hf and Nd isotopic through highly attenuated continental crust because i) some basalts
compositions and depleted trace elements relative to the Jurassic yield geochemical signatures that approach seamounts (T-MORB),

A C
4.0 10000
Ecuador (Abitagua, Zamora, Rosa Florida)
Peru (Ilo Batholith) Peraluminous
3.5 Jurassic arc Ecuador (Azafrán, Chinguál)
1000
Colombia (Ibagué, Segovia)
3.0
Colombia (Santander Massif)
Metalluminous
Al/(Na+K)

100
2.5
Ecuador, Colombia
2.0 Triassic anatec tes 10
1.5
1
1.0
Peru (Allincapac) Jurassic back-arc
0.5
0.5 1.0 1.5 2.0 2.5 3.0 0.1
Cs Rb Ba Th U Nb Ta La Ce Pb Pr Sr Nd Zr Hf Sm Eu Ti Tb Y Tm Yb
Al/(Ca+Na+K) D
B 100
Azafrán, Chinguál Batholith
Rosa Florida, Abitagua, Zamora
10
Ibagué, Segovia
10
Na2O + K2O - CaO (wt%)

8 Peru (Allincapac)
Jurassic back-arc

6
1
4
Alkalic
2 c
lci 0.1
i- ca La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Yb Lu
k al lin
e
0 Al ka
al
lc- Calcic
Peru (Ilo Batholith)
-2 Ca Jurassic arc

50 55 60 65 70 75 80
SiO2 (wt%)

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 Mišković 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
32 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

A B
εNdi (whole rock)
Th (ppm) 6
100 Gz
4

10 2 Gz
High-K calc-alkaline

0
Calc-alkaline
1
-2
Rhyo e, grano
gran

Peru
Coastal Batholith
Bas site, dio
and
lite,
it

-4 Arequipa Terrane
altic
e

0.1
dacit iorite

and ite

Peru
-6
e,

Tholeiite Arequipa Terrane


d

esite
r

0.01 -8
,

0 20 40 60 80 100 120 140 160 180 200


Co (ppm)
Time (Ma)
C D
εHfi (zircon) (La/Yb)n
16 35
DM
14
12 Peru 30
Arequipa Terrane
10 and back-arc
25
8
6 20
4
2 15
0
10
-2 Peru
-4 Coastal Batholith
Arequipa Terrane and 5
-6 North of Arequipa
-8 0
100 120 140 160 180 200 220 -10 -5 0 5 10 15
εNdi (whole rock)
Time (Ma)

Cretaceous ultramafic - mafic complexes


Peltetec (Ecuador)
Raspas (Ecuador)
Arquía Unit (Colombia)
Barragán Unit (Colombia)
Cretaceous Volcanic Rocks
Alao Arc (Ecuador)
Quebradagrande Complex (Colombia)
Cretaceous Intrusions
Eastern Cordillera (Colombia)
Jurassic Intrusions
Azafrán and Chinguál (Ecuador)
Zamora, Abitagua (Ecuador)
Ibagué, Segovia (Colombia)
Santander Massif (Colombia)

Fig. 16. Geochemical and isotopic data from Jurassic intrusions, Early Cretaceous volcanic rocks, and M-HP/LT mafic and ultramafic 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 Mišković 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 33

ii) most volcanic rocks erupted in submarine conditions, and iii) no con- a continental arc that erupted through thinned continental crust, and
tinental detritus is found to the west of the Quebradagrande Complex. iii) extension prevailed during the Early Cretaceous. No evidence exists
Pindell and Kennan (2009) draw the Quebradagrande Arc as an oceanic for a Jurassic oceanic Plateau within Colombia or Ecuador.
arc above an east dipping subduction zone until 125 Ma, although this The Arquía, Barragán, and Raspas complexes are considered to be
date is not based on robust geochronological evidence from north- equivalent because i) they are located in similar structural positions,
western South America. outboard of Triassic anatectites and volcanic arc rocks, ii) they are com-
The simplest explanation for the geochemical, isotopic and geochro- posed of basalts and gabbros that have been metamorphosed to varying
nological trends obtained from the Jurassic and Early Cretaceous igne- degrees under M-HP/LT conditions, iii) they yield peak metamorphic
ous rocks is that they formed above the same east-dipping subduction and retrogression dates that are similar, and iv) they yield MORB–
zone, which was retreating oceanward during 209–194 Ma, and after seamount geochemical signatures. The Raspas Complex now forms
~ 145 Ma until ~ 114 Ma (E.g. Cochrane, 2013; Fig. 19). Extension of part of the para-autochthonous Amotape Terrane, which detached
the continental crust during ~ 145–114 Ma was sufficient to generate from rocks that are now exposed in the Cordillera Real. The Raspas Com-
mafic magmas with T-MORB geochemical characteristics and marine plex hosts HP-LT rocks of oceanic plateau and MORB affinity (Arculus
environments, which are found intercalated within the Alao and et al., 1999; Bosch et al., 2002; John et al., 2010), which were exhumed
Quebradagrande sequences. Kennan and Pindell (2009) refer to this following peak metamorphism at 130–126 Ma (John et al., 2010).
extensional feature as the Colombian Marginal Seaway. This time corre- Geochemical and isotopic compositions also suggest that the Barragán
sponds with heating of some fault blocks within the southern Cordillera (HP-LT) and Arquía (MP-LT) units originated as MORB and seamounts
Real (Fig. 7a) to temperatures of up to ~500 °C, which was synchronous (Bustamante et al., 2012), while their 40Ar/39Ar dates suggest that
with a significant increase in εHFi (zircon) in the magmatic rocks. they were exhuming and cooled below ~ 400 °C during 120–112 Ma.
Heating is interpreted to be a consequence of sedimentary burial during We consider these units to have been metamorphosed within the
extension, combined with an increase in geothermal gradients, which same subduction zone, which was exhumed to variable degrees along-
may have been significant in places due to the proximity of magma. strike of northwestern South America (Fig. 19). Exhumation could
Other fault blocks (Fig. 7b; e.g. northern Colombia) were not being have been triggered by forced return flow (Gerya et al., 2002), which
heated at this time, presumably because they did not reside in a part may have occurred during a compressive event, combined with the
of the extensional system that was being buried. Finally, some faulted inherent buoyancy of oceanic plateau rocks relative to MOR-derived
units (e.g. south of the Ibagué Fault in the Cordillera Real; Villagómez lithosphere.
and Spikings, 2013) were cooling at ~140 Ma (Fig. 7c), perhaps because Rifting of northwestern South America during the Early Cretaceous
they were exhumed during extension. may have been sufficient to detach continental slivers (e.g. the Chaucha
This interpretation is similar to that of Toussaint and Restrepo Terrane in Ecuador), and we suggest that the greenschist facies ultra-
(1994), Cooper et al. (1995), Sarmiento-Rojas et al. (2006) and Pindell mafic–mafic rocks of the Peltetec Complex formed during advanced
and Kennan (2009), who suggest that the rocks of the Eastern Cordillera continental rifting as E-MORB crust. The Peltetec Complex has not
of Colombia (Fig. 1) underwent back-arc extension (NNE–SSW exten- been metamorphosed in a subduction zone, and its current structural
sional axis; e.g. Mora et al., 2006), during the Cretaceous, associated position between the Guamote Sequence and the Alao Arc in Ecuador
with an arc that formed rocks which are now exposed within the Cordil- indicates they were obducted during a compressional event. Their
40
lera Central (Quebradagrande Complex). Vásquez and Altenberger Ar/39Ar dates suggest that they formed at ~ 134 Ma, and hence they
(2005) and Vásquez et al. (2010) report that gabbroic intrusions ex- formed before the subduction channel that hosted the Arquía, Barragán
posed in the Eastern Cordillera formed within a rift setting during and Raspas complexes was exhumed.
135 Ma and 121 Ma (Table 2; Fig. 11). The MORB and OIB-type chemis-
try of these rocks (Vásquez et al., 2010) suggests that back-arc extension
was significant. 5.6. Compression during the Early Cretaceous
The Guamote Sequence of Litherland et al. (1994) does not host any
magmatic rocks, and is composed of metasedimentary rocks that were Dense mineral assemblages and detrital zircon fission track dates
deposited after 155 Ma (Fig. 13). The Braziliano and Sunsas aged U-Pb with very short lag-times suggests that rocks deposited within the
ages of detrital zircons shows that the sediments were derived from Upper Magdalena Valley (Fig. 1; Vergara and Prössl, 1994; Sarmiento
cratonic South America. Furthermore, their detrital age signature is in- and Rangel, 2004) and Oriente (Ruiz et al., 2007; Martin-Gombojav
distinguishable from that obtained from arenites within the Alao Arc, and Winkler, 2008) basins during 120–115 Ma were derived from an
Quebradagrande Complex and the Upano Unit (Fig. 13). The Guamote exhuming cordillera located to the west. Restrepo et al. (2009) suggest
Sequence is currently separated from the Alao Arc by ultramafic–mafic that Aptian–middle Albian sandstones of the Abejorral Fm., which are
rocks of the Peltetec Complex, although it is not unreasonable to suggest located along the western flank of the Cordillera Central, were derived
that it once formed a part of the South American Margin, and was from a proto-Central Cordillera to the east. Time–Temperature paths
emplaced either by i) strike-slip displacement from more southern for some faulted blocks suggest that they cooled rapidly at 117–
latitudes (see Section 5.7), or by ii) rifting away from north-western 107 Ma (Fig. 7c), while other faulted units appear to have continued
South America, followed by re-accretion (Fig. 19). the heating trend from earlier in the Cretaceous (Fig. 7a). Villagomez
Litherland et al. (1994) suggest that prior to 140 Ma, continental et al. (2013) interpret these data as evidence for compression, which ex-
crust of the Chaucha Terrane lay outboard of a west-facing island arc humed some fault blocks, forming a proto-Cordillera Central. This time
(Alao Arc), which was separated from South America by oceanic crust period coincides with the timing of retrogression of the east-dipping
that was subducting beneath South America, forming the Azafrán subduction zone during exhumation, and compression at this time
Pluton. Subsequently, the same authors suggest that these terranes col- obducted blueschists and eclogites of the Raspas, Arquía and Barragán
lided together during 140–120 Ma, during the compressive Peltetec complexes onto the South American margin. Compression may have
Event. Similarly, Villagómez and Spikings (2013) suggest that the colli- occurred due to an increase in the convergence rates of the oceanic
sion of a series of seamounts or oceanic plateau blocked the Jurassic sub- Caribbean Plate (Kennan and Pindell, 2009) and South America, as a
duction zone, terminating the Jurassic arc, and that a new subduction consequence of the opening of the South Atlantic, which started at
zone formed outboard of the hypothetical plateau, forming the west- ~ 120 Ma (Eagles, 2007). Detrital zircon fission track dates from the
facing Quebradagrande Arc. However, the data compiled here suggest Oriente Basin of Ecuador (Fig. 1; Ruiz et al., 2004) suggest that a second
that i) arc magmatism during the Jurassic–Early Cretaceous formed compressive event may have occurred within Ecuador at ~ 100 Ma,
above a single subduction zone, ii) the Quebradagrande Complex was which would be coeval with compression in Peru (Mégard, 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
34 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

A B
8 La/Yb
Peraluminous
7 30
Barragán

Metalluminous
6
Al/(Na+K)
Triassic Jurassic Arquía
5 25 Batholiths Raspas
anatectes
4 Calc-alkaline Peltetec
Jurassic 20
3 arc arc Alao
2 Azafran Quebradagrande
15 Batholith
1
0 10
0 0.5 1.0 1.5 2.0 2.5
Island arc
Al/(Ca+Na+K)
C 5 tholeiite
Ocean plateau tholeiite
Nb/La
1.6 0
1.4 0 200 400 600 800
E-MORB seamounts Zr/Th
1.2 D
1.0
0.8 Quebradagrande Complex
MORB 1000
0.6 Alao Arc
connental arc
0.4
oceanic 100
0.2
arc
0
0 0.5 1.0 1.5 2.0 2.5 3.0 10
(La/Sm)n
E
10 1

N-MORB
0.1
Cs Rb Ba Th U Nb Ta La Ce Pb Pr Sr Nd Zr Hf Sm Eu Ti Tb Y Tm Yb
1

N-MORB
0.1
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Yb Lu
F G
Blueschist 10
Barragan Complex
Amphibolite
100
Arquía Complex Amphibolite
1
10

N-MORB
1 0.1
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Yb Lu

N-MORB
0.1
Cs Rb Ba Th U Nb Ta La Ce Pb Pr Sr Nd Zr Hf Sm Eu Ti Tb Y Tm Yb
H
Blueschist I
Raspas Complex 10
Eclogite
100
Peltetec Complex Basalt, gabbro

10 1

1 N-MORB
0.1
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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 35

Maresch et al. (2009) report metamorphic zircon dates of 116– 2011) and Triassic (Cajamarca Complex; e.g. Cochrane et al., 2014a)
106 Ma from anatectites within HP-LT metagabbros at La Rinconada, U-Pb dates, which are intruded by the ~ 90 Ma Antioquia Batholith.
Margarita Island (southern Caribbean). Those rocks are interpreted as However, in our opinion these are insufficient to assign a terrane
back-arc basin crust that was subducted and retrogressed during 116– status to these rocks, and it is not proven that they are derived from
106 Ma. The coincidence in the timing of metamorphism in northwest- several hundred km to the south, as is advocated in many models (e.g.
ern South America and Margarita Island suggests that they may have Kennan and Pindell, 2009). The simplest interpretation of the data pre-
formed in the same tectonic setting. However, Maresch et al. (2009) sented here is that the current west-to-east juxtaposition of M-HP/LT
suggest that the HP-LT rocks at Margarita Island formed on a west- rocks, an Early Cretaceous arc, Palaeozoic and Triassic basement, latest
dipping slab of continental back-arc crust. The same arc polarity is Jurassic arc and Jurassic arc, is a consequence of Jurassic–Early Creta-
shown by Pindell et al. (2005) at 119 Ma, although Pindell and ceous attenuation of the margin, followed by a switch to compression
Kennan (2009) suggest that the polarity of the subduction zone flipped starting at ~ 120–115 Ma. The assemblage of Palaeozoic and Triassic
during 125–120 Ma, during the transition from their Trans-American rocks in Colombia is identical in composition, field relationships
Arc, to the Caribbean Arc, and the separation of North America from and age to those found throughout the Loja Terrane (a term used by
Gondwana. However, it is difficult to account for the outboard position Litherland et al., 1994) of Ecuador, and they are interpreted here as
of M-HP/LT metamorphic rocks relative to the Alao and Quebrada- basement that was exhumed during Jurassic–Early Cretaceous exten-
grande arcs within an east-facing arc system, and the simplest explana- sion. Jurassic intrusive rocks are not found in these regions because
tion of the very large amount of data obtained from Jurassic and Early they did not overlie the arc axis at that time (Fig. 19). Furthermore, U-
Cretaceous rocks of the northern Andes suggest that subduction was Pb zircon ages spanning between 900 and 1700 Ma, with a peak at
east-dipping beneath South America. ~ 1200 Ma (orthogneiss, La Miel Unit; Villagómez et al., 2011) within
Ordovician rocks suggest that Precambrian rocks may underlie this
5.7. The Chaucha Terrane and the Tahamí Terrane suspect terrane. Consequently, we do not include these rocks as terranes
in our reconstruction (Figs. 18a and 19). Similar to the case of the
As previously described, the Chaucha Terrane includes the Guamote Chaucha Terrane, models which place the Jurassic subduction zone out-
sequence (Fig. 11), which hosts quartzites and slates that contain Juras- board of a hypothetical Tahamí Terrane (e.g. Pindell and Kennan, 2009;
sic ammonites, some of which were deposited after 155 ± 6.1 Ma Fig. 18b) create a distance between the trench and the Jurassic arc in
(Fig. 13). These rocks lie outboard of the Peltetec Fault, and probably South America that is too large.
represent a rifted fragment of South America, which re-accreted during
compression at 120–110 Ma, or perhaps at a later time. The N-S extent 5.8. Comparison with Peru (145–115 Ma)
of these units beneath the volcanic cover rocks is unknown, and other
suspect occurrences include i) a small inlier of slates within the north- With the exception of far northern Peru, no intrusive or volcanic
ern Interandean Depression in Ecuador (Fig. 1; near the town of rocks have been documented within Peru that yield zircon U-Pb
Chota, 0°15’N), ii) a faulted sliver of graphitic schists in the southern dates that lie between 145 and 115 Ma. The period between 145 and
Western Cordillera of Ecuador (2°30’S), and iii) undated granites of 110 Ma was characterised by extensional tectonic events along the
the Manu Inlier, located between the Amotape Terrane and the Eastern Peruvian margin, generating deep sedimentary basins in northern
Cordillera of Ecuador (3°30’S). The Chaucha Terrane is shown on the Peru during 145–130 Ma (Chicama Basin), and in Central Peru during
tectonic reconstructions of Litherland et al. (1994) as a large continental 130–110 Ma (Jaillard and Soler, 1996). Kennan and Pindell (2009)
block that is exotic to South America. Similarly, Pindell and Kennan draw highly oblique and sinistral plate convergence directions between
(2009) show the Chaucha Terrane on their reconstructions (Fig. 18b) South America and the Farallon Plate at 130 Ma, which would result in
as a continental microplate, with sufficient dimensions to displace the slow subduction north of the Huancabamba Deflection, and no net
Jurassic subduction zone westwards by N100 km. However, the distance convergence along coastal Peru, accounting for a lack of magmatism in
between their Jurassic subduction zone and South America (Fig. 18b) is that region. Pindell and Kennan (2009) draw orthogonal subduction
too large to account for the distribution of Jurassic arc rocks within between their Caribbean Plate and the Peruvian coastline at 125–120
Ecuador. We consider all of these exposures to be para-autochthonous Ma, although this is unlikely given the lack of arc rocks of this age.
to South America, and equivalent to lithologies exposed to the east of The Aptian calc-alkaline Celica continental arc (Feininger and
the Peltetec Fault. Their current dispersion over a strike-distance of Bristow, 1980; Lebrat et al., 1987) formed in northwestern Peru and
~ 300 km may be a result of strike-slip segmentation of several rifted far southwestern Ecuador, and its forearc Celica-Lancones Basin is pre-
continental slivers during the amalgamation of the Caribbean Large Ig- served (Fig. 1; Jaillard et al., 1999). The magmatic rocks are poorly
neous Province at ~ 75 Ma (Vallejo et al., 2006; Spikings et al., 2010), dated, and they may have formed above the same subduction zone
and we do not consider these rocks to belong to a separate terrane, that gave rise to the Quebradagrande and Alao arcs. These rocks may
and use of the term “Chaucha Terrane” is misleading. In our interpreta- represent the earliest manifestation of magmatism in Peru that lead to
tion, the latest Jurassic–Early Cretaceous Guamote sequence formed the formation of the Coastal Batholith starting at ~115 Ma.
within the forearc of the Alao Arc and Upano Units during extension,
above an east-dipping subduction zone. 6. The tectonic history of northwestern South America during
The Tahamí Terrane (Antioquia Terrane in Pindell and Kennan, 115–75 Ma
2009) of northern Colombia (Restrepo and Toussaint, 1982) is widely
considered to be fault bounded by the Otú-Pericos and Palestina faults Only small volumes of magmatism are recorded in Ecuador and
to the east, the San Jeronimo Fault to the west, and the Ibagué Fault in Colombia during ~115–100 Ma, which corroborates the highly oblique
the south (Fig. 1). It is considered to be a distinct terrane because i) it dextral convergence angles of the Caribbean Plate with South America
is separated from Sunsas-aged (~1 Ga) crust in the east, and ii) it does (Pindell and Kennan, 2009). Barragán et al. (2005) report plateau
not expose Jurassic arc rocks, unlike the region to the east. The base- (whole rock) 40Ar/39Ar dates of 110–82 Ma from low volumes of alkali
ment rocks in this suspect terrane yield Ordovician (Villagómez et al., 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; Villagómez et al., 2011; Cochrane, 2013; Rodriguez and Zapata, 2013) and
M-HP/LT mafic and ultramafic rocks (Litherland et al., 1994; Arculus et al., 2002; Bosch et al., 2002; John et al., 2010; Villagómez 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 fields are from Jolly et al. (2001), and the
Th-Co classification 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
36 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

A 209 - 194 Ma 189 - 145 Ma

North
America Yucatan Yucatan

Central and
20°N southern
Mexico
Santander
Massif
20°N
S

120mm/a
F
P-SP

120mm/a I
10°N
Arc exposed today
10°N Arc not formed, burried A
or tectonically eroded
Z
Present day coastline

0°N

Arequipa
0°N Terrane
N CC
IB

Nazca Nazca
300km Ridge Ridge
300km
100°W 90°W 100°W 90°W

B 190 Ma 158 Ma

North
America Yucatan Central and
Southern Yucatan
Mexico

Ch 20°N
20°N
120mm/a 120mm/a Ch
T

T
10°N
10°N C

0°N
0°N
1000 km 1000 km

110°W 100°W 90°W 110°W 100°W 90°W

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 189–145 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: Ibagué
Batholith, IB: Ilo Batholith, P-SPF: Peltetec-Silvia Pijao Fault (this is the Jurassic–Early Cretaceous palaeomargin; Vallejo et al., 2006), S: Segovia Batholith, Z: Zamora Batholith. B) Sub-
duction zones and reconstruction of Pindell and Kennan (2009), which assume that the Tahamí 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: Tahamí 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 37

189-145 Ma 145 - 141 Ma


Continental arc volcanism Extension of the continental margin

Jurassic Batholith Foliated


(E.g. Abitagua, Ibagué) arc (E.g. Azafrán Batholith)

W E
rollback

active
arc
active
arc Triassic anatectites,
undifferentiated
Palaeozoic rocks

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

Raspas Complex Quebradagrande


Guamote Peltetec Upano Arquía Complex
Alao Arc Complex
Sequence Unit Unit (marine) Barragán Complex
marine
rollback rollback

active active
arc asthenospheric arc asthenospheric
upwelling forming upwelling forming
transitional crust transitional crust

115-100 Ma Closure of fore/intra/back arc basins, obduction of M-HP/LT rocks, rock uplift and exhumation

Future Caribbean Peltetec Fault Zone Future Caribbean Silvia-Pijao


Cosanga Otú-Pericos
suture (75 Ma), suture (75 Ma), Fault
Fault Fault
Ingapirca Fault Cauca-Almaguer
marine Fault marine
compression compression

Fig. 19. Schematic models for the tectonic evolution of the northwestern South American margin, which fit the geochronological, geochemical and sedimentological data. These models
propose that the Jurassic arc axis during 185–145 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-mafic and mafic rocks of the Peltetec Unit between those arcs and re-accreted continental slivers.

suggest that the basalts formed during asthenospheric upwelling during the Albian. Extension is supported by gravity data, which reveal
caused by the detachment of an east-dipping slab subsequent to terrane an arch-like structure of dense material within a rift setting, which is
collision, implying that an east-dipping subduction zone existed be- considered to be imbricated oceanic crust (Jones, 1981). Jaillard and
tween an accreting terrane and South America. Our model suggests Soler (1996) report that extension was punctuated with compressional
that the long-lived Jurassic–Early Cretaceous subduction zone simply events, and that volcanic activity which fed the Casma Basin ceased by
retreated westwards, and was not “pinched-off” between two crustal the Late Albian, during the compressional Mochica Phase of Mégard
blocks (Fig. 19). We suggest that if any asthenospheric upwelling oc- (1984). Abundant subduction related magmatic activity along Peru
curred, it was driven by steepening of the slab during its prolonged res- starting at ~ 115 Ma is accounted for in the reconstructions of Pindell
idence in the mantle in the absence of subduction after ~110 Ma, along a and Kennan (2009) by the north-east directed subduction of their
highly oblique margin. newly formed Caribbean Plate beneath South America, forming a
Plutons of the Coastal Batholith in Central Peru and coastal Arequipa mainly strike-slip boundary north of the Huancabamba Deflection,
started intruding at ~115 Ma. The magmatic units along the coastline of and amagmatic tectonics in the northern Andes.
central Peru are considered to have formed by shallow melting of Acidic magmatism in the Northern Andes between 100 and 75 Ma is
mantle-derived basalts (Atherton, 1990; Atherton and Petford, 1996), dominated by the large Antioquia Batholith (Fig. 10; concordant zircon
which were originally derived from partial melting in the mantle U-Pb ages of 95–85 Ma; Villagómez et al., 2011; Fig. 20) in the northern
wedge above a subduction zone (e.g. McCourt, 1981). The intrusions Cordillera Central, which intrudes through undifferentiated Palaeozoic
crystallised within a deep (~10 km) marginal basin that formed during rocks and the Triassic Cajamarca Fm. Other scattered continental arc
continental margin extension over an along-strike distance of ~1600 km intrusions that yield zircon U-Pb ages in this age range include the
(Atherton and Petford, 1996), which formed the Casma marginal basin Córdoba Batholith (80 Ma; Villagómez 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
38 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

Central, and the Tangula Batholith (zircon U-Pb 92 ± 1 Ma; These rocks are equivalent to the Aruba Tonalite (Island of Aruba; zircon
Schütte et al., 2010) of the southern Cordillera Real, which intrudes U-Pb ages range between 89 and 87 Ma; Wright and Wyld, 2004; Van
continental crust. Villagómez et al. (2011) and Villagómez and der Lelij et al., 2010), which intrudes Turonian basalts that erupted
Spikings (2013) utilise geochemical data to suggest that these intru- above an oceanic hot-spot (White et al., 1999). The Rio Cala Group
sions formed by minor subduction of oceanic crust beneath South (Fig. 20; Vallejo et al., 2009) stratigraphically (high-Mg andesites,
America. Pindell and Kennan (2009) utilise plate reconstructions to basalts, and turbidites) overlies the oceanic plateau within Ecuador,
suggest that the Antioquia Batholith formed too close to the ocean- and its intra-oceanic character has been determined by isotopic and
continent boundary (b 100 km) to be arc magmas, and instead geochemical studies (Cosma et al., 1998; Mamberti, 2001; Mamberti
formed by melting of a slab-tip during subduction initiation, et al., 2003; Allibon et al., 2005; Vallejo et al., 2006), combined with
although they acknowledge that this is speculative. dense mineral assemblages that reveal no input from differentiated con-
tinental crust (Hughes et al., 1998; Vallejo et al., 2009). No U-Pb dates
6.1. The formation of the Caribbean Large Igneous Province and its collision have been obtained from the igneous rocks of the Rio Cala Group, al-
with South America. though the basal La Portada Fm. yields Santonian (85.89–83.5 Ma;
Gradstein et al., 2004) biostratigraphic ages (Kerr et al., 2002). Similarly,
6.1.1. Geochemistry and geochronology Campanian radiolaria have been found intercalated within island arc
The basement of the forearcs and Western Cordillera of Ecuador and lavas of the Ricaurte Arc in southern Colombia (Western Cordillera;
Colombia consists of ultramafic and mafic rocks, and their chemical Spadea et al., 1989). The association of Santonian–Maastrichtian intra-
compositions suggest that they formed above an oceanic hot-spot oceanic island-arc rocks overlying oceanic hot-spot derived rocks is doc-
(Fig. 1; Reynaud et al., 1999; Lapierre et al., 2000; Hughes and umented throughout the circum-Caribbean region (e.g. Frost and Snoke,
Pilatasig, 2002; Kerr et al., 2002; Mamberti et al., 2003, 2004; Kerr and 1989; Donnelly et al., 1990).
Tarney, 2005; Hastie and Kerr, 2010). Within Ecuador, the bulk of the The similar geochemistry and dates obtained from the Western
plateau rocks are represented by the Piñon (forearc), Pallatanga and Cordilleras and the forearcs suggest that these rocks were derived
San Juan (Western Cordillera) formations, and in Colombia the same li- from a single terrane (e.g. Luzieux et al., 2006). We refer to the accreted
thologies have been given numerous names, although here we will use allochthons within the forearcs and Western Cordilleras of Ecuador and
the terms Volcanic Formation for the mafic rocks of the Western Cordil- Colombia as the Caribbean Large Igneous Province (CLIP), which in-
lera and forearc (Kerr et al., 1997), and Amaime Fm. for the basalts cludes the equivalent Pallatanga-Piñon and Calima terranes (Figs. 1
exposed within the Cauca-Patía Valley (Fig. 1). No geochemical differ- and 20). The basaltic basement of the Chocó-Panamá Terrane located
ences are found between the basement of the forearc and the Western north of the Garrapatas Fault (Fig. 1) is younger and accreted to South
Cordillera. America during the Miocene, and thus it is not included in this
Radiometric data from the San Juan Fm. include a Sm-Nd isochron discussion.
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 6.1.2. Time of initial accretion with South America
(Mamberti et al., 2004), and a concordant U-Pb zircon date of 87.10 ± Estimates of the timing of accretion of the CLIP onto northwestern
1.66 Ma (Vallejo et al., 2006). The U-shaped 40Ar/39Ar age spectrum re- South America are 85–65 Ma (Aspden et al., 1987; Lebrat et al., 1987;
veals the presence of excess 40Ar, and this date is discarded. The discrep- Kerr et al., 2002; Spikings et al., 2005), or 75–65 Ma (Spikings et al.,
ancy between the Sm/Nd and U-Pb dates suggests that the densely 2001; Hughes and Pilatasig, 2002; Jaillard et al., 2004; Luzieux
faulted San Juan Unit is mapped incorrectly, and it includes intercalated et al., 2006; Vallejo et al., 2006; Spikings et al., 2010; Van der Lelij
but unrelated sequences. Jaillard et al. (2004) suggest that the older et al., 2010; Villagómez and Spikings, 2013).
Sm/Nd date is from an Early Cretaceous, allochthonous oceanic plateau, Thermochronological analyses of the Cordillera Real and the Cordillera
and they name it the San Juan-Multitud Terrane. We suggest that the Central reveal the onset of extremely rapid cooling at 75–73 Ma (Fig. 7b,
foliated gabbros that yield the older Sm/Nd age are a detached fragment c; Spikings et al., 2001, 2010; Villagómez and Spikings, 2013). Cooling is
of the anastomosed Peltetec Complex. Luzieux et al. (2005) report a pla- interpreted to be due to exhumation at rates of 1–1.6 km/My during
teau hornblende 40Ar/39Ar from the Piñon Fm. of 88.0 ± 1.6 Ma, suggest- 75–65 Ma, and high cooling rates continued until ~ 55 Ma. Recent
ing that it is equivalent to the basement of the Western Cordillera. (Cochrane et al., 2014b) and new (this study) t–T paths generated from
Amphibolites of the Totoras Amphibolite in the central Western apatite U-Pb data reveal the onset of rapid cooling at 80–75 Ma
Cordillera of Ecuador formed by metamorphism of an oceanic plateau (Fig. 7a). Palaeomagnetic data from the Piñon and San Lorenzo blocks
at 800–850 °C and 6–9 kbar (Jaillard et al., 2004; Beaudon et al., 2005). of coastal Ecuador (Fig. 1) record 40–50° of clockwise rotation during
Vallejo et al. (2006) report a hornblende plateau 40Ar/39Ar date of 73–70 Ma (Luzieux et al., 2006), which was synchronous with rapid
84.69 ± 2.23 Ma, and suggest that this records the timing of retrogres- exhumation of the same basement rocks located closer to South
sion through 550–500 °C, at the base of the oceanic plateau. Concordant America (Spikings et al., 2005).
zircon U-Pb ages (gabbros; Villagómez et al., 2011) and plateau 40Ar/39Ar Redbeds of the Tena Formation, located in the Subandean zone
dates of basaltic groundmass (Kerr et al., 1997) suggest that the oceanic (Fig. 1) of Ecuador are the oldest within the foreland basin to host
plateau rocks of Colombia crystallised during 100–92 Ma, which is significant quantities of metamorphic mineral grains derived from
slightly older than the ages from Ecuador. The crystalline ages and geo- high elevations to the west (Ruiz et al., 2004). Furthermore, fission-
chemical compositions reveal a strong affinity with the oceanic hot- track dates of detrital zircons within the Tena Fm. are indistinguish-
spot related rocks of the Caribbean Large igneous Province (e.g. Sinton able from their depositional ages, revealing extremely high exhuma-
et al., 1998), and they are considered to be detached allochthons of tion rates in the Cordillera Real. At the same time, there is a reduction
that province (e.g. Kerr et al., 1997). in the detrital supply from cratonic South America to the east. Within
Field studies and geochemical data show that the oceanic plateau the forearc of Ecuador, the Late Campanian–Maastrichtian (70–
rocks were intruded by a primitive arc sequence (Fig. 20). The oldest 65 Ma; Gradstein et al., 2004) Yunguilla Fm. was being deposited in
of these arc units are acidic intrusions of the Pujilí Granite (Ecuador; zir- a basin floored by oceanic crust of either the Pallatanga Unit
con U-Pb 85.5 ± 1.4; Vallejo et al., 2006), the Buga Granite (Colombia; (Hughes and Pilatasig, 2002), or the San Juan-Multitud Terrane
zircon U-Pb 90.6 ± 1.3 Ma–92.1 ± 1.3 Ma; Villagómez et al., 2011), (Jaillard et al., 2004). The turbidites were sourced from metamorphic
the Santa Fe Tonalite (Colombia; Sm-Nd age 98.0 ± 9.1 Ma; Weber rocks of the Eastern Cordillera (Vallejo et al., 2009). Jaillard et al.
et al., 2011), and the Altamira Gabbro (zircon U-Pb 88.9 ± 1.5; Zapata (2004) described coeval quartz-free pelagic cherts, and quartz rich
et al., 2011), all of which intrude the hot-spot related basement basalts. 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 39

Ecuador. Both of these sequences overlie oceanic mafic rocks, and 7. Conclusions
Jaillard et al. (2004) suggest that they define two separate terranes
that accreted before 71 Ma, and during 69–65 Ma. Within 1. Geochemical and isotopic analyses suggest that high-temperature
Colombia, rapid cooling and exhumation during 75–65 Ma was syn- metamorphism within Central American terranes (e.g. Maya
chronous with the deposition of the siliciclastic, El Cobre, Block) at 250 Ma occurred during compression driven by terrane
Monserrate, La Tabla and Cimarrona Fms. during the Campanian– accretion (e.g. Weber et al., 2007) along central Western Pangaea.
Maastrichtian in the retro-foreland Magdalena Valley (Villamil, Extension prevailed along the Peruvian margin at 250 Ma, and
1999), and the Nogales Fm. in the forearc, which hosts a large pro- thus it is likely that no continental terranes lay outboard of Peru at
portion of metamorphic detritus derived from the Cordillera Central that time.
(Moreno and Pardo, 2003). 2. Magmatic underplating and anatexis of continental crust during
These data were used by Luzieux et al. (2006), Vallejo et al. 240–225 Ma occurred during progressive thinning of the continen-
(2006), Spikings et al. (2010) and Villagomez et al. (2013) to suggest tal lithosphere during rifting along western Pangaea. Rifting ad-
that mafic rocks that originated above an oceanic hot-spot at 99– vanced to complete separation of continental crust by ~ 216 Ma,
87 Ma formed a single terrane, which along with its overlying arc ac- and the formation of oceanic lithosphere between the conjugate
creted against the margin of northwestern South America at ~ 75 Ma margins of northwestern South America and basement terranes of
(Fig. 20). This interpretation implies that the intra-oceanic arc (E.g. Central America (e.g. Oaxaquia). The rifting event is recorded by
Rio Cala Group) formed by west-dipping subduction beneath the amphibolitised tholeiitic basaltic dykes and extensive tracts of
buoyant hot-spot derived rocks, prior to its collision with South migmatites and S-type granites within the conjugate margins. The
America, which is consistent with the models of Burke (1988), Kerr rift axis propagated southwards, and extension is recorded along
et al. (1997), Spikings et al. (2001), Van der Lelij et al. (2010), and western Peru (the Mitu Aulocagen), Bolivia, western Argentina,
Zapata et al. (2011). However, this is inconsistent with the recon- Chile and southern Brazil. Rifting along northwestern South
struction of Pindell and Kennan (2009), who draw an east dipping America started as a back-arc basin to a Permian arc, and represents
subduction zone at 84 Ma beneath South America. The reconstruc- the early break-up of western Pangaea, leading to the separation of
tions of Pindell and Kennan (2009) suggest that the Rio Cala Arc is North and South America by ~180 Ma.
a continental arc, which is inconsistent with geochemical, isotopic 3. Metaluminous, I-type arc magmatism commenced in northwestern
and sedimentological data. Estimates of collision at ~ 85 Ma (e.g. South America at ~209 Ma, due to east-dipping subduction of the
Kerr et al., 2002; Spikings et al., 2005) were based around a mis- Farallon Plate. The arc axis migrated oceanward at some time dur-
interpretation of thermochronological data. ing 194–189 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
6.1.3. The nature of the CLIP–South America suture the Peruvian margin (~ 216–135 Ma) also started to migrate
Within Ecuador, the suture between the 90 and 87 Ma CLIP and oceanward at ~175 Ma, resulting in coeval arc and back-arc rocks.
the pre-existing continental margin is mainly obscured beneath Arc migration is considered to be a result of slab-retreat along the
volcanic rocks and intermontane basins within the Interandean De- western margin of South America, which caused the continental
pression (Fig. 1). The suture may be represented by the Pujilí Fault margin to extend, thinning the continental crust and generating
zone, which exposes the serpentinized Pujilí Melange, whose matrix progressively more isotopically juvenile arcs. The onset of latest
is the Pallatanga Unit. Two gabbroic inliers (E.g. Guayllabamba Triassic–Jurassic subduction beneath Colombia and Ecuador may
Basin) occur within the Interandean Depression, although the ages young towards the south, although this trend cannot be extended
of these rocks are unknown. Foliated metamorphic crust erupts as towards southern Peru, across the Huancabamba Deflection.
xenoliths within volcanoes along the western side of the Inter- 4. The present-day gap in Jurassic arc rocks north of the Arequipa
andean Depression (Bruet, 1987), suggesting that the basement to Terrane in Peru is considered to be a result of tectonic erosion
the valley may be a complex assemblage of dissected Early Creta- (Clift et al., 2003), and thus we propose that an arc did form in
ceous margin, and accreted CLIP. The Arquía Complex in Colombia that region. We suggest that the alternative hypothesis that the
is faulted against the Amaime Fm., which floors the Cauca-Patía arc was tectonically displaced northwards and now forms part of
Valley (Fig. 1) across the Cauca-Almaguer Fault. According to our Colombia (Bayona et al., 2010) is unlikely because the Jurassic arc
model (Fig. 19), the Cauca-Almaguer Fault is equivalent to the in Colombia and Ecuador is not temporally duplicated.
Ingapirca Fault (Litherland et al., 1994; Fig. 1) in Ecuador (the west- 5. Trench retreat of the east dipping-subduction zone accelerated
ern margin of the Guamote sequence). No reliable age dates have along northwestern South America at ~ 144 Ma, and extension
been obtained from the basalts of the Amaime Fm., although their during 144–115 Ma formed syn-tectonic granitoid intrusions with-
chemical composition suggests that they formed above an oceanic in Ecuador, attenuated the continental margin forming thin intra-
hot-spot (Kerr et al., 1997), and are identical to the Volcanic, arc basins characterised by transitional crust, and resulted in an
Pallatanga and Piñon fms. The nature of the contact between the oceanward migration of the arc axes, which became progressively
Buga Batholith (90.6 ± 1.3 Ma–92.1 ± 1.3 Ma; Villagómez et al., more isotopically juvenile and geochemically depleted. Arc rocks
2011) and the Amaime Fm. is uncertain, although it may be intrusive, of the Quebradagrande Complex and Alao arc erupted through
in which case the Amaime Fm. is older than ~ 91 Ma. We suggest that thin continental crust during the Early Cretaceous within a marine
these rocks form part of the CLIP, and the suture in Colombia is the environment. Rapid extension may have rifted some narrow con-
Cauca-Almaguer Fault. tinental slivers (e.g. the Guamote Sequence) from the margin.
Spikings and Crowhurst (2004) and Winkler et al. (2005) show that Back-arc magmatism is sporadically preserved within the Eastern
the Interandean Valley formed in Ecuador at ~ 6 Ma within a Cordillera of Colombia (136–121 Ma). Highly oblique and sinistral
dextral transcurrent setting (e.g. Winter and Lavenú, 1989), im- convergence directions between the Peruvian margin and the
plying that the accreted CLIP was closer to the Early Cretaceous Farallon Plate lead to a magmatic gap in Peru during ~135–115 Ma.
palaeomargin before ~ 6 Ma. We suggest that the CLIP-South 6. The distribution and composition of sedimentary rocks, combined
America suture, which started forming at ~ 75 Ma, now exists as with detrital thermochronology suggests that the margin of north-
a complex melange zone that forms the basement of the Inter- western South America was placed under compression at ~115 Ma.
andean Depression, and is perhaps more intact as the Cauca- Compression of the attenuated, weak, hot crust juxtaposed arc
Almaguer Fault within Colombia. 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
40 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

80° 70° 60° 50° 40° 80° 70° 60° 50° 40°

Caribbean
Plate
CA/NA Rio Cala 10°
10° Ar Arc
Proto- Oceanic
CA/NA Caribbean plateau B A

? P

illera
Oceanic 0° Mainly amagmatic

plateau within South America

-cord
CA/SA

Proto
T

10° 10°
CA/SA South America
?
South America

c o tho
c o tho

ba
as lith
ba
as lith

95-85 Ma

tal
100 Ma
tal

20° 20°

90° 80° 70° 60° 50° Late Cretaceous arcs


A: Antioquia Batholith, Ar: Aruba Batholith,
20° B: Buga Granodiorite, P: Pujilí Granite,
T: Tangula Batholith
Caribbean Early Cretaceous subduction zone
Plate CA/NA M-HP/LT rocks (Arquía, Barragán, Raspas complexes)
Oceanic Ar 10° Peltetec Complex
plateau Early Cretaceous arcs
CA/SA

A Northeast facing
C
PP

B
0° West facing, continental
P (Quebradagrande Complex, Alao Arc)
South America
T
co tho

75-70 Ma
ba
as lith

10°
tal

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, modified after Villagómez 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 (Müller et al., 1993). The
reconstruction draws the east-facing Rio Cala Arc forming above the oceanic plateau as it approaches South America. PPC: Piñon-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 and Pindell, 2009) resulted in only very minor intra-plate
faulted against the western margin of these compressed sequences, magmatism (Barragán et al., 2005) during 115–100 Ma, which
and represent a subduction channel that started to exhume from may have occurred during the steepening of an antiquated slab.
peak eclogitic conditions at 130–126 Ma. These eclogites and Net convergence with an orthogonal component along the
blueschists retrogressed through ~400 °C at 120–112 Ma, and it is Peruvian margin started forming the Coastal Batholith at ~115 Ma
likely that they were obducted onto the margin during compression within a marginal basin that formed during extension.
that started at ~115 Ma. These rocks originally formed parts of the 9. Oceanic plateau rocks of the Caribbean Large Igneous Province
same slab, and varying trench-parallel metamorphic facies reflect erupted through the Farallon Plate at near equatorial latitudes
exhumation from varying depths. Models which invoke west- (Luzieux et al., 2006) during 100–87 Ma, and these rocks migrated
dipping slabs after 125 Ma do not account for the spatial juxtaposi- approximately eastwards relative to South America. Subduction
tion of M-HP/LT rocks and their associated arcs. beneath northwestern South America during this period was
7. We suggest that there is very little evidence for the existence of restricted to northern Colombia and southern Ecuador, distal from
large allochthonous continental terranes (Tahamí and Chaucha) the leading edge of the approaching plateau. However, a majority
outboard of northwestern South America during the Jurassic– of the margin was amagmatic, and the intervening oceanic litho-
Early Cretaceous. These suspect terranes are not required to fit the sphere was consumed by west-dipping subduction beneath the
data presented here, and the geochemical, isotopic, sedimentologi- oceanic plateau, forming the Rio Cala intra-oceanic arc, and numer-
cal and thermochronological data can be accounted for by having a ous scattered intrusions. Collectively, the plateau and overlying arc
single east-dipping slab that is retreating oceanward at variable are referred to as the Caribbean Large Igneous Province.
rates during 189–115 Ma. 10. The Caribbean Large Igneous Province first collided with South
8. Highly oblique dextral convergence angles between the margin of America at ~ 75 Ma, resulting in the detachment and clockwise
northwestern South America and the Caribbean Plate (Kennan 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 41

forearc plains and the Western Cordillera. Other detached Boekhout, F., Sempere, T., Spikings, R., Schaltegger, U., 2013a. Late Paleozoic to Jurassic
chronostratigraphy of coastal southern Peru: temporal evolution of sedimentation
allochthons form the basement to the Leeward Antilles along the along an active margin. Journal of South American Earth Sciences 47, 179–200.
southern Caribbean plate boundary. Allochthons with South Boekhout, F., Roberts, N.M.W., Gerdes, A., Schaltegger, U., 2013b. A Hf-isotope perspective
America are sutured against the Early Cretaceous margin via the on continent formation in the south Peruvian Andes. The Geological Society of
London Special Publications 389. http://dx.doi.org/10.1144/SP389.6.
Cauca-Almaguer (Colombia) and Ingapirca (Ecuador) faults. Within Bosch, D., Gabriele, P., Lapierre, H., Malfere, J.-L., Jaillard, E., 2002. Geodynamic significance
Ecuador, the suture is represented by a melange, which is mainly of the Raspas Metamorphic Complex (SW Ecuador): geochemical and isotopic con-
buried beneath the Interandean Depression. straints. Tectonophysics 345, 83–102.
Botero, A., 1963. Contribución al conocimiento de la geología de la zona central de
Antíoquia. Anales Facultad de Minas (Medellín) 57, 101.
Bourgois, J., Toussaint, J.-F., Gonzales, H., Azema, J., Calle, B., Desmet, A., Murcia, L.A.,
Acknowledgements Acevedo, A.P., Parra, E., Tournon, J., 1987. Geological history of the Cretaceous
ophiolitic complexes of Northwestern South America (Colombian Andes).
Tectonophysics 143, 307–327.
We thank Arturo Egüez, Etienne Jaillard, Alfredo Buitron, Byron Bozkurt, E., Park, L.R.G., 1994. Southern Menderes Massif: an incipient metamorphic core
Pelicita and Luis Lopez for assistance in the field in the cordilleras of complex in western Anatolia, Turkey. Journal of the Geological Society 151, 213–216.
Ecuador, and Ecopetrol S.A., Andres Mora, Andreas Kammer, Agustin Bristow, C.R., 1973. Guide to the geology of the Cuenca Basin, southern Ecuador.
Ecuadorian Geological and Geophysical Society, Quito.
Cardona, Jaime Corredor, Jaime Castellanos, Wilson Casallas, and Burke, K., 1988. Tectonic evolution of the Caribbean. Annual Review Earth and Planetary
Luis Quiroz for their assistance during field work in Colombia. The Sciences 16, 201–230.
manuscript was improved by the thorough and helpful reviews of Victor 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 in-
Ramos, Maria Helbig and an anonymous reviewer. trusive rocks from the Garzón Massif, Upper Magdalena Valley and Central Cordillera,
southern Colombia. Boletin de Geología 32, 93–109.
Bustamante, A., Juliani, C., Hall, C.M., Essene, E.J., 2011. 40Ar/39Ar ages from blueschists of
Appendix A. Supplementary data the Jambaló region, Central Cordillera of Colombia: implications on the styles of ac-
cretion in the Northern Andes. Geologica Acta 9, 351–362.
Bustamante, A., Juliani, C., Essene, E.J., Hall, C.M., Hyppolito, T., 2012. Geochemical con-
Supplementary data to this article can be found online at http://dx.
straints on blueschist-facies rocks of the Central Cordillera of Colombia: the Andean
doi.org/10.1016/j.gr.2014.06.004. Barragán region. International Geology Review 54, 1013–1030.
Cameron, K.L., Lopez, R., Ortega-Gutiérrez, F., Solari, L.A., Keppie, J.D., Schulze, C., 2004. U-
Pb geochronology and Pb isotopic compositions of leached feldspars: constraints on
References the origin and evolution of Grenville rocks from eastern and southern Mexico. Geo-
logical Society of America Memoirs 197, 755–769.
Allibon, J., Monjoie, P., Lapierre, H., Jaillard, E., Bussy, F., Bosch, D., 2005. High Mg- Cardona, A., Cordani, U., Sanchez, A., 2007. Metamorphic, geochronological and con-
basalts in the Western Cordillera of Ecuador: evidence of plateau root melting straints from the pre-Permian basement of the eastern Peruvian (10°S): a Paleozoic
during Late Cretaceous arc magmatism, In: Sempéré, T. (Ed.), Proceedings of extensional–accretionary orogen? 20th Colloquium on Latin American Earth Sci-
the Sixth International Symposium on Andean Geodynamics, Program and Ab- ences, April 11–13, Kiel, Germany, pp. 29–30.
stracts, IRD ÉditionsUniversitat de Barcelona, Instituto Geológico y Minero de Cardona, A., Valencia, V., Garzón, A., Montes, C., Ojeda, G., Ruiz, J., Weber, M., 2010. Permian
España, Barcelona, Spain, p. 3335. to Triassic I to S-type magmatic switch in the northeast Sierra Nevada de Santa Marta
Alvarez, J.A., 1983. Geologia de la Cordillera Central y e1 occidente colombiano y and adjacent regions, Colombian Caribbean: tectonic setting and implications within
petroquimica de 10s intrusivos granitoides Mesocenozoicos. Boletin Geological Pangea paleogeography. Journal of South American Earth Sciences 29, 772–783.
INGEOMINAS, Bogota, Colombia, p. 26. Castro, A., Moreno-Ventas, I., Fernández, C., Vujovich, G., Gallastegui, G., Heredia, N.,
Arculus, R.J., Lapierre, H., Jaillard, E., 1999. Geochemical window into subduction and ac- Martino, R.D., Becchio, R., Corretgé, L.G., Díaz-Alvarado, J., García-Arias, M., Liu, D.-Y.
cretion processes: Raspas metamorphic complex, Ecuador. Geology 27, 547–550. , 2011. Petrology and SHRIMP U-Pb zircon geochronology of Cordilleran granitoids
Aspden, J.A., Litherland, M., 1992. The geology and Mesozoic collisional history of the Cor- of the Bariloche area, Argentina. Journal of South American Earth Sciences 32,
dillera Real, Ecuador. Tectonophysics 205, 187–204. 508–530.
Aspden, J.A., McCourt, W.J., Brook, M., 1987. Geometrical control of subduction-related Cawood, P.A., McCausland, P.J.A., Dunning, G.R., 2001. Opening Iapetus: constraints from
magmatism: the Mesozoic and Cenozoic plutonic history of Western Colombia. Jour- the Laurentian margin in Newfoundland. Geological Society of America Bulletin
nal of the Geological Society 144, 893–905. 113, 443–453.
Aspden, J.A., Bonilla, W., Duque, P., 1995. The El Oro metamorphic complex, Ecuador: ge- Centeno-Garcia, E., Keppie, J.D., 1999. Latest Paleozoic-early Mesozoic structures in the
ology and economic mineral deposits. Nottingham, British Geological Survey, Over- central Oaxaca Terrane of southern Mexico: deformation near a triple junction.
seas Geology and Mineral Resources, 67, p. 63. Tectonophysics 301, 231–242.
Atherton, M.P., 1990. The Coastal Batholith of Peru: the product of rapid recycling of ‘new’ Chamberlain, K.R., Schmitt, A.K., Swapp, S.M., Harrison, T.M., Swoboda-Colberg, N.,
crust formed within rifted continental margin. Geological Journal 25, 337–349. Bleeker, W., Peterson, T.D., Jeffferson, C.W., Khudoley, A.K., 2010. In situ U-Pb SIMS
Atherton, M.P., Petford, N., 1996. Plutonism and the growth of Andean Crust at 9°S from (IN-SIMS) micro-baddeleyite dating of mafic rocks: method with examples. Precam-
100 to 3 Ma. Journal of South American Earth Sciences 9, 1–9. brian Research 183, 379–387.
Bahlburg, H., Vervoort, J.D., Du Frane, S.A., Bock, B., Augustsson, C., Reimann, C., 2009. Chappell, B.W., White, A.J.R., 1974. Two contrasting granite types. Pacific Geology 8,
Timing of crust formation and recycling in accretionary orogens: insights learned 173–174.
from the western margin of South America. Earth-Science Reviews 97, 215–241. Cherniak, D.J., Lanford, W.A., Ryerson, F.J., 1991. Lead diffusion in apatite and zircon using
Baldock, M.W., 1982. Geology of Ecuador. Explanatory Bulletin of the national geological ion implantation and Rutherford Backscattering techniques. Geochimica et
map of the Republic of Ecuador, scale 1: 1,000,000 Direccion General de Geologia y Cosmochimica Acta 55, 1663–1673.
Minas, Quito, Ecuador, 11, p. 54. Chew, D.M., Schaltegger, U., Košler, J., Whitehouse, M.J., Gutjahr, M., Spikings, R.A.,
Balkwill, H.R., Paredes, F.I., Rodriguez, G., Almeida, J.P., 1995. Northern part of Oriente Miškovíc, A., 2007. U-Pb geochronologic evidence for the evolution of the
Basin, Ecuador: reflection seismic expression of structures. In: Tankard, A.J., Suarez, Gondwanan margin of the north-central Andes. Geological Society of America Bulle-
R., Welsink, H.J. (Eds.), Petroleum Basins of South America: American Association of tin 119, 697–711.
Petroleum Geologists, 62, pp. 559–571. Chew, D.M., Magna, T., Kirkland, C.L., Miskovic, A., Cardona, A., Spikings, R., Schaltegger, U.,
Barragán, R., Baby, P., Duncan, R., 2005. Cretaceous alkaline intra-plate magmatism in the 2008. Detrital zircon fingerprint of the Proto-Andes: evidence for a Neoproterozoic
Ecuadorian Oriente Basin: geochemical, geochronological and tectonic evidence. active margin? Precambrian Research 167, 186–200.
Earth and Planetary Science Letters 236, 670–690. Chiaradia, M., Vallance, J., Fontboté, L., Stein, H., Schaltegger, U., Coder, J., Richards, J.,
Barredo, S., Chemale, F., Marsicano, C., Ávila, J.N., Ottone, E.G., Ramos, V.A., 2012. Tectono- Villeneuve, M., Gendall, I., 2009. U-Pb, Re-Os and 40Ar/39Ar geochronology of the
sequence stratigraphy and U-Pb zircon ages of the Rincón Blanco Depocenter, north- Nambija Au-skarn and Pangui porphyry Cu deposits, Ecuador: implications for the Ju-
ern Cuyo Rift, Argentina. Gondwana Research 21, 624–636. rassic metallogenic belt of the northern Andes. Mineralium Deposita 44, 371–387.
Bayona, G., Jimenez, G., Silva, C., Cardona, A., Montes, C., Roncancio, J., Cordani, U., 2010. Clift, P.D., Hartley, A.J., 2007. Slow rates of subduction erosion and coastal underplating
Paleomagnetic data and K-Ar ages from Mesozoic units of the Santa Marta massif: a along the Andean margin of Chile and Peru. Geology 35, 503–506.
preliminary interpretation for block rotation and translations. Journal of South Clift, P.D., Pecher, I., Kukowski, N., Hampel, A., 2003. Tectonic erosion of the Peruvian
American Earth Sciences 29, 817–831. forearc, Lima Basin, by subduction and Nazca Ridge collision. Tectonics 22. http://
Beaudon, E., Martelat, J.E., Amortegui, A., Lapierre, H., Jaillard, E., 2005. Metabasites de la dx.doi.org/10.1029/2002TC001386.
cordillere occidentale d'Equateur, temoins du soubassement oceanique des Andes Cochrane, R., 2013. U-Pb thermochronology, geochronology and geochemistry of NW
d'Equateur. Comptes Rendus Geoscience 337, 625–634. South America: rift to drift transition, active margin dynamics and implications for
Beutel, E.K., 2009. Magmatic rifting of Pangaea linked to onset of South America plate mo- the volume balance of continents(PhD thesis) Terre & Environment, 118. University
tion. Tectonophysics 468, 149–157. of Geneva, Switzerland, p. 191.
Boekhout, F., Spikings, R., Sempere, T., Chiaradia, M., Ulianov, A., Schaltegger, U., 2012. Me- Cochrane, R., Spikings, R., Gerdes, A., Ulianov, A., Mora, A., Villagómez, D., Putlitz, B.,
sozoic arc magmatism along the southern Peruvian margin during Gondwana break- Chiaradia, M., 2014a. Permo-Triassic anatexis, continental rifting and the disassembly
up and dispersal. Lithos 146-147, 48–64. of western Pangaea. Lithos 190–191, 383–402.

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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
42 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

Cochrane, R., Spikings, R.A., Chew, D., Wotzlaw, J.-F., Chiaradia, M., Tyrell, S., Schaltegger, Goldsmith, R., Marvin, R.F., Mehnert, H.H., 1971. Radiometric ages in the Santander Mas-
U., Van der Lelij, R., 2014b. High temperature (N350 °C) thermochronology and sif, Eastern Colombia, Colombian Andes. Professional paper United States Geological
mechanisms of Pb loss in apatite. Geochimica et Cosmochimica Acta 127, 39–56. Survey 750-D, pp. 44–90.
Cocks, L.R.M., Torsvik, T.H., 2002. Earth geography from 500 to 400 million years ago: a Golonka, J., Bocharova, N.Y., 2000. Hot spot activity and the break-up of Pangea.
faunal and palaeomagnetic review. Journal of the Geological Society 159, 631–644. Palaeogeography Palaeoclimatology Palaeoecology 161, 49–69.
Collins, W.J., 2002. Hot orogens, tectonic switching, and creation of continental crust. Ge- Gómez, J., Nivia, A., Montes, N.E., Jimenez, D.M., Tejada, M.L., Sepulveda, J., Osorio, J.A.,
ology 30, 535–538. Gaona, T., Diederix, H., Uribe, H., Mora, M., 2007. Geological map of Colombia. Escala
Collins, W.J., Richards, S.W., 2008. Geodynamic significance of S-type granites in circum- 1:1'000.000. Ingeominas, 2nd Edition, Bogota.
Pacific orogens. Geology 36, 559–562. Gómez-Cruz, A.D.J., Moreno-Sánchez, M., Pardo, A., 1995. Edad y origen del “Complejo
Collins, W.J., Belousova, E.A., Kemp, A.I.S., Murphy, J.B., 2011. Two contrasting Phanerozoic metasedimentario Aranzazu-Manizales” en los alrededores de Manizales
orogenic systems revealed by hafnium isotope data. Nature Geoscience 4, 333–337. (Departamento de Caldas, Colombia). Geología Colombiana 19, 83–93.
Colony, R.J., Sinclair, J.H., 1932. Metamorphic and igneous rocks of eastern Ecuador. An- González, H., 1980. Geología de las Planchas 167 (Sonson) e 187 (Salamina). Boletín
nals of the New York Academy of Sciences 34, 1–53. Geológico de Ingeominas, Informe 1760.
Cooper, B., et al., 1995. Basin development and tectonic history of the Eastern Cordillera Gradstein, F.M., Ogg, J.G., Smith, A.G., et al., 2004. A geological timescale 2004. Cambridge
and Llanos Basin. Colombia. American Association of Petroleum Geologists 79, University Press, Cambridge.
1421–1443. Grajales-Nishimura, J.M., Centeno-Garcia, E., Keppie, J.D., Dostal, J., 1999. Geochemistry of
Cordani, U.G., Brito-Neves, B.B., D'Agrella, M.S., 2003. From Rodinia to Gondwana: a re- Paleozoic basalts from the Juchatengo complex of southern Mexico: tectonic implica-
view of the available evidence from South America. Gondwana Research 6, 275–283. tions. Journal of South American Earth Sciences 12, 537–544.
Cosma, L., Lapierre, H., Jaillard, E., Laubacher, G., Bosch, D., Desmet, A., Mamberti, M., Hampel, A., 2002. The migration history of the Nazca Ridge along the Peruvian active mar-
Gabriele, P., 1998. Petrographie et geochimie des unites magmatiques de la Cordillère gin: a re-evaluation. Earth and Planetary Science Letters 203, 665–679.
Occidentale d'Equateur (0°30′S): implications tectoniques. Bulletin de la Societe Harris, C., Faure, K., Diamond, R.E., Scheepers, R., 1997. Oxygen and hydrogen isotope 966
Geologique de France 169, 739–751. geochemistry of S- and I-type granitoids: the Cape Granite suite, South Africa. Chem-
Dalmayrac, B., Laubacher, G., Marocco, R., 1980. Géologie des Andes péruviennes. ical Geology 143, 95–114.
Caractères généraux de l'évolution géologique des Andes péruviennes: Travaux et Harrison, T.M., Célérier, J., Aikman, A.B., Hermann, J., Heizler, M.T., 2009. Diffusion of 40Ar
Documents de l'ORSTOM, 122, p. 501. in muscovite. Geochimica et Cosmochimica Acta 73, 1039–1051.
Dasch, L., 1982. U-Pb Geochronology of the Sierra de Perijá, Venezuela(PhD thesis) Case Hartmann, L.A., Santos, J.O.S., 2004. Predominance og high Th/U, magmatic zircon in
Western Reserve University (183 pp.). Brazilian Shield sandstones. Geology 32, 73–76.
De Haller, A., Corfu, F., Fontboté, L., Schaltegger, U., Barra, F., Chiaradia, M., Frank, M., Hastie, A.R., Kerr, A.C., 2010. Mantle plume or slab window? Physical and geochemical
Alvarado, J.Z., 2006. Geology, Geochronology, and Hf and Pb Isotope data of the constraints on the origin of the Caribbean oceanic plateau. Earth-Science Reviews
Raúl-Condestable Iron Oxide-Copper-Gold Deposit, Central Coast of Peru. Economic 98, 283–293.
Geology 101, 281–310. Hastie, A.R., Kerr, A.C., Pearce, J.A., Mitchell, S.F., 2007. Classification of altered volcanic is-
Dehler, S.A., 2012. Initial rifting and breakup between Nova Scotia and Morocco: insight land arc rocks using immobile trace elements: development of the Th-Co discrimina-
from new magnetic models. Canadian Journal of Earth Sciences 49, 1385–1394. tion diagram. Journal of Petrology 48, 2341–2357.
Demouy, S., Paquette, J.-L., Blanquat, M., de, S., Benoit, M., Belousova, E.A., O'Reilly,, S.Y., Helbig, M., Keppie, J.D., Murphy, J.B., Solari, L.A., 2012. U-Pb geochronological constraints
García, F., Tejada, L.C., Gallegos, R., Sempere, T., 2012. Spatial and temporal evolution on the Triassic–Jurassic Ayu Complex, southern Mexico: derivation from the western
of Liassic to Paleocene arc activity in southern Peru unravelled by zircon U-Pb and Hf margin of Pangea-A. Gondwana Research 22, 910–927.
in-situ data on plutonic rocks. Lithos 155, 183–200. Herbert, H.J., Pichler, H., 1983. K-Ar ages of rocks from the Eastern Cordillera of Ecuador.
Dickinson, W.R., Lawton, T.F., 2001. Carboniferous to Cretaceous assembly and fragmenta- Deutschen Geologischen Gesellschaft 134, 483–493.
tion of Mexico. Geological Society of America Bulletin 113, 1142–1160. Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M., Reyes-Harker, A., Stockli, D.F., 2010.
Donnelly, T., Beets, D., Carr, M., Jackson, T., Klaver, G., Lewis, J., Maury, R., Schellekens, H., Linking sedimentation in the northern Andes to basement configuration, Mesozoic
Smith, A., Wadge, G., Westercamp, D., 1990. History and tectonic setting of the Carib- extension, and Cenozoic shortening: evidence from detrital zircon U-Pb ages in the
bean magmatism. In: Dengo, G., Case, J. (Eds.), The Caribbean Region: Boulder, Colo- Eastern Cordillera of Colombia. Geological Society of America Bulletin 122,
rado, Geological Society of America, Geology of North America H, pp. 339–374. 1423–1442.
Ducea, M.N., Gehrels, G.E., Shoemaker, S., Ruiz, J., Valencia, V.A., 2004. Geologic evolution Hughes, R.A., Pilatasig, L.F., 2002. Cretaceous and Tertiary terrane accretion in the Cordil-
of the Xolapa Complex, southern Mexico: evidence from U-Pb zircon geochronology. lera Occidental of the Ecuadorian Andes. Tectonophysics 345, 29–48.
Geological Society of America Bulletin 116, 1016–1025. Hughes, R., Bermúdez, R., Espinel, G., 1998. Mapa Geologico de la Cordillera Occidental del
Eagles, G., 2007. New angles on South Atlantic opening. Geophysical Journal International Ecuador entre 0º–1° S: Quito, Ecuador. Corporación de Desarrollo e Investigación
168, 353–361. Geológica, Minera y Metalúrgica–Ministerio de Energia Ecuador–British Geological
Elías-Herrera, M., Ortega-Gutiérrez, F., 2002. Caltepec fault zone: an Early Permian dextral Survey, scale 1:200,000.
transpressional boundary between the Proterozoic Oaxacan and Paleozoic Acatlán Jaillard, E., Soler, P., 1996. Cretaceous to early Paleogene tectonic evolution of the
complexes, southern Mexico, and regional implications. Tectonics 21. http://dx.doi. northern Central Andes (0–1°S) and its relations to geodynamics. Tectonophysics
org/10.1029/200TC001278. 259, 41–53.
Feininger, T., 1980. Eclogite and related high-pressure regional metamorphic rocks from Jaillard, E., Soler, P., Carlier, G., Mourier, T., 1990. Geodynamic evolution of the northern
the Andes of Ecuador. Journal of Petrology 21, 107–140. and central Andes during early to middle Mesozoic times: a Tethyan model. Journal
Feininger, T., Bristow, C.R., 1980. Cretaceous and Paleogene geologic history of coastal of the Geological Society 147, 1009–1022.
Ecuador. Geologische Rundschau 69, 849–874. Jaillard, E., Laubacher, G., Bengston, P., Dhondt, A.V., Bulot, L.G., 1999. Stratigraphy and
Feininger, T., Silberman, M.L., 1982. K-Ar geochronology of basement rocks on the north- evolution of the Cretaceous forearc Celica-Lancones basin of southwestern Ecuador.
ern flanks of the Huancabamba Deflection, Ecuador. Open File Report. United States Journal of South American Earth Sciences 12, 51–68.
Geological Survey, 82, p. 206. Jaillard, E., Ordoñez, M., Suárez, J., Toro, J., Iza, D., Lugo, W., 2004. Stratigraphy of the late
Feininger, T., Barrero, D., Castro, N., 1972. Geología de Antioquia et Caldas–Sub-zona II-B. Cretaceous–Paleogene deposits of the cordillera occidental of central Ecuador:
Boletin Geología Bogota 20, 173. geodynamic implications. Journal of South American Earth Sciences 17, 49–58.
Fortey, N.J., 1990. Petrographic data and course notes for the Cordillera Real Project, Jaimes, E., de Freitas, M., 2006. An Albian–Cenomanian unconformity in the northern
Ecuador. British Geological Survey Technical Report WG/90/14/R (67 pp.). Andes. Evidence and tectonic significance. Journal of South American Earth Sciences
Frost, C., Snoke, A., 1989. Tobago, West Indies, a fragment of a Mesozoic oceanic island arc: 21, 466–492.
petrochemical evidence. Journal of the Geological Society of London 146, 953–964. John, T., Scherer, E.E., Schenk, V., Herms, P., Halama, R., Garbe-Schönberg, D., 2010.
Frost, B.R., Barnes, C.G., Collins, W.J., Arculus, R.J., Ellis, D.J., Frost, C.D., 2001. A geochemical Subducted seamounts in an eclogite-facies ophiolite sequence: the Andean
classification of granitic rocks. Journal of Petrology 42, 2033–2048. Raspas Complex, SW Ecuador. Contributions to Mineralogy and Petrology 159,
Funck, T., Jackson, H.R., Louden, K.E., Dehler, S.A., Wu, Y., 2004. Crustal structure of the 265–284.
northern Nova Scotia rifted margin (eastern Canada). Journal of Geophysical Re- Jolly, W.T., Lidiak, E.G., Dickin, A.Ap., Wu, T.W., 2001. Secular geochemistry of central
search Solid Earth 109. http://dx.doi.org/10.1029/2004JB003008. Puerto Rican island arc lavas: constraints on Mesozoic tectonism in the eastern Great-
Gabriele, P., 2002. HP terranes exhumation in an active margin setting: geology, petrology er Antilles. Journal of Petrology 42, 2197–2214.
and geochemistry of the Raspas Complex in SW Ecuador(PhD thesis) Université de Jones, P.R., 1981. Crustal structure of the Peru continental margin and adjacent Nazca
Lausanne. plate, 9°S. In: Kulm, L.D., Dymond, E., Dasch, J., Hussong, D.M. (Eds.), Nazca Plate:
Gerbi, C.C., Johnson, S.E., Koons, P.O., 2006. Controls on low-pressure anatexis. Journal of Crustal Formation and Andean Convergence. Memoir Geological Society of America,
Metamorphic Petrology 24, 107–118. 154, pp. 423–444.
Gerdes, A., Zeh, A., 2009. Zircon formation versus zircon alteration — new insights from Kennan, L., Pindell, J., 2009. Dextral shear, terrane accretion and basin formation in the
combined U-Pb and Lu-Hf in-situ LA-ICP-MS analyses, and consequences for the in- Northern Andes: best explained by interaction with a Pacific-derived Caribbean
terpretation of Archean zircon from the Central Zone of the Limpopo Belt. Chemical Plate? Geological Society of London Special Publications 328, 487–531.
Geology 261, 230–243. Keppie, J.D., Gutierrez, F.O., 1999. Middle American Precambrian basement: a missing
Gerdes, A.G., Montero, P.M., Bea, F.B., Fershater, G.F., Borodina, N.B., Osipova, T.O., piece of the reconstructed 1-Ga orogen. In: Ramos, V.A., Keppie, J.D. (Eds.),
Shardakova, G.S., 2002. Peraluminous granites frequently with mantle-like isotope Laurentia-Gondwana connections before Pangaea, Boulder, Colorado. Geological So-
compositions: the continental-type Murzinka and Dzhabyk batholiths of the eastern ciety of America Special Paper, 336.
Urals. International Journal of Earth Sciences 91, 3–19. Keppie, J.D., Dostal, J., Miller, B.V., Ortega-Rivera, A., Roldan-Quintana, J., Lee, J.W.K., 2006.
Gerya, T.V., Stöckhert, B., Perchuk, A.L., 2002. Exhumation of high-pressure meta- Geochronology and geochemistry of the Francisco Gneiss: Triassic continental rift
morphic rocks in a subduction channel: a numerical simulation. Tectonics 21, tholeiites on the Mexican Margin of Pangea metamorphosed and exhumed in a ter-
1056. tiary core complex. International Geology Reviews 48, 1–16.

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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 43

Kerr, A.C., Tarney, J., 2005. Tectonic evolution of the Caribbean and northwestern South Nivia, A., Marriner, G.F., Kerr, A.C., Tarney, J., 2006. The Quebradagrande Complex: a Lower
America: the case for accretion of two Late Cretaceous oceanic plateaus. Geology Cretaceous ensialic marginal basin in the Central Cordillera of the Colombian Andes.
33, 269–272. Journal of South American Earth Sciences 21, 423–436.
Kerr, A.C., Marriner, G.F., Tarney, J., Nivia, A., Saunders, A.D., Thirlwall, M.F., Sinton, C. Noble, S.R., Aspden, J.A., Jemielita, R., 1997. Northern Andean crustal evolution: new U-Pb
W., 1997. Cretaceous basaltic terranes in Western Colombia: elemental, chrono- geochronological constraints from Ecuador. Geological Society of America Bulletin
logical and Sr-Nd isotopic constraints on petrogenesis. Journal of Petrology 38, 109, 789–798.
677–702. Oliver, N.H.S., Zakowski, S., 1995. Timing and geometry of deformation, low-pressure
Kerr, A.C., Aspden, J.A., Tarney, J., Pilatasig, L.F., 2002. The nature and provenance of accret- metamorphism and anatexis in the eastern Mt. Lofty Ranges: the possible role of ex-
ed oceanic terranes in western Ecuador: geochemical and tectonic constraints. Jour- tension. Australian Journal of Earth Sciences 42, 501–507.
nal of the Geological Society 159, 577–594. Oliveros, V., Feraud, G., Aguirre, L., Fornari, M., Morata, D., 2006. The early Andean mag-
Kirsch, M., Keppie, J.D., Murphy, J.B., Solari, L.A., 2012. Permian–Carboniferous arc matic province (EAMP): Ar-40/Ar-39 dating on mesozoic volcanic and plutonic
magmatism and basin evolution along the western margin of Pangaea: geochemical rocks from the Coastal Cordillera, Northern Chile. Journal of Volcanology and Geo-
and geochronological evidence from the eastern Acatlán Complex, southern Mexico. thermal Research 157, 311–330.
Geological Society of America Bulletin 124, 1607–1628. Omarini, R.H., Sureda, R.J., Gotze, H.J., Seilacher, A., Pfluger, F., 1999. Puncoviscana folded
Kirsch, M., Helbig, M., Keppie, J.D., Murphy, J.B., Lee, J.K.W., Solari, L.A., 2014. A Late Trias- belt in northwestern Argentina: testimony of Late Proterozoic Rodinia fragmentation
sic tectonothermal event in the eastern Acatlán Complex, southern Mexico, synchro- and pre-Gondwana collisional episodes. International Journal of Earth Sciences 88,
nous with a magmatic arc hiatus: the result of flat-slab subduction? Lithosphere 6, 76–97.
63–79. Orrego, A., Restrepo, J.J., Toussaint, J.F., Linares, E., 1980. Datación de un esquisto sericítico
Lapierre, H., Bosch, D., Dupuis, V., Polvé, M., Maury, R.C., Hernandez, J., Monié, P., de Jambaló, Cauca. , 25. Geología Universidad Nacional, pp. 133–134 (Special
Yeghicheyan, D., Jaillard, E., Tardy, M., de Lépinay, B.M., Mamberti, M., Desmet, A., Publications).
Keller, F., Sénebier, F., 2000. Mantle plume events in the genesis of the peri- Ortega-Obregon, C., Solari, L., Gomez-Tuena, A., Elias-Herrera, M., Ortega-Gutierrez, F.,
Caribbean Cretaceous oceanic plateau province. Journal of Geophysical Research Macias-Romo, C., 2013. Permian–Carboniferous arc magmatism in southern
105, 8403–8421. Mexico: U-Pb dating, trace element and Hf isotopic evidence on zircons of earliest
Leal-Mejia, H., Draper, J.C.M.I., Shaw, R.P., 2011. Phanerozoic gold metallogeny in the subduction beneath the western margin of Gondwana. International Journal of
Colombian Andes. Proceedings let's talk ore deposits, SGA biennial meeting, Antofa- Earth Sciences. http://dx.doi.org/10.1007/s00531-013-0933-1.
gasta, Chile. Pankhurst, R.J., Rapela, C.W., Fanning, C.M., 2000. Age and origin of coeval TTG, I- and S-
Lebrat, M., Mégard, F., Dupuy, C., Dostal, J., 1987. Geochemistry and tectonic setting of pre- type granites in the Famatinian belt of NW Argentina. Transactions of the Royal Soci-
collision Cretaceous and Paleogene volcanic rocks of Ecuador. Geological Society of ety of Edinburgh: Earth Sciences 91, 151–168.
America Bulletin 99, 569–578. Parson, L.M., Wright, I.C., 1996. The Lau-Havre-Taupo back-arc basin: a southward-
Litherland, M., Aspden, J.A., Jemielita, R.A., 1994. The metamorphic belts of Ecuador. Over- propagating, multi-stage evolution from rifting to spreading. Tectonophysics 263,
seas Memoir of the British Geological Survey, 11, p. 147 (Nottingham, England). 1–22.
Luzieux, L.D.A., Heller, F., Spikings, R., Vallejo, C.F., Winkler, W., 2006. Origin and Creta- Peccerillo, A., Taylor, S.R., 1976. Geochemistry of Eocene calc-alkaline volcanic rocks from
ceous tectonic history of the coastal Ecuadorian forearc between 1°N and 3°S: paleo- the Kastamonu area, Northern Turkey. Contributions to Mineralogy and Petrology 58,
magnetic, radiometric and fossil evidence. Earth and Planetary Science Letters 249, 63–81.
400–414. Pindell, J.L., Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and
Mamberti, M., 2001. Origin and Evolution of Two Distinct Cretaceous Oceanic Plateaus Ac- northern South America in the mantle reference frame: an update. Geological Society
creted in Western Ecuador (South America): Petrological, Geochemical and Isotopic of London Special Publications 328, 1–55.
Evidence(Ph.D. thesis) Université de Lausanne, p. 241. Pindell, J.L., Kennan, L., Maresch, W.V., Stanek, K.P., Draper, G., Higgs, R., 2005. Plate kine-
Mamberti, M., Lapierre, H., Bosch, D., Jaillard, E., Ethien, R., Hernandez, J., Polvé, M., 2003. matics and crustal dynamics of circum-Caribbean arc–continent interactions: tecton-
Accreted fragments of the Late Cretaceous Caribbean-Colombian Plateau in Ecuador. ic controls on basin development in Proto-Caribbean margins. In: Lallemant, A.,
Lithos 66, 173–199. Sisson, V.B. (Eds.), Caribbean–South American Plate Interactions. Geological Society
Mamberti, M., Lapierre, H., Bosch, D., Jaillard, E., Hernandez, J., Polvé, M., 2004. The Early of America Special Paper, 394, pp. 7–52.
Cretaceous San Juan Plutonic Suite, Ecuador: a magma chamber in an oceanic pla- Polliand, M., Schaltegger, U., Frank, M., Fontboté, L., 2005. Formation of intra-arc
teau? Canadian Journal of Earth Sciences 41, 1237–1258. volcanosedimentary basins in the western flank of the central Peruvian Andes during
Maniar, P.D., Piccoli, P.M., 1989. Tectonic discrimination of granitoids. Geological Society Late Cretaceous oblique subduction: field evidence and constraints from U-Pb ages
of America Bulletin 101, 635–643. and Hf isotopes. International Journal of Earth Sciences 94, 231–232.
Maresch, W.V., Kluge, R., Baumann, A., Pindell, J.L., Krückhans-Lueder, G., Stanek, K., 2009. Pratt, W.T., Duque, P., Ponce, M., 2005. An autochthonous geological model for the eastern
The occurrence of high-pressure metamorphism on Margarita Island, Venezuela: a Andes of Ecuador. Tectonophysics 399, 251–278.
constraint on Caribbean-South America interaction. Geological Society of London Ramos, V.A., 2008. The basement of the Central Andes: the Arequipa and related terranes.
Special Publications 328, 705–741. Annual Reviews of Earth and Planetary Sciences 36, 289–324.
Martínez, A.M.C., 2007. Petrogenesis and evolution of Aburra Ophiolite, Colombian Andes, Ramos, V.A., 2009. Anatomy and global context of the Andes: main geologic features and
Central Range(Ph.D. thesis) University of Brasilia, p. 178. the Andean orogenic cycle. The Geological Society of America Memoir 204, 31–65.
Martin-Gombojav, N., Winkler, W., 2008. Recycling of Proterozoic crust in the Andean Ramos, V.A., Aleman, A., 2000. Tectonic evolution of the Andes. In: Cordani, U.G., Milani, E.
Amazon foreland of Ecuador: implications for orogenic development of the Northern J., Thomaz Filha, A., Campos, D.A. (Eds.), Tectonic Evolution of South America. 31st In-
Andes. Terra Nova 20, 22–31. ternational Geological Congress, Rio de Janerio, pp. 635–685.
Maya, M., González, H., 1995. Unidades Litodémicas en la Cordillera Central de Colombia. Rapela, C.W., Pankhurst, R.J., Fanning, C.M., Herve, F., 2005. Pacific subduction coeval with
Informe Unidad Operativa Medellín, Ingeominas, Colombia, pp. 44–57. the Karoo mantle plume: the early Jurassic subcordilleran belt of northwestern Pata-
McCourt, W.J., 1981. The geochemistry and petrography of the Coastal Batholith of Peru, gonia. In: Vaughan, A.P.M., Leat, P.T., Pankhurst, R.J. (Eds.), Terrane processes at the
Lima Segment. Journal of the Geological Society 138, 407–420. margins of Gondwana. Geological Society of London, 246, pp. 217–239 (Geological
McCourt, W.J., Aspden, J.A., Brook, M., 1984. New geological and geochemical data from Society Special Publications).
the Colombian Andes: continental growth by multiple accretion. Journal of the Geo- Reitsma, M.J., 2012. Reconstructing the Late Paleozoic–Early Mesozoic Plutonic and Sedi-
logical Society 831–845. mentary Record of South-East Peru: Orphaned Back-Arcs Along the Western Margin
Mégard, F., 1984. The Andean orogenic period and its major structures in central and of Gondwana(PhD thesis) Terre & Environment, 111. University of Geneva,
northern Peru. Journal of the Geological Society of London 141, 893–900. Switzerland, p. 226.
Mišković, A., Schaltegger, U., 2009. Crustal growth along a non-collisional cratonic mar- Restrepo, J.J., Toussaint, J.F., 1976. Edades radiométricas de algunas rocas de Antioquia,
gin: a Lu-Hf isotopic survey of the Eastern Cordilleran granitoids of Peru. Earth and Colombia. Publicacion Especial Geología. , 12. Universidad Nacional de Medellin,
Planetary Science Letters 279, 303–315. pp. 1–11.
Mišković, A., Spikings, R.A., Chew, D.M., Košler, J., Ulianov, A., Schaltegger, U., 2009. Restrepo, J.J., Toussaint, J.F., 1982. Metamorfismos superpuestos em la Cordillera Central
Tectonomagmatic evolution of Western Amazonia: geochemical characterization de Colombia. V Congreso Latino-Americano de Geología, Buenos Aires, Argentina,
and zircon U-Pb geochronologic constraints from the Peruvian Eastern Cordilleran pp. 505–512.
granitoids. Geological Society of America Bulletin 121, 1298–1324. Restrepo, J.J., Toussaint, J.F., 1988. Terranes and continental accretion in the Colombian
Montes, C., Guzman, G., Bayona, G., Cardona, A., Valencia, V., Jaramillo, C., 2010. Clockwise Andes. Episodes 7, 189–193.
rotation of the Santa Marta massif and simultaneous Paleogene to Neogene deforma- Restrepo, J.J., Toussaint, J.F., Gonzáles, H., Cordani, U., Kawashita, K., Linares, E., Parica, C.,
tion of the Plato-San Jorge and Cesar-Ranchería basins. Journal of South American 1991. Precisiones geocronológicas sobre el occidente Colombiano. Simposio sobre
Earth Sciences 29, 832–848. magmatismo andino y su marco tectónico. Memorias, Tomo 1, Manizales,
Moreno, M., Pardo, A., 2003. Stratigraphical and sedimentological constraints on Western Colombia, pp. 1–22.
Colombia: implications on the evolution of the Caribbean plate. In: Bartolini, C., Restrepo, J.J., Ordóñez-Carmona, O., Moreno-Sánchez, M., 2009. A comment on “The
Buffler, R.T., Blickwede, J. (Eds.), The Circum-Gulf of Mexico and the Caribbean: Hy- Quebradagrande Complex: a Lower Cretaceous ensialic marginal basin in the Central
drocarbon Habitats, Basin Formation, and Plate Tectonics. American Association of Cordillera of the Colombian Andes” by Nivia et al. Journal of South American Earth
Petroleum Geologists Memoir, 79, pp. 891–924. Sciences 28, 204–205.
Mourier, T., Laj, C., Mégard, F., Roperch, P., Mitouard, P., Farfan Medrano, A., 1988. An ac- Restrepo, J.J., Ordoñez-Carmona, O., Armstrong, R., Pimentel, M.M., 2011. Triassic meta-
creted continental terrane in northwestern Peru. Earth and Planetary Science Letters morphism in the northern part of the Tahamí Terrane of the central cordillera of
88, 182–192. Colombia. Journal of South American Earth Sciences 32, 497–507.
Mukasa, S.B., 1986. Zircon U-Pb ages of super-units in the Coastal batholith, Peru: impli- Restrepo-Pace, P.A., Cediel, F., 2010. Northern South America basement tectonics and im-
cations for magmatic and tectonic processes. Geological Society of America Bulletin plications for paleocontinental reconstructions of the Americas. Journal of South
97, 241–254. American Earth Sciences 29, 764–771.

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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
44 R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx

Restrepo-Pace, P.A., Ruiz, J., Gehrels, G., Cosca, M., 1997. Geochronology and Nd isotopic Spadea, P., Espinosa, E., Orrego, A., 1989. High-Mg extrusive rocks from the Romeral
data for Grenville-age rocks in the Colombian Andes: new constraints for Late zone ophiolites in the south western Colombian Andes. Chemical Geology 77,
Proterozoic–Early Paleozoic paleocontinental reconstructions of the Americas. Earth 303–321.
and Planetary Science Letters 150, 427–441. Spikings, R.A., Crowhurst, P.V., 2004. (U-Th)/He thermochronometric constraints on the
Reynaud, C., Jaillard, E., Lapierre, H., Mamberti, M., Mascle, G.H., 1999. Oceanic plateau and late Miocene-Pliocene tectonic development of the northern Cordillera Real and the
island arcs of southwestern Ecuador: their place in the geodynamic evolution of Interandean Depression, Ecuador. Journal of South American Earth Sciences 17,
northwestern South America. Tectonophysics 307, 235–254. 239–251.
Riding, J.B., 1989. A palynological investigation of nine rock samples from Ecuador Spikings, R.A., Seward, D., Winkler, W., Ruiz, G.M., 2000. Low-temperature
(Maguazo Unit). British Geological Survey Technical Report WH/89/361/R (4 pp.). thermochronology of the northern Cordillera Real, Ecuador: tectonic insights
Riel, N., Guillot, S., Jaillard, E., Martelat, J.-E., Paquette, J.-L., Schwartz, S., Gonclaves, P., from zircon and apatite fission track analysis. Tectonics 19, 649–668.
Duclaux, G., Thebaud, N., Lanari, P., Janots, E., Yuquilema, J., 2013. Metamorphic and Spikings, R.A., Winkler, W., Seward, D., Handler, R., 2001. Along-strike variations in
geochronological study of the Triassic El Oro metamorphic complex, Ecuador: impli- the thermal and tectonic response of the continental Ecuadorian Andes to the
cations for high-temperature metamorphism in a forearc zone. Lithos 156–159, collision with heterogeneous oceanic crust. Earth and Planetary Science Letters
41–68. 186, 57–73.
Rodriguez, G., Zapata, G., 2013. Comparative analysis of the Barroso Formation and Spikings, R.A., Winkler, W., Hughes, R.A., Handler, R., 2005. Thermochronology of alloch-
Quebradagrande Complex: a volcanic arc tholeiitic–calcoalcaline, segmented by thonous terranes in Ecuador: unravelling the accretionary and post-accretionary his-
the fault system Romeral in Northern Andes. Boletin Ciencias de la Tierra 33, tory of the Northern Andes. Tectonophysics 399, 195–220.
39–58. Spikings, R.A., Crowhurst, P.V., Winkler, W., Villagomez, D., 2010. Syn- and post accretion-
Romero, D., Valencia, K., Alarcón, P., Peña, D., Ramos, V.A., 2013. The offshore basement of ary cooling history of the Ecuadorian Andes constrained by their in-situ and detrital
Peru: evidence for different igneous and metamorphic domains in the forearc. Journal thermochronometric record. Journal of South American Earth Sciences 30, 121–133.
of South American Earth Sciences 42, 47–60. Sun, S.S., McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts:
Romeuf, N., Aguirre, L., Soler, P., Féraud, G., Jaillard, E., Ruffet, G., 1995. Middle Jurassic vol- implications for mantle composition and processes. Geological Society, London, Spe-
canism in the Northern and Central Andes. Revisita Geologica de Chile 22, 245–259. cial Publications 42, 313–345.
Rosas, S., Fontboté, L., Tankard, A., 2007. Tectonic evolution and paleogeography of the Taylor, S.R., McLennan, S.M., 1995. The geochemical evolution of the continental crust. Re-
Mesozoic Pucará Basin, central Peru. Journal of South American Earth Sciences 24, views of Geophysics 33, 241–265.
1–24. Torres, R., Ruíz, J., Patchett, P.J., Grajales-Nishimura, J.M., 1999. Permo-Triassic continental
Rosenbaum, G., Giles, D., Saxon, M., Betts, P.G., Weinberg, R.F., Dubox, C., 2005. Subduction arc in eastern Mexico; tectonic implications for reconstructions of southern North
of the Nazca Ridge and the Inca Plateau: insights into the formation of ore deposits in America. In: Bartolini, C., Wilson, J.L., Lawton, T.F. (Eds.), Mesozoic Sedimentary and
Peru. Earth and Planetary Science Letters 239, 18–32. Tectonic History of North-Central Mexico. Geological Society of America Special
Rubatto, D., 2002. Zircon trace element geochemistry: partitioning with garnet and the Paper, 340, pp. 191–196.
link between U-Pb ages and metamorphism. Chemical Geology 184, 123–138. Toussaint, J.F., Restrepo, J.J., 1994. The Colombian Andes during Cretaceous times. In:
Ruiz, J., Tosdal, R.M., Restrepo, P.A., Murillo-Muñetón, G., 1999. Pb isotope evidence for Salfity, J.A. (Ed.), Cretaceous Tectonics of the Andes. Earth Evolution Series. Vieweg
Columbia–southern Mexico connections in the Proterozoic. In: Ramos, V.A., Keppie, and Teubner Verlag, pp. 61–100.
J.D. (Eds.), Laurentia–Gondwana Connections Before Pangea. Geological Society of Vallejo, C., Spikings, R.A., Luzieux, L., Winkler, W., Chew, D., Page, L., 2006. The early inter-
America Special Paper, 336, pp. 183–197. action between the Caribbean Plateau and the NW South American Plate. Terra Nova
Ruiz, G.M.H., Seward, D., Winkler, W., 2004. Detrital thermochronology; a new perspec- 18, 264–269.
tive on hinterland tectonics, an example from the Andean Amazon Basin, Ecuador. Vallejo, C., Winkler, W., Spikings, R.A., Luzieux, L., Heller, F., Bussy, F., 2009. Mode and
Basin Research 16, 413–430. timing of terrane accretion in the forearc of the Andes of Ecuador. The Geological So-
Ruiz, G., Seward, D., Winkler, W., Maria, A.M., David, T.W., 2007. Evolution of the Amazon ciety of America Memoir 204, 197–216.
Basin in Ecuador with special reference to hinterland tectonics: data from zircon Van der Lelij, R., 2013. Reconstructing North-Western Gondwana With Implications
fission-track and heavy mineral analysis. Developments in Sedimentology 58, for the Evolution of the Iapetus and Rheic Oceans: A Geochronological,
907–934. Thermochronological and Geochemical study(PhD thesis) Terre & Environment,
Sarmiento, L.F., Rangel, A., 2004. Petroleum systems of the Upper Magdalena Valley, 121. University of Geneva, Switzerland, p. 221.
Colombia. Marine and Petroleum Geology 21, 373–391. Van der Lelij, R., Spikings, R.A., Kerr, A.C., Kounov, A., Cosca, M., Chew, D., Villagomez, D.,
Sarmiento-Rojas, L.F., Van Wess, J.D., Cloetingh, S., 2006. Mesozoic transtensional basin 2010. Thermochronology and tectonics of the Leeward Antilles: evolution of the
history of the Eastern Cordillera, Colombian Andes: inferences from tectonic models. southern Caribbean Plate boundary zone. Tectonics 29. http://dx.doi.org/10.1029/
Journal of South American Earth Sciences 21, 383–411. 2009tc002654.
Saur, W., 1950. Mapa geológica del Ecuador 1:500 000. Quito, Universidad Central and Vásquez, M., Altenberger, U., Romer, R.L., Sudo, M., Moreno-Murillo, J.M., 2010. Magmatic
Dirección de Minería. evolution of the Andean Eastern Cordillera of Colombia during the Cretaceous: Influ-
Schaaf, P., Weber, B., Weis, P., Groß, A., Ortega-Gutiérrez, F., Köhler, H., 2002. The Chiapas ence of previous tectonic processes. Journal of South American Earth Sciences 29,
Massif (México) revised; new geologic and isotopic data for basement characteristics. 171–186.
In: Miller, H. (Ed.), Contributions to Latin-American Geology: Neues Jahrbuch für Vásquez, M., Altenberger, U., 2005. Mid-Cretaceous extension-related magmatism in
Geologie und Paläontologie. Abhandlungen, 225, pp. 1–23. the eastern Colombian Andes. Journal of South American Earth Sciences 20,
Scheuber, E., Gonzalez, G., 1999. Tectonics of the Jurassic–Early Cretaceous magmatic arc 193–210.
of the north Chilean Coastal Cordillera (22°–26°S): a story of crustal deformation Vergara, L., Prössl, K.F., 1994. Dating the Yaví Formation (Aptian, Upper Magdalena Valley,
along a convergent plate boundary. Tectonics 18, 895–910. Colombia), palynological results. In: Etayo, F. (Ed.), Estudios Geológicos del Valle Su-
Schlische, R.W., 2002. Progress in understanding the structural geology, basin evolution, perior del Magdalena. Departamento de Geociencias, Universidad Nacional de
and tectonic history of the eastern North American rift system. Intorsvik In: Colombia, Bogotá, p. 14.
LeTourneau, P.M., Olsen, P.E. (Eds.), The Great Rift Valleys of Pangaea in Eastern Villagómez, D., Spikings, R., 2013. Thermochronology and tectonics of the Central and
North America, 1. Columbia University Press, New York. Western Cordilleras of Colombia: Early Cretaceous–Tertiary evolution of the North-
Schütte, P., Chiaradia, M., Beate, B., 2010. Geodynamic controls on Tertiary arc magmatism ern Andes. Lithos 168, 228–249.
in Ecuador: constraints from U-Pb zircon geochronology of Oligocene–Miocene intru- Villagómez, D., Spikings, R., Magna, T., Kammer, A., Winkler, W., Beltrán, A., 2011. Geo-
sions and regional age distribution trends. Tectonophysics 489, 159–176. chronology, geochemistry and tectonic evolution of the Western and Central cordil-
Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto, V., Jacay, J., Arispe, O., Néraudeau, D., leras of Colombia. Lithos 125, 875–896.
Cárdenas, J., Rosas, S., Jiménez, N., 2002. Late Permian–Middle Jurassic lithospheric Villamil, T., 1999. Campanian–Miocene tectonostratigraphy, depocenter evolution and
thinning in Peru and Bolivia, and its bearing on Andean-age tectonics. Tectonophysics basin development of Colombia and western Venezuela. Palaeogeography,
345, 153–181. Palaeoclimatology, Palaeoecology 153, 239–275.
Sheppard, G., Bushnell, G.H.S., 1933. Metamorphic rocks of the eastern Andes near Vinasco, C.J., Cordani, U.G., Gonzales, H., Weber, M., Pelaez, C., 2006. Geochronological,
Cuenca, Ecuador. Geological Magazine 70, 321–330. isotopic, and geochemical data from Permo-Triassic granitic gneisses and granitoids
Sherlock, S.C., Kelley, S.P., 2002. Excess argon evolution in HP-LT rocks: a UVLAMP study of the Colombian Central Andes. Journal of South American Earth Sciences 21,
of phengite and K-free minerals, NW Turkey. Chemical Geology 182, 619–636. 355–371.
Shervais, J.W., 1982. Ti-V plots and the petrogenesis of modern and ophiolitic lavas. Earth Viscarret, P., Wright, J., Urbani, F., 2009. New U-Pb zircon ages of El Baúl Massif, Cojedes
and Planetary Science Letters 59, 101–118. State, Venezuela. Revisita Técnica de la Facultad de Ingenieria Universidad del Zulia
Sinton, C.W., Duncan, R.A., Storey, M., Lewis, J., Estrada, J.J., 1998. An oceanic flood basalt 32, 210–221.
province within the Caribbean plate. Earth and Planetary Science Letters 155, Weber, B., Iriondo, A., Premo, W.R., Hecht, L., Schaaf, P., 2007. New insights into the histo-
221–235. ry and origin of the southern Maya block, SE Mexico: U-Pb-SHRIMP zircon geochro-
Sláma, J., Košler, J., Condon, D., Crowley, J.L., Gerdes, A., Hanchar, J.M., Horstwood, M.S.A., nology from metamorphic rocks of the Chiapas massif. International Journal of Earth
Morris, G.A., Nasdala, G.A., Norberg, L., Schaltegger, U., Schoene, B., Turbrett, M.N., Sciences 96, 253–269.
Whitehouse, M.J., 2008. Plešovice zircon–A new natural reference material for U-Pb Weber, M., Cardona, A., Valencia, V., García-Casco, A., Tobón, M., Zapata, S., 2010. U/Pb de-
and Hf isotopic microanalysis. Chemical Geology 249, 1–35. trital zircon provenance from late Cretaceous metamorphic units of the Guajira Pen-
Solari, L.A., Dostal, J., Ortega-Gutierrez, F., Keppie, J.D., 2001. The 275 Ma arc-related La insula, Colombia: tectonic implications on the collision between the Caribbean arc
Carbonera stock in the northern Oaxacan Complex of Southern Mexico: U-Pb geo- and the South American margin. Journal of South American Earth Sciences 29,
chronology and geochemistry. Revista Mexicana de Ciencias Geológicas 18, 149–161. 805–816.
Solari, L.A., Gómez-Tuena, A., Ortega-Gutiérrez, F., Ortega-Obregón, C., 2011. The Chaucús White, R.V., Tarney, J., Kerr, A.C., Saunders, A.D., Kempton, P.D., Pringle, M.S., Klaver, G.T.,
Metamorphic Complex, central Guatemala: geochronological and geochemical con- 1999. Modification of an oceanic plateau, Aruba, Dutch Caribbean: implications for
straints on its Paleozoic–Mesozoic evolution. Geological Acta 9, 329–350. the generation of continental crust. Lithos 46, 43–68.

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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004
R. Spikings et al. / Gondwana Research xxx (2014) xxx–xxx 45

Wiedenbeck, M., Allé, P., Corfu, F., Griffin, 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, Dr. Roelant van der Lelij completed his MSc in the Universi-
trace element and REE analyses. Geostandards and Geoanalytical Research 19, 1–23. ty of Geneva in 2008 where studied the tectonic evolution of
Winkler, W., Villagomez, D., Spikings, R., Abegglen, P., Tobler, S., Egüez, A., 2005. The the South Caribbean Plate Boundary. As part of his PhD pro-
Chota basin and its significance for the inception and tectonic setting of the ject between 2008 and 2013 in the University of Geneva, he
Interandean depression in Ecuador. Journal of South American Earth Sciences 19, worked on the Phanerozoic tectonic history of igneous and
5–19. metamorphic basement rocks exposed in the Santander
Winter, Th., Lavenú, A., 1989. Morphological and microtectonic evidence for a major ac- Massif of Colombia and the Merida Andes of Venezuela. Since
tive right-lateral strike–slip fault across central Ecuador (South America). Annales March 2014 he manages the new K/Ar geochronology labo-
Tectonicae 3, 123–139. ratory in the Geological Survey of Norway in Trondheim,
Wright, J.E., Wyld, S.J., 2004. Aruba and Curaçao: remnants of a collided Pacific oceanic which aims to constrain the history of brittle tectonics and
plateau? Initial geologic results from the BOLIVAR project. Eos Transactions of the landscape evolution of onshore and offshore Norway.
American Geophysical Union 85, 47.
Xu, H., Ma, C., Ye, K., 2007. Early cretaceous granitoids and their implications for the col-
lapse of the Dabie orogen, eastern China: SHRIMP zircon U-Pb dating and geochem-
istry. Chemical Geology 240, 238–259.
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 Dr. Cristian Vallejo graduated at the Escuela Politécnica
tectonics. Geological Society of America Bulletin 103, 817–828. Nacional-Quito, and obtained a PhD from ETH-Zürich, study-
Zapata, J.P., Restrepo, J.J., Martens, U., Cardona, A., Brito, R., 2011. Geochronology and geo- ing the Geodynamics of the Western Andes of Ecuador and
chemistry of the basic sequence of Altamira, Antioquia, Western Cordillera of its relationship with the collision of the Caribbean Plateau.
Colombia. 14th Congresso Colombiano de Geología, Medellin, Colombia. After his PhD he worked as a Research Fellow on the Strati-
Zerfass, H., Chemale, F., Schultz, C.L., Lavina, E., 2004. Tectonics and sedimentation in graphic Development of Slope Systems Consortium Research
Southern South America during Triassic. Sedimentary Geology 166, 265–292. 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 Politécnica Nacional,
Quito.
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-Zürich, and as tenured re-
search staff at the University of Geneva where he manages Prof. Wilfried Winkler has held a position of senior lecturer
the 40Ar/39Ar laboratory. His research has focussed on and professor for sedimentary petrology, basin analysis and
thermochronology and geochronology of the Andean cordil- sedimentology at the ETH Zurich since 1988. He graduated
leras in Ecuador, Colombia, Venezuela, Peru and Chile. More in 1977 at the University of Fribourg, Switzerland, where he
recently, Richard has focussed his research efforts on bulk subsequently obtained a PhD in geology in 1981. From
and in-situ U-Pb thermochronology of accessory phases. 1981 to 1988, he was a postdoctoral fellow and lecturer at
both Fribourg and Basel Universities. His main research in-
terests are the bearings of tectonics and climate on basin
sedimentation, and provenance studies in different plate tec-
tonics settings. He carried out research in the Alps,
Carpathians, Pyrenees, Northern Andes (Ecuador, Peru),
Dr. Ryan Cochrane graduated in geology at the University of Sinai Peninsula (Egypt), Central Asian Orogenic Belt
Johannesburg in 2005, after which he worked as a gold ex- (Mongolia) and Myanmar.
ploration 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 geo-
chemistry and thermochronology from the University of
Geneva in mid-2013. Ryan immediately started work at Prof. Bernado Beate is a Professor of Geology at the Escuela
Thomson Reuters GFMS in London as a Mine Economics An- Politécnica Nacional, Quito, Ecuador. His main research inter-
alyst with a focus on precious metals and mining research, ests include the petrology of volcanic rocks, geothermal sys-
including the maintenance of mine economics products. Re- tems and ore geology, and he is a specialist in mapping
cently, Ryan was promoted to a Senior Analyst and focusses volcanic systems and mineral alteration assemblages. Prof.
on the economics of mining, and deriving corporate valua- Beate also operates as a consultant for the mineral resource
tions for precious metal mining companies and geothermal industry, and has worked in Ecuador,
Colombia, Peru, Bolivia, Paraguay and New Zealand.

Dr. Diego Villagómez is qualified as an Engineer in Geology


(EPN, Quito, Ecuador) and a PhD in Geology from the Univer-
sity 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 geother-
mal energy projects, defining 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 (290–75 Ma), Gondwana Research (2014), http://dx.doi.org/10.1016/j.gr.2014.06.004

You might also like