You are on page 1of 20

Gondwana Research 66 (2019) 207–226

Contents lists available at ScienceDirect

Gondwana Research

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

Cretaceous extensional and compressional tectonics in the Northwestern


Andes, prior to the collision with the Caribbean oceanic plateau
S. Zapata a,b,c,⁎, A. Cardona d, J.S. Jaramillo a, A. Patiño a, V. Valencia e, S. León a, D. Mejía a,
A. Pardo-Trujillo f, J.P. Castañeda a
a
Facultad de Minas, Universidad Nacional de Colombia, Cr. 80 # 65-223, Medellín, Colombia
b
Corporación Geológica ARES, Calle 44A N. 53-96, Bogotá D.C., Colombia
c
Institute of Earth and Environmental Sciences, University of Potsdam, Karl-Liebknecht-Str. 24-25, 14476 Potsdam, Germany
d
Departamento de Procesos y Energía, Universidad Nacional de Colombia, Cr. 80 # 65-223, Medellín, Colombia
e
School of Earth and Environment, Washington State University, Pullman, United States
f
Instituto de Investigaciones en Estratigrafía-IIES, Universidad de Caldas, Calle 65 N° 26-10, Manizales, Colombia

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

Article history: The Cretaceous units exposed in the northwestern segment of the Colombian Andes preserve the record of exten-
Received 29 December 2017 sional and compressional tectonics prior to the collision with Caribbean oceanic terranes. We integrated field,
Received in revised form 8 October 2018 stratigraphic, sedimentary provenance, whole rock geochemistry, Nd isotopes and U-Pb zircon data to under-
Accepted 10 October 2018
stand the Cretaceous tectonostratigraphic and magmatic record of the Colombian Andes. The results suggest
Available online 22 November 2018
that several sedimentary successions including the Abejorral Fm. were deposited on top of the continental base-
Handling Editor: R.D. Nance ment in an Early Cretaceous backarc basin (150–100 Ma). Between 120 and 100 Ma, the appearance of basaltic
and andesitic magmatism (~115–100 Ma), basin deepening, and seafloor spreading were the result of advanced
Keywords: stages of backarc extension. A change to compressional tectonics took place during the Late Cretaceous
Northern Andes (100–80 Ma). During this compressional phase, the extended blocks were reincorporated into the margin, closing
Paleogeography the former Early Cretaceous backarc basin. Subsequently, a Late Cretaceous volcanic arc was built on the conti-
Cretaceous nental margin; as a result, the volcanic rocks of the Quebradagrande Complex were unconformably deposited
Extension on top of the faulted and folded rocks of the Abejorral Fm. Between the Late Cretaceous and the Paleocene
Convergent margins
(80–60 Ma), an arc-continent collision between the Caribbean oceanic plateau and the South-American conti-
Provenance
nental margin deformed the rocks of the Quebradagrande Complex and shut-down the active volcanic arc. Our
results suggest an Early Cretaceous extensional event followed by compressional tectonics prior to the collision
with the Caribbean oceanic plateau.
© 2019 International Association for Gondwana Research. Published by Elsevier B.V. All rights reserved.

1. Introduction accretion of exotic terranes, changes in the relative plate-convergence


rates, and variations in the slab subduction angle (Cochrane et al.,
Accretionary orogens are characterized by continuous long-lived 2014b; Kerr et al., 1996; Nivia et al., 2006; Restrepo and Toussaint,
subduction systems and the alternation between extensional and com- 1988; Spikings et al., 2015; Villagómez and Spikings, 2013; Zuluaga
pressional tectonics (Cawood et al., 2009; Royden, 1993; Uyeda and et al., 2015). However, the tectonic mechanisms and the precise tempo-
Kanamori, 1979). The accretion of oceanic and continental terranes re- ral constraints of these extensional and compressional phases remain
sults in a complex superposition of multiple sedimentary, magmatic, controversial (e.g. Nivia et al., 2006; Restrepo et al., 2009; Spikings
and deformational events (Cawood et al., 2009). et al., 2015).
The tectonic evolution of the Northern Andes is considered a typical The Cretaceous tectonic reconstructions proposed for the
accretionary orogeny related to different phases of extension and com- Northern Andes can be divided into two opposite models summa-
pression, in a subduction system active since the Jurassic (Ramos, 1999; rized in Fig. 2. The first model proposes continuous compression re-
Restrepo and Toussaint, 1991; Spikings et al., 2015). The change be- lated to the Early Cretaceous accretion of at least one oceanic
tween contrasting geological configurations have been related to the terrane. This accretionary event was followed by the growth of a
Late Cretaceous volcanic arc along the continental edge of the NW
⁎ Corresponding author at: University of Potsdam, Institute of Earth and Environmental
South American margin (Spikings et al., 2015; Toussaint and
Sciences, Karl-Liebknecht-Str. 24-25, 14476 Potsdam, Germany. Restrepo, 1996; Villagómez et al., 2011; Villagómez and Spikings,
E-mail address: szapata@uni-potsdam.de (S. Zapata). 2013). The second model proposes the existence of an Early

https://doi.org/10.1016/j.gr.2018.10.008
1342-937X/© 2019 International Association for Gondwana Research. Published by Elsevier B.V. All rights reserved.
208 S. Zapata et al. / Gondwana Research 66 (2019) 207–226

Cretaceous extensional continental arc coeval with the formation of In this contribution, we present new field observations, sedimentary
a backarc basin (141–115 Ma) (Kennan and Pindell, 2009; Nivia provenance, geochemical, U-Pb data from the Early Cretaceous volcano-
et al., 2006; Spikings et al., 2015; Villagómez et al., 2011). Both sedimentary record and from the Permo-Triassic continental basement
models suggest the collision of one or more oceanic terranes with exposed between 5°53′N and 5°30′N, in the western flank of the Central
the continental margin between the Late Cretaceous and the Paleo- Cordillera, in the Colombian Andes. These results together with pub-
cene (80–60 Ma). lished data are used to reconstruct the Cretaceous tectonic evolution

Atlantic Ocean GP

SNM
SP

SMF
SM

OPF
RFS
GF

CPB
CC
MV
EC

CV Fig. 2

WC

200 km
RFS

Paleogene plutonic rocks


Romeral fault zone
(Quebradagrande C. and Arquía C.)
Late Cretaceous plutonic rocks
Abejorral Fm and other
CR siliciclastic Cretaceous rocks
Eastern Cordillera basin
Allochthonous oceanic basement
Jurassic igneous rocks
Continental basement

Fig. 1. Geological provinces of the Northern Andes. The black square shows the location of the geological map presented in Fig. 2. WC: Western Cordillera; CC: Central Cordillera; EC:
Eastern Cordillera; CR: Cordillera Real; CV: Cauca Valley; MV: Magdalena Valley; GP: Guajira Peninsula; SNM: Sierra Nevada de Santa Marta; SP: Serrania de Perija; SM: Santander
Massif; CPB: Choco-Panama Block; GP: Garrapatas Fault; RFS: Romeral Fault System; OPF: Otu-Pericos Fault; SMF: Santa Marta Fault.
S. Zapata et al. / Gondwana Research 66 (2019) 207–226 209

of the Colombian Andes. Our model suggests an extensional tectonic re- 2012; Maya and Gonzalez, 1995; Vinasco et al., 2006) (Fig. 1). The Cen-
gime during the Early Cretaceous and constraints the time of the transi- tral Cordillera siliciclastic successions are characterized by coarse-
tion into a Late Cretaceous compressional regime coeval with arc grained rocks deposited in fluvial-deltaic environments overlain by
magmatism. This transition to compressional tectonics was responsible finer grain lithologies deposited in marine conditions. These sedimen-
for the onset of the North Andean compressional tectonic settings, prior tary units were deposited between the Berrasian and the Aptian, as
to the collision with the Caribbean plateau. the fossil record suggests (Gomez et al., 1995; Gomez et al., 2002;
González, 2001; Quiroz, 2005) (Fig. 1).
2. Geological framework These sedimentary rocks are grouped in several units including the
Abejorral Fm., Valle Alto Fm., Aranzazu-Manizales Complex; and the
The Cretaceous tectono-stratigraphic record of the Colombian Andes San Luis, Berlín, la Soledad and Segovia sediments (Gomez et al., 1995;
is exposed along the three main Cordilleras and the intervening river González, 2001; Quiroz, 2005). These units have been related to differ-
valleys (Cauca and Magdalena Valleys) (Fig. 1). To the east, the Magda- ent tectonic settings including a passive continental margin (Gomez
lena Valley and the Eastern Cordillera include up to 2 km of Berrasian to et al., 2002; Pardo-Trujillo et al., 2002; Toussaint and Restrepo, 1996),
Aptian marine siliciclastic rocks deposited in extensional basins a backarc basin (Nivia et al., 2006) (Fig. 2B), and a foreland basin
(Sarmiento-Rojas et al., 2006) with limited magmatic activity (Spikings et al., 2015; Villagómez et al., 2011) (Fig. 2A).
(Vasquez and Altenberger, 2005). These marine siliciclastic sequences The Central Cordillera metamorphic basement is composed of low to
are overlain by fine-grained siliciclastic rocks and carbonates deposited medium grade meta-sedimentary rocks intruded by Permo-Triassic
in a shallow platform environment. These fine-grained rocks were de- granitoids and amphibolites; these metamorphic rocks are grouped in
posited during a tectonically-quiescent stage after the Albian (Villamil the Cajamarca Complex (Gómez et al., 2015; Maya and Gonzalez,
and Pindell, 1999). Subsequently, the stratigraphic record suggests a 1995). The Cretaceous sedimentary rocks and the metamorphic base-
change to transitional-deltaic and fluvial depositional environments. ment of the Central Cordillera are intruded by several intermediate-
This change from marine to deltaic and fluvial deposition has been felsic plutonic bodies (Fig. 1), including the Antioquia Batholith and
linked to a Late Cretaceous collisional event, between the Great Arc of the Altavista Stock. These intrusives have yield U-Pb crystallization
the Caribbean and the NW of the South American margin (Erlich et al., ages between 90 and 63 Ma (Correa et al., 2006; Ibañez-Mejia et al.,
2003; Escalona and Mann, 2011; Gomez et al., 2003). 2007; Leal-Mejia, 2011; Villagómez et al., 2011).
The sedimentary record in the Central Cordillera includes discontin- Different deformed Permo-Triassic to Cretaceous metamorphic and
uous exposures of Lower Cretaceous siliciclastic rocks deposited on top volcano-sedimentary rocks are exposed at the western segment of the
of a pre-Cretaceous metamorphic and igneous basement (Martens et al., Central Cordillera and along the Cauca Valley. These units are limited

A) Spikings et al. 2015 and Villagómez et al. 2013 B) Spikings et al. 2015, Villagómez et al. 2013 and Nivia et al. 2006

190 - 145 Ma Jurassic arc 160- 145 Ma

145 - 130 Ma
145- 130 Ma
Abejorral rift

130 - 115 Ma Quebradagrande arc


125- 100 Ma
Quebradagrande rift
Silvia Pijao San Jerónimo
115 - 100 Ma Fault Fault
Quebradagrande
collision
100 - 90 Ma 100 - 80 ma
Basin closure

Spikings et al. 2015 and Villagómez et al. 2013


90 - 80 Ma Late Cretaceous plutonic rocks
Quebradagrande Complex
Abejorral Fm.
Western Cordillera
Allochthonous oceanic basement
75 - 70 Ma, Western Cordillera collision
Cauca Almaguer Arquía Complex
Fault
Jurassic igneous rocks
Continental basementt

Fig. 2. Tectonic models proposed for the Cretaceous evolution of the Northern Andes. (A), Compressional tectonic scenario, modified from Spikings et al. (2015), Restrepo and Toussaint
(1988), and Villagómez et al. (2011). (B), Extensional tectonic scenario, modified from Villagómez et al. (2011), Spikings et al. (2015) and Nivia et al. (2006).
210 S. Zapata et al. / Gondwana Research 66 (2019) 207–226

by the faults of the Cauca-Romeral Fault System (RFS) (Figs. 1 and 3) The Quebradagrande Complex includes a series of discontinuous N-S
(González, 2001; Maya and Gonzalez, 1995; Restrepo and Toussaint, trending deformed basaltic to dacitic lavas and pyroclastic rocks inter-
1988; Vinasco Vallejo and Cordani, 2012). bedded with mudstones, chert, conglomerates and lithic sandstones
The San Jerónimo fault (SJF) separates the pre-Cretaceous metamor- (Gomez-Cruz et al., 1995; Gomez et al., 2002; Restrepo-Moreno et al.,
phic basement of the Central Cordillera from the Cauca Ophiolitic Com- 2009). The age of the Quebradagrande complex has been constrained
plex and the Quebradagrande Complex (Alvarez, 1983; Borrero et al., between 97 and 120 Ma based on fossil occurrences and two U-Pb zir-
2012; Maya and Gonzalez, 1995; Pardo-Trujillo et al., 2011; con ages from a tuff and a dioritic intrusive (Cochrane et al., 2014b;
Villagómez et al., 2011) (Figs. 1 and 3). The Cauca Ophiolitic complex Villagómez et al., 2011). However, younger Campanian-Maastrichtian
is composed of several discontinuous fault-bounded bodies compose fossil ages have also been reported in several localities suggesting a
of gabbroic plutons, basaltic pillow lavas, and ultramafic rocks; which more complicated accumulation history (Gomez et al., 2002;
have been considered as fragments of a dismembered ophiolitic com- González, 1980) (Fig. 3). Tonalitic and gabbroic plutonic bodies intrud-
plex. The age of this ophiolite has been assumed as Cretaceous due to ing the volcanic rocks of the Quebradagrande have U-Pb crystallization
its close relation with the Quebradagrande Complex (Alvarez, 1983; ages between 78 and 90 Ma (Jaramillo et al., 2017; Villagómez et al.,
González, 1980; Maya and Gonzalez, 1995). 2011). Previous tectonic models have related the Quebradagrande

A B
Jaramillo et al. 2017

SJF

CAF Fig 2B

Fig 4

CAF

Cambumbia
SPF Stock

Pacora
Stock
SJF

Miocene and Quaternary units


Paleogene plutonic rocks
Late Cretaceous plutonic rocks
SPF

Quebradagrande Complex
F

Cauca Ophiolithic Complex


SJ

Abejorral Fm. and other


siliciclastic Cretaceous rocks
Arquia Complex
Western Cordillera allochthonous
oceanic basement
Pre-Cretaceous continental basement
U-Pb geochronology Stratigraphic sections in Fig. 5 Upper Cretaceous fossils
Geochemestry Lower Cretaceous fossils

Fig. 3. (A). Simplified geological map of the Cauca Valley, the map includes published Cretaceous fossil locations (Gomez et al., 2002; González, 2001; González et al., 1988). Black boxes
indicate the location of the study area and the location of the data presented in Jaramillo et al. (2017); (B) geological map from the study area modified from Modified from Gómez et al.
(2015) and González (1980); stars represent the collected samples, and the numbered boxes represent the location of the stratigraphic sections presented in Fig. 5. The black box indicates
the location of the mapping area presented in Fig. 4. SJF: San Jerónimo Fault; SPF: Silvia Pijao Fault; and CAF: Cauca-Almaguer Fault.
S. Zapata et al. / Gondwana Research 66 (2019) 207–226 211

Complex and the Cauca Ophiolitic Complex to the formation of a back- 30–20 μm and 10 Hz, respectively. U and Th concentration were moni-
arc basin or to an accreted oceanic arc (Cochrane et al., 2014b; Nivia tored by comparing to NIST 610 trace element glass.
et al., 2006; Toussaint and Restrepo, 1996; Villagómez et al., 2011) In the case of the zircons recovered from magmatic rocks, the analy-
(Fig. 2). ses were conducted after reviewing cathodoluminescence (CL) analysis,
The Silvia Pijao Fault (SPF) separates the Quebradagrande complex and zircon rims and well-defined cores were used to constrain the zir-
on the east from different Triassic granitic rocks, including the Pacora con crystallization history (Valencia et al., 2005). In detrital samples,
and the Cambumbia Stocks. Also west of the SPF have been documented only the cores of the grains were analyzed to avoid complex zircon his-
Early Cretaceous high to middle-pressure metamorphic rocks grouped tories (Gehrels et al., 2006).
in the Arquía Complex (Bustamante et al., 2011b, 2011a; Cochrane For all samples, probability density plots and weight-average ages
et al., 2014a; Maya and Gonzalez, 1995; Ruiz-Jimenez et al., 2012; were obtained with ISOPLOT 3.62 (Ludwig, 2007). For the detrital sam-
Vinasco Vallejo and Cordani, 2012). The Triassic granites have been con- ples, only the grains with discordance b 20% and errors b 5% were con-
sidered as remnants of the continental margin separated during the sidered. The volcanic ages were obtained using the grains with the
opening of a back-arc basin (Nivia et al., 2006) (Fig. 2B). The Early Cre- lower errors and avoiding inherited ages. The maximum depositional
taceous metamorphic rocks of the Arquia Complex have been consid- ages were calculated using a weighted average age from the three youn-
ered as a part of a subduction/accretion complex formed during the gest zircons that overlap at 2 sigma (Dickinson and Gehrels, 2009). An-
growth of the Quebradagrande oceanic arc (Cochrane et al., 2014a; alytical results are presented in Table A1.
Spikings et al., 2015) (Fig. 2A).
Finally, the Cauca Almaguer Fault (CAF) separates the units em- 3.3. Whole-rock geochemistry
bedded within the Romeral Fault System from the Cretaceous rocks
from the Western Cordillera (Aspden et al., 1987; Barrero, 1979; Whole rock chemical analyses were conducted in volcanic and
Kerr et al., 1997) (Figs. 1 and 3). The Western Cordillera basement in- plutonic rocks from the Abejorral Fm. and the Cauca Ophiolitic Com-
cludes mafic rocks interbedded with deep-marine sedimentary rocks plex. Samples were analyzed with an inductively coupled plasma
deposited during the formation of an Early Cretaceous oceanic pla- mass spectrometry (ICPMS) at Acme Analytical Laboratories Ltd. in
teau, in a position southwest of the present day position. The Vancouver, Canada. A 0.2 g aliquot was weighed into a graphite cru-
volcano-sedimentary basement of the Western Cordillera was in- cible and mixed with 1.5 g of LiBO2 flux. The crucibles are placed in
truded by oceanic arc plutons during the Late Cretaceous (Escalona an oven and heated to 1050 °C for 15 min. The molten sample was
and Mann, 2011; Rodriguez and Zapata, 2013; Villagómez et al., dissolved in 5% HNO 3. Calibration standards and reagent blanks
2011; Weber et al., 2015; Zapata, 2009). This oceanic plateau col- were added to the sample sequence. Sample solutions were aspi-
lided with the northwestern margin of South America between the rated into an ICP emission spectrograph (Jarrel Ash Atom Comb
Late Cretaceous and the Paleocene (Bayona et al., 2011; Escalona 975) for determining major oxides and certain trace elements (Ba,
and Mann, 2011; Hincapié-Gómez et al., 2018; Pindell et al., 1998; Nb, Ni, Sr, Sc, Y and Zr), while the sample solutions are aspirated
Spikings et al., 2015) (Fig. 2). into an ICP-MS (Perkins–Elmer Elan 6000) for determination of the
trace elements, including rare earth elements. The Results are pre-
3. Methods sented in the Table A2. Standard SO-8 were measured (Table A3)
with standard deviations of 0.01 for TiO 2 , 0.753 for Nb and 0.739
Geological mapping was carried out in an area of ~60 km (Fig. 4). for Y.
Representative samples from the Quebradagrande Complex, Abejorral
Fm., Cauca Ophiolitic Complex, and the Cajamarca Complex were col- 3.4. Nd isotopes
lected for petrography, U-Pb zircon geochronology, whole rock geo-
chemistry, and Nd isotopes. Complementary cartographic observations The analyses were done on a Thermo-Finnigan Neptune
and sampling were also conducted up to 50 km south of this area multicollector system at Washington State University in Pullman, and
(Fig. 3B). the results are presented in Table 3. The procedures for sample prepara-
tion and Nd dilution can be found in Gaschnig et al. (2011). The Sm and
3.1. Conglomerate clast counting and sandstone petrography Nd isotope analyses followed procedures described in Vervoort and
Blichert-Toft (1999). Sm concentrations were corrected using
147
Conglomerate clast counting was performed following the ribbon Sm/152Sm: 0.56081, and Nd was corrected for mass fractionation
counting method (Howard, 1993). Clasts b 2 cm in size were excluded using 146Nd/144Nd: 0.7219 and normalized using the Ames Nd standard
from the analysis results are presented in Table 1. 300 points were (±0.000020 2σ average reproducibility). The εNd values were calcu-
counted in sandstone samples following both the Gazzi-Dickinson and lated using present-day values of 143Nd/144Nd: 0.512630 and
147
the Indiana methods (Dickinson, 1985). Additional high-resolution pet- Sm/144Nd: 0.160 for CHUR (Bouvier et al., 2008).
rographic discrimination of quartz was performed according to the
methodology proposed by Basu et al. (1975). Petrographic results are 4. Field observations and petrography
presented in Table 2.
Most of the units within the study area record various stages of duc-
3.2. U-Pb geochronology tile and brittle deformation, which resulted in highly deformed rocks.
The main characteristics of four major units are described from the
U-Pb geochronology samples include Granitoids from the Cajamarca east to the west as follows (Fig. 4).
Complex and volcanic and siliciclastic rocks from the Quebradagrande
Complex and the Abejorral Fm. After cathodoluminescence imaging, 4.1. Gneisses, granitoids, and schists
the Laser Ablation Inductively Coupled Plasma Mass Spectrometry
(LA-ICP-MS) U-Pb analyses were conducted at Washington State Uni- Gneisses and schistose rocks with an NW-SE foliation were mapped
versity. The analyses were conducted at Washington State University within the study area (Fig. 4). The gneissic rocks include two different
using a New Wave Nd:YAG UV 213-nm laser coupled to a Thermo foliated metagranitoids locally known as the Pantanillo and Abejorral
Finnigan Element 2 single collector, double-focusing, magnetic sector gneisses. These gneisses are composed of quartz (4–56%), plagioclase
ICP-MS at WSU. Operating procedures and parameters were similar to (6–10%), K-feldspar (10–15%), and different amounts of muscovite
those of Chang et al. (2006). Laser spot sizes and repetition rates were and biotite with minor sillimanite (24–12%) defining the main foliation.
212 S. Zapata et al. / Gondwana Research 66 (2019) 207–226

Fig. 4. Geological map and cross-section of the mapping area. White and black stars denote the sample location. SJF: San Jeronimo Fault.

The Abejorral and the Pantanillo gneisses have yield Permo-Triassic U- + feldspar schist with muscovite + quartz + feldspar schists. Both
Pb zircon crystallization ages (Vinasco et al., 2006). The schists include units are included in the Cajamarca Complex (Maya and Gonzalez,
centimetric to decametric intercalations of chlorite + amphibole 1995), which is the oldest basement of the Central Cordillera.

Table 1
Results of conglomerate clast counting analysis.

Unit Quartz (%) Sandstones (%) Mudstones (%) Chert (%) Metamorphic (%) Plutonic (%) Volcanic (%)

Lower Abejorral Member 85 0 14 0 0 0 0


Upper Abejorral Member 32 16 16 0 25 2 9
Quebradagrande Complex 6 45 5 26 1 0 17
S. Zapata et al. / Gondwana Research 66 (2019) 207–226 213

Table 2
Sandstone framework analysis. Qm: monocrystalline quartz; Qsed: polycrystalline sedimentary quartz; Qpd: polycrystalline diffuse quartz; Qpf: polycrystalline foliated quartz; Qct: chert;
Pl: plagioclase; Fk: orthoclase; Lsl: sedimentary lithic siltstone; Lsl: sedimentary lithic mudstone; Lmm: muscovite metamorphic lithic; Lmb: biotite metamorphic lithic; Lmc: chlorite
metamorphic lithic; Lma: amphibolitic metamorphic lithic; Lmn: gneissic metamorphic lithic; Lh: hypabyssal lithic; Lvf: felsic volcanic lithic; Lvv: vitreous volcanic lithic; Msc: muscovite;
Cl: chlorite; Px: pyroxene; Bt: biotite; Hm: heavy minerals; opaques; Qp2–3: polycrystalline quartz 2–3 grains; Qp + 3: polycrystalline quartz N3 grains; Qn: non-undulatory quartz; and
Qund: undulatory extension quartz.

West Lower Abejorral Member Upper Abejorral Member Quebradagrande


Abejorral Complex

SL-046 SZ-009F SZ-009a AP-005 SZ-009e SZ-009h AP-034 AP-036 SZ-022 SZ-027 JPC 002 JPC012

Qm (%) 54 39 40 59 70 42 27 23 58 49 68 58
Qsed (%) 1 0 0 0 0 1 1 1 1 0 9 6
Qpd (%) 0 2 1 5 3 4 15 15 1 7 0 0
Qpf (%) 0 0 0 1 1 0 3 7 0 0 0 0
Qpoli 9 24 25 11 18 17 11 9 1 21 0 0
(%)
QCt (%) 0 26 15 2 3 11 16 6 0 2 0 0
Pl (%) 0 0 0 0 0 0 1 2 0 0 5 19
F (%) 0 0 3 3 0 0 0 0 0 0 0 0
Lss (%) 7 0 0 1 0 0 6 11 0 1 0 6
Lsl (%) 8 0 0 0 0 0 6 1 0 0 0 0
Lmm (%) 3 0 0 5 0 0 0 6 2 1 0 5
Lmg (%) 0 0 0 0 0 0 0 1 1 0 0 0
Lmb (%) 0 0 0 0 0 0 3 0 0 0 0 0
Lmc (%) 0 0 0 0 0 0 3 4 0 0 0 0
Lma (%) 0 0 0 0 0 0 0 7 0 0 0 0
Lmn (%) 0 0 0 0 0 0 0 1 0 2 0 0
Lp (%) 0 0 0 0 0 0 4 1 0 1 0 0
Lvf (%) 0 0 0 0 0 0 0 1 0 0 0 0
Lvv (%) 0 0 0 0 0 0 0 0 2 2 0 0
Lvi (%) 0 0 0 0 0 0 2 4 0 1 0 0
Msc (%) 6 5 16 11 5 12 1 0 25 6 12 3
Cl (%) 10 0 0 0 0 0 0 3 0 0 0 0
PX (%) 0 0 0 0 0 0 0 0 2 1 0 0
Bt (%) 0 2 0 0 0 10 1 0 7 6 0 0
Hm (%) 0 0 1 1 0 1 0 0 1 0 0 0
Op (%) 3 2 0 0 0 2 0 0 0 0 6 4

High resolution quartz analysis (Basu et al., 1975)

West Abejorral Lower Abejorral Member Upper Abejorral Member Quebradagrande


Complex

SL-046 SZ-009F SZ-009a AP-005 SZ-009e SZ-009h AP-034 AP-036 SZ-022 SZ-027 JPC 002 JPC012

Qp2–3 (%) 6 18 17 7 9 20 1 6 1 15 4 4
Qp + 3 (%) 7 38 32 17 18 23 61 58 4 24 3 6
Qn (%) 6 31 23 27 29 34 16 11 82 40 23 75
Qund (%) 81 13 27 49 44 23 22 25 13 21 71 16

4.2. Abejorral Formation as the most abundant component. This member rests on top of
the Abejorral gneiss on the east and is overthrusting the schists of
Due to the discontinuous nature of the field exposures and the over- the Cajamarca Complex in the west (Fig. 4). (2) The Upper
imposed deformation (i.e. folding and faulting), we were not able to Abejorral Member is mainly composed of thin tabular beds of
measure a continuous stratigraphic section within the study area. How- black mudstones with planar lamination, interbedded with thin
ever, the more representative lithologies of this formation were levels of medium to fine-grained muddy sandstones and metric
mapped, sampled, described, and correlated with more continuous pub- levels (1–5 m) of tabular beds of black chert and siliceous
lished stratigraphic sections, in equivalent structural positions outside mudstones. A series of andesitic lava flows and pyroclastic beds
of the study area (Figs. 3B and 5). are interbedded with the sedimentary rocks of the Upper Abejorral
Siliciclastic rocks of the Abejorral Formation can be divided into Member. Additionally, several metric-scale sub-volcanic rocks
two informal members. (1) The Lower Abejorral Member includes were observed intruding mudstones of the Upper Abejorral
thick to very thick (0.5–3 m) tabular beds of structureless matrix- Member. These volcanic rocks exhibit porphyritic and
supported and poorly-sorted conglomerates, with pebble-sized glomeroporphydic textures. Plagioclase is euhedral and is replaced
rounded clasts and a medium-grained sandy matrix, with quartz by carbonates. The pyroxenes are replaced by chlorite.

Table 3
Sm-Nd isotopic composition of the Cretaceous igneous rocks from the Central Cordillera.
147
Sample Sm (ppm) Nd (ppm) Sm/144Nd Error abs 143
Nd/144Nd Error abs ε(0) T1 (Ma) ε(T1)

SL-0013 3.02 8.95 0.20430 0.000007 0.513149 0.000008 10.12 100 10.027
JS-050 0.82 1.7 0.29200 0.000024 0.513215 0.000007 11.41 100 10.318
DM-017 1.65 3.89 0.25760 0.000004 0.513164 0.000008 10.41 100 9.679
JS-023 1.41 3.69 0.23090 0.000004 0.513133 0.000007 9.812 100 9.420
SZ-012 1.9 6.95 0.16531 0.513046 8.115 112 8.563
214 S. Zapata et al. / Gondwana Research 66 (2019) 207–226

~50 km ~3 km ~5 km
A) Section 1 B) Section 2 C) Section 3 D) Section 4 ~30 km E) Section 5
Quebrada Honda stratigraphic section Rio Pozito stratigraphic section Guayaquil stratigraphic section Monteverde stratigraphic section Quebrada Arenosa stratigraphic section
Gomez et al.1995

Lithology
Thickness
Grain size Grain size

Lithology
Lithology

Lithology
Thickness

Grain si ze Grain size

Lithology
Thickness
Thickness
Grain size

Thickness
(m)
(m)

(m)

(m)
vcS

(m)
mS

Cobb
Pebb
Gran
Clay

vfS

vcS
mS

Cobb
Pebb

vcS
vcS

Gran

vcS
Silt

Clay

vfS

mS
mS

mS
Cobb
Cobb

cS

Cobb
Pebb
Pebb

fS

Pebb
Gran
Gran

Clay

Gran
Clay

vfS

Clay
vfS

vfS
Silt

Silt
Silt

cS
cS

fS
fS

200

2000

250
200
200
100
1500

Marine shelf
Albian

Marine shelf

Marine shelf
200
150
100
Marine shelf

1000

150

Proximal alluvial
100 50

Proximal alluvial

Proximal turbiditic fans


50
500
Shoreface

50
50
0

Berrasian

10 0
0

0
0

Upper Abejorral Member (110 - 100 Ma). Lower Abejorral Member (130 - 110 Ma).
Chert and siliceous Medium-Fine sandstone Mudstone
mudrock Coarse sandstone Conglomerate

Ammonites Bivalves Plant remains Cross stratification Trough stratification Planar lamination

Fig. 5. Stratigraphic sections of the Abejorral Fm. and correlatable sedimentary formations north and south of the studied region. The locations of the stratigraphic section are presented in
Fig. 3B. The stratigraphic columns are arranged from north to south. The distances between the sections are indicated on the top of the figure.

Deformational structures within the two members include asym- fabrics. The gabbros are massive and texturally heterogeneous, vary-
metric folds with rounded hinges accompanied by several NW in- ing from fine-grained to pegmatitic. They are composed of plagio-
verse fault planes dipping to the northeast (Fig. 4). clase (28–59%) and augite pyroxene (9–33%), which ophitic and
This informal subdivision between Lower and the Upper members subophitic textures and opaques (1–5%). The pyroxene is replaced
can also be applied to the published stratigraphic sections, north and by amphibole (hornblende or actinolite). Plagioclase composition is
south of the study area (Gomez et al., 1995; Gomez et al., 2002; labradorite (An50-An60) but is commonly replaced by pumpellyite
González, 1980; Quiroz, 2005) (Figs. 3B and 5). Within these strati- and epidote. Intrusive relations with the basaltic rocks and the mica-
graphic sections, the Lower Member is composed of oligomictic ceous schist of the Cajamarca Complex are marked by the presence of
quartz-conglomerates interlayered with sandstones and is unconform- small dikes, silicification, and crystal size variations near the
ably accumulated on top of metamorphic rocks of the Cajamarca Com- contacts.
plex; abundant plant remains have been described within this The volcanic rocks are composed of clinopyroxene (35%) with
member. The Upper Member is composed of fine mudstones and variolitic and ophitic textures commonly replaced by amphibole,
minor siltstones, which include Aptian-Albian gastropods and ammo- saussuritized plagioclase (39%) with labradorite composition (An55)
nites (Gomez et al., 2002; González, 1980) (Fig. 5). embedded within an aphanitic matrix (6%). The clinopyroxenes exhibit
ophitic textures, whereas olivine (3%) is highly altered to serpentine,
4.3. Cauca Ophiolitic Complex calcite and chlorite.

This unit is composed of discontinuously exposed gabbroic 4.4. Quebradagrande Complex


serpentized peridotites and basaltic rocks in fault-contact with
volcano-sedimentary rocks from the Quebradagrande Complex The volcanic and sedimentary rocks from the Quebradagrande Com-
(Fig. 4). This faulted contact is characterized by local mylonitic plex crop out in the western segment of the study area in the Campanas
S. Zapata et al. / Gondwana Research 66 (2019) 207–226 215

creek (Fig. 4). Based on the dominant lithology a volcanic and a sedi- a succession of black shales, chert, very fine grain sandstones, and lava
mentary member can be discriminated. The volcanic member is mainly flows. Aptian-Albian fossils have been recognized in the shale levels of
composed of andesitic and basaltic lava flows locally autobrecciated, this succession (González, 1980) (Fig. 4).
with interbedded agglomerates (Fig. 6D). These volcanic beds exhibit Towards the west, this succession has fault contacts with the Triassic
porphyritic and glomeroporphiric textures, with plagioclase and pyrox- plutonic rocks of the Pacora and Cambumbia stocks (Fig. 3B). Farther
ene as phenocrysts in a crystalline or vitreous matrix. Black to gray lam- south similar successions of mudstones, sandstones and basaltic rocks
inated beds of mudstones are intercalated with these volcanic rocks. with Albian-Aptian fossils have been described (Alvarez, 1983)
Locally these units are found as proto-mylonites and mylonites, with (Fig. 3B). These rocks have been included as part of the Quebradagrande
neoformed chlorite and white mica surrounding feldspar Complex (González, 1980; Maya and Gonzalez, 1995). However, the de-
porphyroclasts. The sedimentary member is composed of thick beds of scribed angular discordance with the Quebradagrande Complex and the
sedimentary breccias, coarse-grained to fine-grained conglomeratic temporal constraints suggest that this unit is correlatable with the
sandstones, interbedded with thinner beds of laminated black and Abejorral Fm.
gray mudstones.
5. Conglomerate clast counts
4.5. West Abejorral Fm.
Clast counts were performed in conglomeratic beds of the Lower and
In the west of the studied area in the Campanas creek, the volcanic the Upper Abejorral Members and in the Quebradagrande Complex
rocks of the Quebradagrande Complex are unconformably overlaying (Fig. 6A, Table 1). The Lower Abejorral Member includes pebble-sized

Sandstones and
A) Clast Counts Quartz conglomerates Chert
Metamorphic rocks Mudstones Volcanic rocks
Quebradagrande

Upper Abejorral

Lower Abejorral

0%
Polycrystalline quartz 100%
(2-3 crystal units per grain)
B) Q C) Q D)
Int aton
or

Quartz arenite
eri

95
Cr

Subarkose

hic
Sublitharenite
ne al

orp
l
nti on
nta

75

tam
Co nsiti

Non-undulatory

Undulatory
me
ic
Feldspathic

quartz
quartz
Tra

ton

igh
Plu

m-h
ft
pli

diu

hic
ose

Lith

tU

Me

orp
Litharenite
c Ark

en
aren

tam
ose

sem
ite

me
Lithi
Ark

Volcanic Rifted Margin


Ba

Low

F 3:1 1:1 1:3 L F L Polycrystalline quartz


Upper Abejorral Lower Abejorral Quebradagrande (>3 crystal units per grain)
West
Abejorral Member Member Complex

E F LV

Qm

Px
LM

500 um
500 um

Fig. 6. (A), Clast counts from the Abejorral Fm. and the Quebradagrande Complex; (B), sandstone classification after Folk (1980); (C), rifted margin provenances after Garzanti et al. (2006);
(D), quartz classification after Basu et al. (1975); (E), polarized thin section from the Upper Abejorral Fm., sample Ap-009e. Lm: metamorphic lithic and Qm: monocrystalline quartz; (F),
polarized thin section from the Upper Abejorral M., sample SZ-027. Lv: volcanic lithic and Px: pyroxene.
216 S. Zapata et al. / Gondwana Research 66 (2019) 207–226

rounded clasts composed of mudstone (15%) and milky colored quartz 7. Geochronology
(85%). The Upper Abejorral Member is characterized by a more diverse
clast composition including milky quartz (32%), sandstones (16%), 7.1. Pre-Cretaceous basement
schistose and gneissic rocks (25%), andesites (9%), and plutonic lithics
(2%). The massive angular breccias of the Quebradagrande Complex in- U-Pb zircon ages were obtained from igneous rocks located in the
clude sandstone (45%), chert (26%), and volcanic fragments (17%), as westernmost segment of the RFS, immediately west of the Abejorral
well as low contents of milky quartz (6%), mudstones (5%) and mica- Formation, the Cauca Ophiolitic Complex and the Quebradagrande
ceous schists (5%). Complex (Fig. 3B). A total of 52 individual zircon crystals from the
Pácora Stock were analyzed (sample SL-057, Fig. 7). Crystal sizes varied
between 30 and 160 μm, with length/width ratios between 1:1 and 1:2.
6. Sandstone petrography Most of the zircon crystals are characterized by oscillatory and zoned
cores with rim overgrowths. Whereas, a portion of the grains is charac-
Sandstones from the Lower Abejorral Member are poorly to very terized by single oscillatory zoning patterns. Th/U ratios are between 0.2
poorly-sorted, medium to coarse-grained, with angular to subangular and 1.55, which is typical of zircons formed in igneous environments
grains in a clay matrix (5–15%). Compositionally, the sandstones are (Rubatto, 2002). Twenty-one zircon crystals yielded a weighted age of
classified as quartz-arenites (Fig. 6B and E). They include abundant 260.5 ± 4.7 Ma (MSWD: 5.2), which we interpreted as the time of igne-
phyllosilicates such as muscovite, biotite, and chlorite (7–23%) and ous crystallization of the sample (Fig. 7). Zircons with older ages of
high contents of mono and polycrystalline quartz (N80%) (Fig. 6B and 530 Ma, 660 Ma and 1210 Ma ages are considered inherited crystals
C). The high-resolution discrimination of quartz after Basu et al. (Table A1).
(1975) shows the presence of monocrystalline quartz with undulatory Sample SG-013 is a granodiorite collected from the Cambumbia
extinction and polycrystalline quartz clasts with more than two grains Stock (Fig. 3B). Thirty-eight analyzed zircon are characterized by sizes
(Fig. 6D). between 50 and 170 μm and are predominantly prismatic with
Sandstones of the Upper Abejorral Member are fine to medium- length/width ratios between of 3:1 and 2:1. CL images showed single
grained, with rounded to subrounded grains in a clay matrix oscillatory zoning patterns. Th/U ratios are between 0.2 and 0.7; these
(≤15%) and are classified as sub-lithoarenites and lithoarenites values are characteristic of igneous zircons (Rubatto, 2002). This rock
(Fig. 6B). Compositionally, the samples include sedimentary lithic yielded an age of 232.9 ± 1.2 Ma (MSWD: 1.8), which is considered to
(quartz-arenites and mudstones up to (15%), metamorphic lithics be the magmatic crystallization age (Fig. 7). Zircons cores older than
(micaceous schists, 8–30%), and lower quartz contents (50–80%) 332 Ma, 355 Ma, and 983 Ma are considered inherited ages (Table A1).
compared to the Lower Member (80–100%). In the quartz type dis-
crimination diagram after Basu et al. (1975), the Abejorral Upper 7.2. Lower Abejorral Member
Member shows affinity with both metamorphic and plutonic sources
(Fig. 6D, Table 2). U-Pb data was obtained from two detrital samples of quartz sand-
A sandstone sample from the West Abejorral Fm. is characterized by stones and a quartz-conglomerate (Fig. 4). A total of 207 single grain
poor to moderately sorted material with rounded to subrounded grains U-Pb ages were obtained from the sample PAN-1. The age distribution
and 10–12% of 7 matrix. Compositionally, the sample is classified as a is characterized by a major Triassic population of ca. 241 Ma, with
sub-litharenite (Fig. 6B), with mono and polycrystalline quartz charac- minor Jurassic (153 Ma), Paleozoic (537 Ma) and Mesoproterozoic
teristic of low to high-grade metamorphic rocks (Basu et al., 1975; ages (1006.5 and 1189.5 Ma). The youngest population has an age of
Bouvier et al., 2008) (Fig. 6C). 149.5 ± 2.7 Ma (Fig. 8). For sample SZ-008, a total of 94 single grain
We analyzed two sandstone samples of the Quebradagrande U-Pb ages were obtained. The age distribution is characterized by
Complex. They are fine-grained sandstones, with angular to highly minor Mesoproterozoic-Neoproterozoic age populations (955 and
angular spherical grains, moderately sorted and matrix content be- 1168 Ma) and a major Permian peak (275.8 Ma), which is the youngest
tween 5 and 8%. Compositionally, they are classified as subarkose population (Fig. 8).
and lithic arkose (Fig. 6B) and are characterized by the presence of
plagioclase (5–19%), sedimentary quartz (6–9%). In the quartz type 7.3. Upper Abejorral Member
discrimination diagram after Basu et al. (1975), the Quebradagrande
Complex shows affinity with both metamorphic and plutonic Three samples of medium-grained lithic sandstones were analyzed
sources (Fig. 6D, Table 2). (Fig. 9). A total of 89, 140 and 104 individual U-Pb ages were obtained

SL-057, Pacora Stock. SG-013, Cambumbia Stock.


295 244

n=52 n=38
240
285

236
275

232

265

228

255
224

245
220

Mean Age= 260.5 ± 4.7 Ma 100 um Mean Age= 232.9 ± 1.2 Ma 100 um
235 216

Fig. 7. U-Pb zircon ages and cathodoluminescence images acquired from Pre-Cretaceous basement west of the Silvia Pijao Fault.
S. Zapata et al. / Gondwana Research 66 (2019) 207–226 217

PAN - 1, Lower Abejorral M. SZ-008, Lower Abejorral M.


Max.Dep : 149.5 +/- 2.7 Ma 241.5 Max.Dep : 236.1+/- 3.45 Ma
275.8
n=207 n=94

40
277.5
241.5
153
275
30
30
184.5

127.5
20 153 20

537 1006
10 10 954 1167.6
1189

0
0 500 1000 1500 2000 2500 3000 3500 0
0 500 1000 1500 2000 2500 3000
Time (Ma) Time (Ma)

SL-046, West Abejorral. JPC-002, West Abejorral.


Max.Dep : 488.1+/- 18 Ma 265
n=92
Max. Dep : 253.3+/- 4.6 Ma
n=107
556
536
160
40 20
265
30 1066

20
10
1050
10

0
0 500 1000 1500 2000 2500 3000 3500 4000 0
0 500 1000 1500 2000 2500 3000 3500 4000
Time (Ma) Time (Ma)
Archean - Neoproterozoic-Paleozoic Paleozoic-Triassic
Neoproterozoic Jurassic
Mezoproterozoic (1250 - 800 Ma)
a) (350 - 800 Ma)
a) (350 - 200 Ma)
a) (200 - 150 Ma)
a)
(3000 - 1250Ma)
a)

Lower
er Creta
etaceous
eous Aptian-Albian Coniacian-Santonian
(150 - 110 Ma)
a) (110 - 90 Ma)
a) (90 - 80 Ma)
a)

Fig. 8. U-Pb detrital zircon ages acquired in samples from the Lower Abejorral Member and the West Abejorral Fm.; detrital populations are presented for each sample.

from samples AP-33, PAN-2 and SZ-023, respectively. All three samples prismatic crystals. The zircons showed complex zoned patterns includ-
are characterized by Early Cretaceous zircon U-Pb age populations of ing homogeneous and oscillatory zoning, with Th/U ratios between 0.09
123.5 Ma (sample AP-33), 103.9 Ma (sample PAN-2) and 104.3 Ma and 0.85. Inherited zircon crystals have ages between 343 Ma and
(sample SZ-023) (Fig. 9). Older zircon U-Pb ages include relatively sim- 2240 Ma. The youngest Cretaceous zircons (n: 27) yielded a weighted
ilar populations of Premo-Triassic (271.25 Ma and 249 Ma), Paleozoic mean age of 111.5 ± 7.9 Ma interpreted as the magmatic crystallization
(463, 512, 520, and 535 Ma). As well as and Paleoproterozoic (1052, age (Fig. 10).
1205, and 1225 Ma), and Archean (3122 Ma) detrital populations Sample SZ-011 is a weathered sample that preserves igneous por-
(Fig. 9). phyritic textures and intrudes the mudstones and fine-grained sand-
stones from the Upper Abejorral Member (Fig. 4). Zircon crystals sizes
7.4. Volcanic rocks of the Upper Abejorral Member are between 50 and 300 μm, with length/width ratios of 3:1. CL images
revealed oscillatory zoning patterns. The Th/U ratios are between 0.2
Zircons from two andesites interlayered with sandstones and mud- and 0.6, which is characteristic of magmatic zircons (Rubatto, 2002;
stones of the Upper Abejorral Member and one porphyritic intrusive Vavra et al., 1999). 16 zircons yielded a weighted mean U-Pb age of
layer were analyzed (Fig. 4). A total of 28 zircon crystals were analyzed 103.5 ± 1.8 Ma interpreted as the magmatic crystallization age (Fig. 10).
from the andesite sample (SZ-012) (Fig. 10). Zircons are between 30
and 250 μm in size. Prismatic zircons have length/width ratios of 3:1. 7.5. West Abejorral Fm.
CL images showed a single oscillatory zoning pattern, characteristic of
an igneous origin (Vavra et al., 1999). Th/U ratios are between 0.2 and We analyzed two sandstones and one andesitic lava from this unit,
0.6, these values are characteristic of magmatic zircons (Rubatto, which is overlain in angular discordance by the volcanic rocks of the
2002). Twenty-two zircon crystals from this sample yielded a weighted Quebradagrande Complex (Figs. 3 and 4). A total of 92 single grain U-
mean age of 103.1 ± 1.5 Ma interpreted as the crystallization age Pb ages were obtained from a quartzose sandstone (sample SL-046),
(Fig. 10). Older ages in zircon cores are Triassic (225 Ma and 238 Ma) the age distribution is characterized by major Neoproterozoic age pop-
and Paleozoic (411 Ma) and are interpreted as inherited ages ulations of ca. 556 Ma and 1050 Ma and less abundant Proterozoic to Ar-
(Table A1). chean ages (2000–3500 Ma) (Fig. 8). Sample JPC-002 is also a quartzose
Sample SZ-018 is an andesitic tuff intercalated within the Upper sandstone. A total of 107 individual zircon U-Pb ages are between 160
Abejorral Member sedimentary rocks (Fig. 4). Zircon crystal sizes are and 3339 Ma with major peaks at 160, 263, 535, and 974–1223 Ma
between 30 and 100 μm, with a length/width ratio of 2:1 for the (Fig. 8).
218 S. Zapata et al. / Gondwana Research 66 (2019) 207–226

AP-033, Upper Abejorral M. SZ-023, Upper Abejorral M.


124.8 154.7
Max.Dep : 123.5+/- 2.3 Ma Max.Dep : 104.3 +/- 2.2 Ma 105 271.25
n=89 n=104

10
124.8 105
462.8 12
271.2 255.5

8 519.7

6 8
5
1051.7
1205.1
4 512.2 1162
1846 1764 3122
4
2

0 0
0 500 1000 1500 2000 2500 3000 3500 4000
0 500 1000 1500 2000 2500 3000
Time (Ma) Time (Ma)

PAN-2, Upper Abejorral M. JPC-008, Quebradagrande C.


Max. Dep : 103.9 +/- 2.8 Ma
Max.Dep : 84.0+/- 0.5 Ma
261.8
n=77
n=140 103.5 249
18
261.8
276
103.5
535.5 30
239.4
10 249
81.2
1225.5 20
1062
6
10 81.2 1016.4
1759.5 561.4

0 0
0 500 1000 1500 2000 2500 3000 3500 0 500 1000 1500 2000 2500 3000
Time (Ma) Time (Ma)
Archean - Neoproterozoic-Paleozoic Paleozoic-Triassic
Neoproterozoic Jurassic
Mezoproterozoic (1250 - 800 Ma)
a) (350 - 800 Ma)
a) (350 - 200 Ma)
a) (200 - 150 Ma)
a)
(3000 - 1250Ma)
a)

er Creta
Lower etaceous
eous Aptian-Albian Coniacian-Santonian
(150 - 110 Ma)
a) (110 - 90 Ma)
a) (90 - 80 Ma)
a)

Fig. 9. U-Pb detrital zircon ages acquired in samples from the Upper Abejorral Member and the Quebradagrande Complex; detrital populations are presented for each sample.

Sample M9 is andesitic lava interlayered with the analyzed sand- crystallization age. Older ages in zircon cores yielded Cretaceous and
stones from the West Abejorral Fm. (Fig. 4). Zircon crystal sizes are be- Proterozoic ages considered inherited zircon ages (Fig. 10).
tween 50 and 150 μm, with length/width ratios between 2:1 and 3:1. CL Sample SL-029 is a dacitic crystal-rich tuff interlayered with fine to
images revealed a simple oscillatory zoning pattern. Th/U ratios are be- medium-grained sediments from the Quebradagrande Complex
tween 0.31 and 0.86. Ten zircons yielded a weighted mean age of 115.7 (Fig. 3). Zircons have crystal sizes between 60 and 130 μm, with
± 7.7 Ma, which is considered to be the maximum age of crystallization length/width ratios of 2:1. CL images revealed simple oscillatory zoning.
(Fig. 8). Inherited zircon crystals yielded Cretaceous and Jurassic ages Th/U ratios varied between 0.1 and 1.1, which are characteristic of igne-
(130 to 184 Ma) (Table A1). ous zircons. This rock yielded a U-Pb zircon age of 85.7 ± 0.4 Ma from 86
zircon crystals (Fig. 10), which is interpreted as the age of crystalliza-
tion. Older ages in zircon cores with Cretaceous, Paleozoic and Protero-
7.6. Quebradagrande complex zoic ages are considered inherited.

Sample JPC-008 is a quartzose sandstone from the sedimentary 8. Geochemistry of volcanic and plutonic rocks
member of the Quebradagrande Complex (Fig. 4). A total of 77 individ-
ual zircon U-Pb ages have ages between 81 Ma and 2700 Ma, with age Major and trace element geochemistry of the igneous rocks from of
peaks at 81, 239, 262 Ma and 1020 Ma (Fig. 9). The youngest detrital zir- the Upper Abejorral Fm. and rocks of the Cauca Ophiolitic Complex are
con age population had a weighted mean age of 84.0 ± 0.5 Ma, which is presented in Table A2. Lost on ignition values from the analyzed sam-
considered the maximum depositional age. ples are around 3.6%. Therefore, we have used immobile trace elements
The sample JPC-13 corresponds to an andesitic lava from the (Pearce, 2008) to avoid the possibility of major element and trace ele-
Quebradagrande Complex (Fig. 4). Zircon crystals have sizes between ment mobility during hydrothermal or low-grade metamorphism.
50 and 170 μm, with length/width ratios between 2:1 and 3:1. CL im-
ages revealed a single oscillatory zoning pattern and Th/U ratios are be- 8.1. Upper Abejorral Member
tween 0.3 and 2.1, which are typical of igneous zircons (Rubatto, 2002).
This sample yielded a U-Pb zircon weighted mean age of 83.2 ± 0.7 Ma Samples from the Upper Abejorral Member are characterized by
obtained from 63 zircon crystals, which is considered the magmatic SiO2 values between 56.84% and 60.35% with total alkali values (Na2O
S. Zapata et al. / Gondwana Research 66 (2019) 207–226 219

Fig. 10. U-Pb mean zircon ages acquired in the volcanic rocks of the Upper Abejorral Member and the Quebradagrande Complex. Cathodoluminescence images are presented for each
sample.

+ K2O) between 4.18% and 4.99% and are classified as andesites. Fe2O3 characterized by enrichment in the Large Ion Lithophile Elements
values are between 4.89% to 5.07% and the magnesium number (Mg#) (LILE) and a well-defined depletion in High Field Strength Elements
between 43.3 and 49.1, which are typical of rocks derived from calc- (HFSE), including a negative anomaly of Nb and Ti (Fig. 11E). These sig-
alkaline magmas. Trace element diagrams from these samples also fol- natures are characteristic of mantle-derived rocks formed in convergent
low an intermediate–calc-alkaline trend in the immobile element margins, as is also suggested by the calc-alkaline character of the ana-
ratio-based Nb/Y vs. Zr/Ti and V vs. Ti discrimination diagrams lyzed samples and the tectonic setting discrimination diagrams
(Fig. 11A and B). (Fig. 11B).
Rare earth element (REE) abundances are characterized by a weak
enrichment in Light Rare Earth Elements (LREE), with (La/Sm)n ratios 8.2. Cauca Ophiolitic Complex
between 1.07 and 1.37, a relatively flat trend of the Heavy Rare Earth El-
ements (HREE), with (Gd/Yb)n values between 1.09 and 1.19, and a In contrast to the intermediate composition of the volcanic rocks of
negative Eu anomaly between 0.81 and 0.83 (Fig. 11C). Multi- the Upper Abejorral Member, the volcanic and plutonic rocks of the
elemental diagrams normalized to the primitive mantle are Cauca Ophiolitic Complex are characterized by a more basic
220 S. Zapata et al. / Gondwana Research 66 (2019) 207–226

500
A B

AB l)
d F ima
Alkali

an rox
rhyolite

0.500

ites

BB b-p
Phonolite 400

Bonin

BA Sla
(&

F al
Rhyolite

)
d ist
AB
Trachyte

BB sla R
V(ppm)

IAT

an b-d
BA nd MO
Dacite e
nolit
r ipho 300
Teph
0.050

Trachy

(a
andesite
Zr/Ti

site
Ande ndesite
ltic a 200
Basa
0.005

Alkali 100 Quebradagrande C. from


Foiditic Rodriguez et al. (2016)
basalt
Basalt Quebradagrande C. from
Jaramillo et al. (2017)
0.001

0
0.01 0.10 1.0 10 0 5 10

Nb/Y Ti(ppm)/1000
100 100
C D

Sample/REE Chondrite
Sample/REE Chondrite

10

1 1

0.1 0.1
La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu

100 100
E F
Sample/MORB

Sample/MORB

10 10

1 1

0.1 0.1

0.01 0.01
Sr K Rb Ba Th Ta Nb Ce P Zr Hf Sm Ti Y Yb Sr K Rb Ba Th Ta Nb Ce P Zr Hf Sm Ti Y Yb
10.4 12
G 10
H Andesites Upper
Abejorral member
DM Ophiolitic gabbros
8
Cauca Ophiolitic gabbros
6 Ophiolitic Ophiolitic basalts
Complex
Ophiolitic basalts
9.6 4
ε Nd

Jurassic arc
ε Nd

2
0
CHUR

-2
9.0
-4
-6

-8
8.4
0 100
0.2 0.6
(La/Sm) n
1.0 1.4 Age (Ma) 200 300

Fig. 11. (A), Nb/Y vs. Zr/Ti rock classification diagram (Pearce, 1996); (B), Ti versus V tectonic discrimination diagram after Shervais (1982). (C), Rare Earth Element (REE) patterns
normalized to Chondrite (Nakamura, 1974) from the analyzed volcanic rocks; (D), Rare Earth Element (REE) patterns normalized to Chondrite (Nakamura, 1974) from the analyzed
plutonic rocks; (E), NMORB normalized multi-elemental patterns of the analyzed volcanic rocks (Pearce et al., 1984); (F), MORB normalized multi-elemental patterns of the analyzed
volcanic rock (Pearce et al., 1984); (G), εNd vs. L/Sm as a monitor of crustal input; (H), εNd values from the analyzed Cretaceous volcanic and plutonic rocks (This study) compare
with εNd values from the Jurassic continental arc (Cochrane et al., 2014a; Bustamante et al., 2017; Quandt et al., 2018).

composition and higher TiO2 contents (0.42–0.47% versus 0.66–3.22%) vs. Y diagrams, the rocks are tholeitic basalts with a single sample
(Fig. 11A). The eight volcanic rocks presented SiO2 values between within the field of andesites (Fig. 11A).
43.87% and 52.57% with total alkalis values (Na2O + K2O) of 1.91% to REE elements are characterized by negative to horizontal trends in
5.35%, which are both characteristic of basalts. Fe2O3 values ranged be- the LREE (Fig. 11C), with (La/Sm)n ratios between 0.37 and 1.07 and a
tween 5.29% to 15.84% and Mg# between 25.5 and 69.6, suggesting that flat trend in the HREE with (Gd/Yb)n between 0.97 and 1.26. Samples
these rocks experienced differentiation. In the Nb/Y vs. Zr/Ti and the Zr also show a weak Eu anomaly (Eu/Eu*: 0.86 to 1.05), with one sample
S. Zapata et al. / Gondwana Research 66 (2019) 207–226 221

being characterized by a strong Eu anomaly. The multi-elemental dia- Cretaceous tectono-stratigraphic domains in the study area, which can
gram normalized to the primitive mantle showed that the lack of Ti or be used to reconstruct discrete tectonic environments that together
Nb anomalies with HFSE contents was similar or less abundant than provide major insights on the geological evolution of the Northern
N-MORB (Fig. 11E). Andes. Zircon U-Pb detrital ages recovered from the Lower Abejorral
The 13 gabbroic samples are characterized by SiO2 values between Member suggest that this unit is younger than ~150 Ma. Moreover,
48.17% and 50.50% and total alkalis (Na2O + K2O) between 2.23% and the volcanic rocks and the detrital zircons of the Upper Abejorral Mem-
4.87%, which classifies them as gabbros and gabbro-diorites ber and the Western Abejorral succession were deposited age between
(Middlemost, 1994). MgO and Fe2O3 varied between 7.03% and 10.09% 115 and 104 Ma (Fig. 2). The composition of the siliciclastic rocks and
and 5.51% to 11.83%, which is also characteristic of a tholeitic trend. their association with Early Cretaceous volcanic rocks suggest that the
Basic and tholeitic patterns are seen in the Nb/Y vs. Zr/Ti (Fig. 11A). West and the Upper Abejorral units were part of the same or similar
The REE patterns of the gabbroic rocks are characterized by deple- sedimentary basins.
tion in the LREE (Fig. 11D), (La/Sm)n between 0.37 and 0.94, flat trends Previous Aptian-Albian temporal constraints obtained for the
in the HREE with (Gd/Yb)n between 0.97 and 1.19, and a negative to Quebradagrande Complex have been established based on the presence
strongly positive Eu anomaly (Eu/Eu*: 0.92 to 2.9), which can be attrib- of Albian-Aptian fossils and zircon U-Pb ages between 114 and 113 Ma
uted to plagioclase fractionation. A slightly negative Nb and a flat trend (Arévalo et al., 2001; Gerardo and Arango, 1975; Gomez et al., 2002;
in the Ti anomaly characterized most of the samples in the multi- Maya and Gonzalez, 1995; Villagómez et al., 2011). U-Pb samples with
element patterns. These trends are characteristic of convergent margins Aptian-Albian ages seem to be more spatially related to the clastic
with limited sediment and fluid inputs (Fig. 11F). rocks of the Abejorral Fm. The Lower Cretaceous volcano-sedimentary
Both the gabbroic and basaltic rocks of the Cauca Ophiolite Complex record described as part of the Quebradagrande Complex in previous
are characterized by a tholeitic character, a N-MORB REE trend, and publications can be equivalent to the Upper Abejorral Member de-
weak Nb and Ti anomalies suggesting that these rocks may have formed scribed in this contribution.
in a distal position in relation to the slab (Dilek et al., 2007; Pearce, Unfortunately, neither our efforts or previous ones have successfully
1996; Taylor and Martinez, 2003). dated the Cauca Ophiolitic Complex. However, this ophiolitic complex is
REE and trace element geochemical characteristics from the volcanic characterized by deformation and metamorphism restricted to local
rocks of the Cauca Ophiolitic Complex bear strong similarities with the high strain zones and below the greenschist facies. In contrast, Triassic
results of other volcanic rocks associated analyzed by Rodríguez et al. mafic units are metamorphosed and associated with to s-type mica-
(2016), farther south, which are characterized by MORB to E-MORB sig- ceous granitoids which are also characterized by more negative εNd sig-
natures; and with less evolved Cretaceous samples reported by natures (Cochrane et al., 2014b; Spikings et al., 2015). A Jurassic
Jaramillo et al. (2017) to the north (Fig. 11B, C, and E). volcano-sedimentary protolith is also unlikely since these Jurassic vol-
canic rocks are also metamorphosed in the amphibolite facies (Blanco-
8.3. Whole rock Sm-Nd isotopes Quintero et al., 2014). More geochronological data is necessary; but nev-
ertheless, we favor a Cretaceous origin for Cauca Ophiolitic Complex
The Sm-Nd isotope composition for both the volcanic and plutonic based on the close relation with the Cretaceous units discussed in this
rocks from the Upper Abejorral Member and the Cauca Ophiolitic Com- contribution and the aforementioned arguments.
plex are presented in Table 3. Initial Nd isotope values were calculated Late Cretaceous volcanic rocks with crystallization ages of 83–85 Ma
from a U-Pb crystallization age of 112 Ma for the Abejorral Formation are unconformably deposited on top of the Early Cretaceous West
(sample SZ-018). For the samples from the Cauca Ophiolitic Complex, Abejorral Fm. This new Late Cretaceous volcanic crystallization ages in
we assumed the same 100–112 Ma age of the Abejorral Fm. (see rocks mapped as the Quebradagrande Complex suggest a younger mag-
Discussion section). matic record for this unit compare to the age proposed by previous au-
The sample of the Upper Abejorral Member is characterized by an thors (Maya and Gonzalez, 1995; Nivia et al., 2006; Spikings et al., 2015;
initial εNd(t) of +8.6. The volcanic rocks of the Cauca Ophiolitic Com- Villagómez and Spikings, 2013) (Fig. 10). These findings are consistent
plex also presented εNd(t) values of +10.0, which are similar to the with Upper Cretaceous fossil occurrences (Pardo-Trujillo et al., 2002)
three gabbroic rocks from this unit, which yielded initial εNd (Fig. 3) and with U-Pb zircon ages from plutonic bodies north and
(t) between +9.4 and +10.32 (Fig. 11H). These values plotted close to south of the study area (Jaramillo et al., 2017; Villagómez et al., 2011)
the depleted mantle field, which suggests that older crustal input is (Fig. 3B).
not significant.
The plot of the La/Sm ratios vs. εNd values from the Cretaceous sam- 9.2. Sedimentary provenance
ples can be used to monitor differentiation or increase of the sedimen-
tary input. The values from the Cauca Ophiolitic complex do not fit in The Lower and Upper members of the Abejorral Fm. are character-
any trend. Although, when the value of the Upper Abejorral Member ized by the presence of micaceous schist, gneissic and quartz clasts in
is considered, a trend can be identified. Therefore, it's possible to suggest the conglomerates and the presence of foliated polycrystalline quartz
that magmatic processes associated Upper Abejorral Fm. may record the and undulatory monocrystalline quartz in the sandstones. These prove-
presence of some crustal input. nance modes suggest a metamorphic terrain as a primary source for
We also included Sm-Nd isotope results from the Jurassic continen- both members. The relations between the type of monocrystalline
tal arc rocks emplaced in pre-Jurassic continental crust (Bustamante quartz and the number of grains of polycrystalline quartz after Basu
et al., 2016; Cochrane et al., 2014a; Quandt et al., 2018; Vinasco et al., et al. (1975) in the Lower Abejorral Member suggests that low-grade
2006). The Cretaceous units exhibit a stronger mantle component com- rocks were a significant portion of this metamorphic source. Detrital zir-
pare to the Jurassic samples (Fig. 11H). con age populations in lower member include Triassic (231–246 Ma),
Permian (274 Ma), Devonian to early Paleozoic and Proterozoic ages.
9. Discussion These age populations are similar to the ages of the Permo-Triassic
and the Paleozoic basement rocks exposed in the Central Cordillera
9.1. Timing of sedimentation and magmatism in the western flank of the and west of the SPF (Cochrane et al., 2014b; Martens et al., 2012;
Central Cordillera Spikings et al., 2015; Villagómez et al., 2011; Vinasco et al., 2006)
(Fig. 3).
The integration of field, petrographic, geochemical and geochrono- Conglomerates and sandstones from the Upper Abejorral Member
logical constraints enables the discrimination of coherent Early show a decrease in quartz content, an increase of siltstone fragments
222 S. Zapata et al. / Gondwana Research 66 (2019) 207–226

and the apparition of volcanic clasts (Fig. 6B and C). Detrital zircons also older ages similar to ages of the Central Cordillera basement (Fig. 9).
have Permo-Triassic and Proterozoic populations and younger Aptian- The presence of a Late Cretaceous detrital zircon age population (ca.
Albian maximum depositional ages (124–100 Ma) compare to the 84 Ma) similar to the magmatic ages of the associated volcanic rocks
lower member. These younger zircon populations are similar to the suggest basin filling coeval or posterior to the volcanic activity (Figs. 8
age of the intercalated volcanic units suggesting a coexistence of and 9).
magmatism and sediment deposition. These results suggest that the
Permo-Triassic and Paleozoic basement exposed towards the eastern 9.3. Tectonic implications
and western limit of the Abejorral Fm. were the main metamorphic
sources of both members of the Abejorral Fm. Field observations, geochemical data, new temporal constraints,
The Upper Abejorral Member was deposited in a basin with coeval stratigraphy and sediment provenance suggest that the Abejorral Fm.
volcanism and sedimentary recycling of the coarse-grained Lower records the development of one or several Early Cretaceous backarc ba-
Abejorral Member (Fig. 12C). The Upper Cretaceous sandstones and sins along the South American margin (Figs. 12E and 13E). The climax of
conglomerates of the Quebradagrande Complex show mixed prove- this extensional event was recorded in these backarc basins during the
nance with sedimentary, volcanic and plutonic lithic fragments Albian-Aptian and preserved within the RFS and in the Central Cordil-
(Fig. 6B and C). Detrital zircon populations include Permo-Triassic and lera of Colombia.

A) 80 - 60 Ma
Z Z´ OPF Paleogene plutonic rocks
Caribbean plateau collision
Quebradagrande Complex
SPF
SJF Cauca Ophiolithic Complex
CAF
CC Late Cretaceous plutonic rocks
WC Upper Abejorral M.
Lower Abejorral M.
Arquia Complex
Oceanic Caribbean plateau
B) 100 - 80 Ma Jurassic igneous rocks
Backarc closure OPF
SJF Continental basement
SPF
Late Cretaceous arc

~40 km
~100 km

C) 110 - 100 Ma
Advanced backarc stage
OPF
SJF

? ?
?
D) 130 - 110 Ma
Backarc opening OPF
SJF

E) 180 - 130 Ma
Jurassic - Early Cretaceous volcanic arc and backarc

Fig. 12. Proposed tectonic evolution of northwestern South America in the Cretaceous (180–65 Ma). SJF: San Jerónimo Fault; SPF: Silvia Pijao Fault; CAF: Cauca-Almaguer Fault; WC: West-
ern Cordillera; and CC: Central Cordillera. Black arrows denote basement erosion and exhumation.
S. Zapata et al. / Gondwana Research 66 (2019) 207–226 223

80°0'0"W 70°0'0"W 80°0'0"W 70°0'0"W 80°0'0"W 70°0'0"W

A) 80 - 60 Ma B) 100 - 80 Ma Proto-Caribbean C) 110 - 100 Ma

arc
Volcanic

10°0'0"N

10°0'0"N
10°0'0"N

SJF
Z Z´
OPF ?

SPF

F
Z

OPF

SJ
SJF

0°0'0"

0°0'0"
0°0'0" Z Z´
CAF

OPF
arc
Volcanic
Backarc
sea floor
spreading

80°0'0"W 70°0'0"W
80°0'0"W 70°0'0"W

D) 130 - 110 Ma E) 180 - 130 Ma

Quebradagrande Complex
10°0'0"N

10°0'0"N
Upper Abejorral Member

Lower Abejorral Member


t if
ra R
dille

Late Cretaceous volcanic arc


Cor
tern
Eas

Oceanic Caribbean plateau

Eastern Cordillera basin


Backarc basins
0°0'0"

Z Z
0°0'0"

Arquía Complex

Z´ Z´
Pre-Cretaceous basement
arc
Volcanic

Fig. 13. Conceptual paleogeographic model illustrating the evolution of northwestern South America during the Cretaceous (140 Ma–65 Ma), modified from the available regional models
(Burke et al., 1984; Jaramillo et al., 2017; Kennan and Pindell, 2009; Pindell et al., 2005; Spikings et al., 2015; Toussaint and Restrepo, 1996; Villamil and Pindell, 1999; Vinasco and Cordani,
2012).

The arrangement of the studied units resembles the geometry of an deltaic to a marine depositional environment between 125 and
inverted backarc basin. This geometry is defined by continental Permo- 100 Ma (Gomez et al., 2002; González, 1980; Quiroz, 2005). This transi-
Triassic blocks in the east and west boundaries of the Cretaceous tion is marked by an upward-fining and thinning pattern, the appear-
volcano-sedimentary strata (i.e. Pacora stock, Cambumbia stock, ance of interlayered chert, and the contrast in the fossil record
Abejorral gneiss and Pantanillo gneiss) (Cochrane et al., 2014a; between the Lower and the Upper Abejorral members (Fig. 5). The
Vinasco et al., 2006); several of these blocks were described and/or provenance of the Upper Abejorral member is characterized by conti-
dated in this contribution (Fig. 3). On both sides of the basin, sedimen- nental basement sources, reworked sedimentary sources and by the ap-
tary rocks from the Abejorral Fm. were deposited on top of the Permo- pearance volcanism coeval with the basin filling.
Triassic basement. In the middle of these continental basement blocks The limited enrichment of LREE recorded and the positive εNd
and the siliciclastic strata are the Ophiolitic remnants of the Cauca values of the volcanic rocks of the Upper Abejorral member (Fig. 11H)
Ophiolitic Complex; these suprasubduction ophiolitic rocks may be suggest that magmatic interactions with the continental crust were lim-
the result of seafloor spreading during advanced stages of the backarc ited, these observations are compatible with the existence of a thinned
extension (Figs. 12C and 13C). continental crust and a structural controlled magmatic emplacement.
We compiled several stratigraphic columns from the Abejorral Fm. This extended crust provided the space for fast emplacement of
in other localities and correlated them with the observations in the magmas at shallow depths, the volcanic rocks in the Upper Abejorral
study area (Fig. 3). These sections record a transition from fluvial- Formation were emplaced in an extended continental crust. The
224 S. Zapata et al. / Gondwana Research 66 (2019) 207–226

MORB to supra-subduction affinity of the Cauca Ophiolitic Complex may Kerr, 2010; Kerr et al., 1997; Pindell and Kenan, 2009; Pindell et al.,
represent more advanced stages of backarc basin extension character- 2005; Spikings et al., 2015; Vallejo et al., 2006; Van Der Lelij et al.,
ized by the formation of oceanic crust (Figs. 12C and 13C). 2010; Villagómez et al., 2011) (Figs. 12A and 13A).
The field relations, the provenance, the transition to marine condi- Previous models have interpreted the Cretaceous evolution of the
tions, the apparition of syn-sedimentary volcanism and the geochemical Colombian Andes as the result of local compressional tectonics due to
character of the ophiolitic remnants can be interpreted as the result of the accretion of exotic oceanic crust (Cediel et al., 2003; Van Der Lelij
the progressive development of a backarc basin. The development of et al., 2010; Pindell and Erikson, 1993; Spikings et al., 2015). In contrast,
this basin was accompanied by the unroofing of continental horst blocks in our model the transition from extension to compression took place
on both sides of the basin (Figs. 12 and 13). prior to the accretion of the exotic oceanic terranes between the Late
Continental extension and associated magmatism may have contin- Cretaceous and the Paleocene. Additionally, the regional extension of
ued until the Albian (100 Ma) as suggested by the youngest magmatic these extensional and compressional tectonic settings suggests that
ages. This tectonic settings differs from previous models, which have the Cretaceous tectonic evolution of the Northern Andes was controlled
considered that during this time the backarc basin was closed due to by plate scale Andean tectonics.
the accretion of an oceanic terrane or to major changes in the subduc-
tion zone (Cediel and Shaw, 2003; Cochrane et al., 2014b; Spikings 10. Conclusions
et al., 2015; Villagómez and Spikings, 2013) (Fig. 2). These previous
models were based on the lack magmatism after 114 Ma and on the co- Integrated field, stratigraphic, provenance, geochemical, and geo-
eval cooling/exhumation events in the Central Cordillera basement and chronological constraints of the Cretaceous sedimentary, volcanic and
in the Arquía Complex. According to these models, the Abejorral Fm. plutonic rocks from the western flank of the Central Cordillera of
was deposited in a foreland basin. The data presented here show that Colombia documents an extensional phase characterized by a transgres-
magmatism was active until 100 Ma; our data also suggest that basin sive basin filling, sedimentary recycling, and volcanism during the Early
filling is more compatible with an extensional tectonic setting Cretaceous (150–100 Ma). The transition from extensional to compres-
(Figs. 12B and 13B). Therefore, published cooling ages may also be sional tectonic settings closed this backarc basin during the Late Creta-
interpreted as extensional-controlled horst exhumation. ceous (100–90 Ma). The volcanic rocks of the Quebradagrande
Regional tectonic models for the western margin of South America Complex and the Antioquia Batholith were formed in a Late Cretaceous
typically show an oblique convergence between South America and volcanic arc built on top of the deformed backarc basin. A final accre-
the Pacific plate during the Early Cretaceous (Burke et al., 1984; tionary event occurred between the Late Cretaceous and the Paleocene
Kennan and Pindell, 2009; Pindell et al., 2005; Villamil and Pindell, due to the accretion of the Caribbean oceanic plateau. Our model docu-
1999). Paleomagnetic data of Jurassic rocks and provenance analysis ments the onset of Cenozoic compressional tectonics in the Northern
on Oligocene strata indicate at least 500 km of northward displacement Andes prior to the accretion of the Caribbean plateau.
of the Central Cordillera; this relative displacement took place during
the Meso-Cenozoic from a southern position (Bayona et al., 2010,
2006; Lamus Ochoa et al., 2013). The documented extensional phase Acknowledgments
was coeval with the northern translation of the studied tectonic blocks.
Therefore, it is feasible that these backarc extensional and compres- We acknowledge COLCIENCIAS for the financial support to the
sional events had a strike-slip component responsible for the latitudinal Corporación Geológica ARES within the program of “Fortalecimiento
translation of these blocks and the subsequent reincorporation to the Institucional” that granted Sebastian Zapata Henao with a scholarship.
margin in a different paleogeographic position (Figs. 12B and 13B). Funding was also received from the National University of Colombia
Between 86 and 83 Ma, andesitic lava flows and coarse-grained sed- projects 25452, 25340, 29182, 18593, 24208, and the Fundación para
iments of the Quebradagrande Complex were deposited in an angular la Promoción de la Investigación y la Tecnología del Banco de la
unconformity over volcanic and sedimentary rocks of the Abejorral República de Colombia, project 3451. Germán Bayona, Gaspar
Fm. This Late Cretaceous magmatic arc also includes intrusive facies Monsalve, students from the EGEO research group and those from dif-
with ages between 90 and 80 Ma exposed north and south of the ferent courses of field geology in the National University of Colombia
study area (Jaramillo et al., 2017; Villagómez et al., 2011) (Fig. 3A). are acknowledged for their discussions and help during fieldwork. We
This younger Quebradagrande Complex records the erosion of the acknowledge Paul Mann and the other anonymous reviewers for
older siliciclastic rocks and post-date the closure of the former backarc reviewing this contribution and for all their helpful comments.
basin (Figs. 12B and 13B).
We proposed a transition to a compressive tectonic setting between Appendix A. Supplementary data
100 and 86 Ma. This compressive tectonic scenario was responsible for
the closure of the former backarc basin and the development of a new Supplementary data to this article can be found online at https://doi.
Late Cretaceous volcanic arc. We also suggest that the large plutonic org/10.1016/j.gr.2018.10.008.
province that formed the Antioquia Batholith in the Central Cordillera
and the Quebradagrande Complex were part of this Late Cretaceous vol- References
canic arc (Figs. 12B and 13B). This transition between extensional and
Alvarez, J., 1983. Geología de la Cordillera Central y el Occidente colombiano y
compressional tectonic settings resulted in the onset of the compression petroquímica de los intrusivos granitoides Mesocenozóicos.
in the Northern Andes. Similar Early Cretaceous extensional tectonics Arévalo, O.J., Mojica, J., Patarroyo, P., 2001. Sedimentitas del Aptiano tardío al sur de Pijao,
and Late Cretaceous compressional tectonics have been extensively Quebrada La Maizena, Flanco occidental de la Cordillera Central, Departamento del
Quindío, Colombia. Geología Colombiana 26, 29–43.
documented along the west South-American margin (Baby et al.,
Aspden, J.A., McCourt, W.J., Brook, M., 1987. Geometrical control of subduction-related
2013; e.g. Jaillard and Soler, 1996; Jaimes and de Freitas, 2006; Martini magmatism: the Mesozoic and Cenozoic plutonic history of Western Colombia. Jour-
et al., 2013, 2011; Mora et al., 2009; Pindell and Erikson, 1993; Salfity nal of the Geological Society of London 144, 893–905.
Baby, P., Rivadeneira, M., Barragán, R., Christophoul, F., 2013. Thick-skinned tectonics in
and Marquillas, 1994; Sarmiento-Rojas et al., 2006; Tunik et al., 2010;
the Oriente foreland basin of Ecuador. Geological Society of London, Special Publica-
Viramonte et al., 1999). tion 377, 59–76.
After the deposition of the Quebradagrande Complex the northern Barrero, D., 1979. Geology of the Central Western Cordillera, West of Buga and Roldanillo,
continental margin of South America collided with an intra-oceanic pla- Colombia. vol. 4. Publicaciones Espec. del Ingeominas, pp. 1–75.
Basu, A., Young, S.W., Suttner, L.J., James, W.C., Mack, G.H., 1975. Re-evaluation of the use
teau derived from the southeast Pacific as part of the Caribbean plate be- of undulatory extinction and polycrystallinity in detrital quartz for provenance inter-
tween the Late Cretaceous and the Paleocene (80–60 Ma) (Hastie and pretation. Journal of Sedimentary Research 45.
S. Zapata et al. / Gondwana Research 66 (2019) 207–226 225

Bayona, G., Rapalini, A., Alvarez, V., 2006. Paleomagnetism in Mesozoic rocks of the North- Gomez, E., Jordan, T., Allmendinger, R.W., 2003. Syntectonic Cenozoic sedimentation in
ern Andes and its implications in Mesozoic Tectonics of Northwestern South America. the northern middle Magdalena Valley Basin of Colombia and implications for exhu-
Earth, Planets and Space 2008, 1–18. mation of the Northern Andes. Geological Society of America Bulletin 117, 547–569.
Bayona, G., Jimenez, G., Silva, C.A., Montes, C., Roncancio, J., Cordani, U., 2010. Paleomag- Gómez, J., Nivia, Á., Montes, N.E., Almanza, M.F., Alcárcel, F.A., Madrid, C., 2015. Notas
netic data and K–Ar ages from Mesozoic units of the SantaMarta massif: a prelimi- explicativas: Mapa Geológico de Colombia. Compil. la Geol. Colomb. Una visión a,
nary interpretation for block rotation and translations. Journal of South American pp. 9–33.
Earth Sciences 29, 817–831. Gomez-Cruz, A.D.J., Moreno-Sánchez, M., Pardo-Trujillo, A., 1995. Edad y Origen del
Bayona, G., Montes, C., Cardona, A., Jaramillo, C., Ojeda, G., Valencia, V., Ayala-Calvo, C., Complejo metasedimentario Aranzazu-Manizales en los Alrededores de Manizales
2011. Intraplate subsidence and basin filling adjacent to an oceanic arc-continent col- (Departamento de Caldas, Colombia). Geología Colombiana 19, 83–93.
lision: a case from the southern Caribbean-South America plate margin. Basin Re- González, H., 1980. Geologia de la planchas 167 (Sonson) y 187 (Salamina)(escala 1:
search 10, 1365–2117. 100,000). Instituto Nacional de Investigaciones Geológico-Mineras.
Borrero, C., Pardo - Trujillo, A., Jaramillo, C., Osorio, J.A., Cardona, A., Flores, J.A., Echeverry, González, H., 2001. Mapa geológico de Antioquia Escala 1: 400.000. Memoria explicativa.
S., 2012. Early Paleogene magmatism in the northern Andes: insights on the effects of Ingeominas, Bogotá.
oceanic plateau-continent. 39, 75–92. González, H., Nuñez, A., Paris, G., 1988. Mapa geologico de Colombia, 1: 1 500 000 1988:
Bouvier, A., Vervoort, J.D., Patchett, P.J., 2008. The Lu–Hf and Sm–Nd isotopic compo- memoria explicativa. Instituto Nacional de Investigacionnes Geologico-Mineras.
sition of CHUR: constraints from unequilibrated chondrites and implications for Hastie, A.R., Kerr, A.C., 2010. Mantle plume or slab window? Physical and geochemical
the bulk composition of terrestrial planets. Earth and Planetary Science Letters constraints on the origin of the Caribbean oceanic plateau. Earth-Science Reviews
273 (1), 48–57. 98, 283–293.
Burke, K., Cooper, C., Dewey, J.F., Mann, P., Pindell, I.L., 1984. Caribbean tectonics and rel- Hincapié-Gómez, S., Cardona, A., Jiménez, G., Monsalve, G., Ramírez-Hoyos, L., Bayona, G.,
ative plate motions. Geological Society of America Bulletin 162, 31–59. 2018. Paleomagnetic and gravimetrical reconnaissance of Cretaceous volcanic rocks
Bustamante, A., Juliani, C., Essene, E.J., Hall, C.M., Hyppolito, T., 2011a. Geochemical con- from the Western Colombian Andes: paleogeographic connections with the Carib-
straints on blueschist- and amphibolite-facies rocks of the Central Cordillera of bean Plate. Studia Geophysica et Geodaetica 62, 485–511.
Colombia: the Andean Barragán region. International Geology Review 54, Howard, J.L., 1993. The statistics of counting clasts in rudites: a review, with examples
1013–1030. https://doi.org/10.1080/00206814.2011.594226. from the upper Palaeogene of southern California. Sedimentology 40, 157–174.
Bustamante, A., Juliani, C., Hall, C.M., 2011b. 40Ar/39Ar ages from blueschists of the Ibañez-Mejia, M., Tassinari, C.C.G., Jaramillo-Mejia, J.M., 2007. U–Pb Zircon Ages of the
Jambaló region, Central Cordillera of Colombia: implications on the styles of accretion “Antioquian Batholith”: Geochronological Constraints of Late Cretaceous Magmatism
in the Northern Andes. Geologica Acta 9, 351–362. in the Central Andes of Colombia.
Bustamante, C., Archanjo, C.J., Cardona, A., Vervoort, J.D., 2016. Late Jurassic to Early Cre- Jaillard, E., Soler, P., 1996. Cretaceous to early Paleogene tectonic evolution of the northern
taceous plutonism in the Colombian Andes: a record of long-term arc maturity. Geo- Central Andes (0–18 S) and its relations to geodynamics. Tectonophysics 259, 41–53.
logical Society of America Bulletin 128, 1762–1779. Jaimes, E., de Freitas, M., 2006. An Albian–Cenomanian unconformity in the northern
Bustamante, C., Archanjo, C.J., Cardona, A., Bustamante, A., Valencia, V.A., 2017. U-Pb Ages Andes: evidence and tectonic significance. Journal of South American Earth Sciences
and Hf Isotopes in Zircons from Parautochthonous Mesozoic Terranes in the Western 21, 466–492. https://doi.org/10.1016/j.jsames.2006.07.011.
Margin of Pangea: Implications for the Terrane Configurations in the Northern Andes. Jaramillo, J.S., Cardona, A., León, S., Valencia, V., Vinasco, C., 2017. Geochemistry and geo-
The Journal of Geology 125 (5), 487–500. chronology from Cretaceous magmatic and sedimentary rocks at 6°35' N, western
Cawood, P.A., Kröner, A., Collins, W.J., Kusky, T.M., Mooney, W.D., Windley, B.F., Kroner, A., flank of the Central cordillera (Colombian Andes): Magmatic record of arc growth
Collins, W.J., Kusky, T.M., Mooney, W.D., Windley, B.F., 2009. Accretionary orogens and collision. Journal of South American Earth Sciences 76, 460–481.
through Earth history. Geological Society of London, Special Publication 318, 1–36. Kennan, L., Pindell, J.L., 2009. Dextral shear, terrane accretion and basin formation in the
https://doi.org/10.1144/SP318.1. Northern Andes: best explained by interaction with a Pacific-derived Caribbean
Cediel, F., Shaw, R.P., 2003. Tectonic assembly of the Northern Andean Block. AAPG Mem- Plate? Geological Society of London, Special Publication 328, 487–531. https://doi.
oir 79, 815–848. org/10.1144/SP328.20.
Cediel, F., Shaw, R., Caceres, C., 2003. Tectonic assembly of the Northern Andean Kerr, A.C., Tarney, J., Marriner, G.F., Nivia, A., Klaver, G., Saunders, A.D., 1996. The geo-
Block. The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habitats, chemistry and tectonic setting of late Cretaceous Caribbean and Colombian. Journal
Basin Formation and Plate Tectonics. Am. Assoc. Pet. Geol. Bull. vol. 79, of South American Earth Sciences 9, 111–120.
pp. 815–848. Kerr, A.C., Marriner, G.F., Tarney, J., Nivia, A., Saunders, A.D., Thirlwall, M.F., Sinton, C.W.,
Chang, Z., Vervoort, J.D., McClelland, W.C., Knaack, C., 2006. U-Pb dating of zircon by LA- 1997. Cretaceous basaltic terranes in Western Colombia: elemental, chronological
ICP-MS. Geochemistry, Geophysics, Geosystems 7, 1–14. and Sr–Nd Isotopic Constraints on petrogenesis. Journal of Petrology 38, 677–702.
Cochrane, R., Spikings, R., Gerdes, A., Ulianov, A., Mora, A., Villagómez, D., Putlitz, B., Lamus Ochoa, F., Bayona, G., Cardona, A., Mora, A., 2013. Provenance of Cenozoic units of
Chiaradia, M., 2014a. Permo-Triassic anatexis, continental rifting and the disassembly the Guaduas syncline: implication in the tectonic evolution of the south of middle
of western Pangaea. Lithos 190–191, 383–402. https://doi.org/10.1016/j. Magdalena valley and adjacent orogens. Boletin de Geología 35, 17–42.
lithos.2013.12.020. Leal-Mejia, H., 2011. Phanerozoic Gold Metallogeny in the Colombian Andes: A Tectono-
Cochrane, R., Spikings, R., Gerdes, A., Winkler, W., Ulianov, A., Mora, A., Chiaradia, M., magmatic Approach.
2014b. Distinguishing between in-situ and accretionary growth of continents along Ludwig, K.C., 2007. User's Manual For Isoplot 3.7. Berkley Geochronology Center, Berkley.
active margins. Lithos 202–203, 382–394. https://doi.org/10.1016/j. Martens, U.C., Restrepo, J.J., Solari, L.A., 2012. Sinifaná metasedimentites and relations
lithos.2014.05.031. with Cajamarca paragneisses of the central cordillera of Colombia. Boletín Ciencias
Correa, A.M., Pimentel, M., Restrepo, J.J., Nilson, A., Ordoñez, O., Martens, U., Laux, J.E., la Tierra 99–110.
Junges, S., 2006. U-Pb Zircon Ages and Nd-Sr Isotopes of the Altavista Stock and the Martini, M., Mori, L., Solari, L., Centeno-García, E., 2011. Sandstone Provenance of the
San Diego Gabbro: New Insights on Cretaceous Arc Magmatism in the Colombian Arperos Basin (Sierra de Guanajuato, Central Mexico): Late Jurassic–Early Cretaceous
Andes. pp. 84–86. Back-Arc Spreading as the Foundation of the Guerrero Terrane. Journal of Geology
Dickinson, W.R., 1985. Interpreting provenance relations from detrital modes of Sand- 119, 597–617. https://doi.org/10.1086/661989.
stones. Proven. Arenites, pp. 333–361. Martini, M., Solari, L., Camprubí, A., 2013. Kinematics of the Guerrero terrane accretion in
Dickinson, W.R., Gehrels, G.E., 2009. Use of U–Pb ages of detrital zircons to infer maxi- the Sierra de Guanajuato, central Mexico: new insights for the structural evolution of
mum depositional ages of strata: a test against a Colorado Plateau Mesozoic database. arc–continent collisional zones. International Geology Review 55, 574–589.
Earth and Planetary Science Letters 288, 115–125. Maya, M., Gonzalez, H., 1995. Unidades Litodemicas De la Cordillera Central. Informe
Dilek, Y., Furnes, H., Shallo, M., 2007. Suprasubduction zone ophiolite formation along the Unidad Operativa Medellin. Ingeominas, pp. 44–57.
periphery of Mesozoic Gondwana. Gondwana Research 11, 453–475. Middlemost, E.A.K., 1994. Naming materials in the magma/igneous rock system. Earth-
Erlich, R.N., Villamil, T., Keens-Dumas, J., 2003. Controls on the Deposition of Upper Creta- Science Reviews 37, 215–224.
ceous Organic Carbon–Rich Rocks From Costa Rica to Suriname. Mora, A., Gaona, T., Kley, J., Montoya, D., Parra, M., Quiroz, L.I., Reyes, G., Strecker, M.R.,
Escalona, A., Mann, P., 2011. Tectonics, basin subsidence mechanism and paleogeography 2009. The role of inherited extensional fault segmentation and linkage in contrac-
of the Caribbean-South American plate boundary zone. Marine and Petroleum Geol- tional orogenesis: a reconstruction of Lower Cretaceous inverted rift basins in the
ogy 28, 28. Eastern Cordillera of Colombia. Basin Research 21, 111–137. https://doi.org/
Folk, R.L., 1980. Petrology of Sedimentary Rocks. Hemphill, Austin, Texas. 10.1111/j.1365-2117.2008.00367.x.
Garzanti, E., Andò, S., Vezzoli, G., Megid, A.A.A., El Kammar, A., 2006. Petrology of Nile Nakamura, N., 1974. Determination of REE, Ba, Fe, Mg, Na and K in carbonaceous and or-
River sands (Ethiopia and Sudan): sediment budgets and erosion patterns. Earth dinary chondrites. Geochimica et Cosmochimica Acta 38, 757–775.
and Planetary Science Letters 252, 327–341. Nivia, A., Marriner, G.F., Kerr, A.C., Tarney, J., 2006. The Quebrada Grande Complex: a
Gaschnig, R.M., Vervoort, J.D., Lewis, R.S., Tikoff, B., 2011. Isotopic evolution of the Idaho Lower Cretaceous ensialic marginal basin in the Central Cordillera of the Colombian
batholith and Challis intrusive province, northern US Cordillera. Journal of Petrology Andes. Journal of South American Earth Sciences 21, 423–436. https://doi.org/
52, 2397–2429. 10.1016/j.jsames.2006.07.002.
Gehrels, G., Valencia, V., Pullen, A., 2006. Detrital zircon geochronology by laser-ablation Pardo-Trujillo, A., Moreno-Sánchez, M., Gomez-Cruz, A., 2002. Stratigraphy of some
multicollector ICPMS at the Arizona LaserChron Center. Paleontol. Soc. Pap. 12, Upper Cretaceous deposits of the Central and Western Cordilleras of Colombia: re-
pp. 67–76. gional implications. International Journal of Tropical Geology, Geography and Ecology
Gerardo, I., Arango, B., 1975. No Title XXVII. 1–113.
Gomez, A., Moreno, A., Pardo, A., 1995. Edad y origen de Complejo metasedimentario de Pardo-Trujillo, A., Cardona, A., Silva, J.C., Borrero, C., Tamayo, J.E., 2011.
Aranzazu-Manizales en los alrededores de Manizales (Departamento de Caldas, Geocronología U/Pb en circones detríticos del Complejo Quebradagrande:
Colombia). Geología Colombiana 19, 83–93. nuevos datos sobre la procedencia de los sedimentos cretáceos en la margen
Gomez, A.J., Moreno, M., Pardo, A., 2002. Afloramientos Fosiliferos del Cretacico Superior NW de Suramérica. XIV Congreso Latinoamericano de Geología y XIII Congreso
en el municipio de Pijao (Borde occidental de la cordillera central). Colombiano de Geología.
226 S. Zapata et al. / Gondwana Research 66 (2019) 207–226

Pearce, J.A., 1996. A user's guide to basalt discrimination diagrams. Trace Elem. Geochem- northwestern South America: from Pangaea to the early collision of the Caribbean
istry Volcan. Rocks Appl. Massive Sulphide Explor.. Geol. Assoc. Canada, Short Course Large Igneous Province (290–75 Ma). Gondwana Research 27, 95–139. https://doi.
Notes 12, p. 113 org/10.1016/j.gr.2014.06.004.
Pearce, J.A., 2008. Geochemical fingerprinting of oceanic basalts with applications to Taylor, B., Martinez, F., 2003. Back-arc basin basalt systematics. Earth and Planetary Sci-
ophiolite classification and the search for Archean oceanic crust. Lithos 100, 14–48. ence Letters 210, 481–497.
Pearce, J.A., Harris, N.W., Tindle, A.G., 1984. Trace element discrimination diagrams for the Toussaint, J.F., Restrepo, J.J., 1996. Evolucion Geologica de Colomnbia–Cretacico.
tectonic interpretation of granitic rocks. Journal of Petrology 25, 956–983. Universidad Nacional De Colombia, Medellin.
Pindell, J., Erikson, J., 1993. The Mesozoic margin of northern South America. In: Salfity, J. Tunik, M., Folguera, A., Naipauer, M., Pimentel, M., Ramos, V.A., 2010. Early uplift and oro-
(Ed.), Cretaceous Tectonics of the Andes. Vieweg Ger., pp. 1–60. genic deformation in the Neuquén Basin: constraints on the Andean uplift from U–Pb
Pindell, I.L., Kenan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and and Hf isotopic data of detrital zircons. Tectonophysics 489, 258–273.
northern South America. Geological Society [London] Special Publication 328, 1–56. Uyeda, S., Kanamori, H., 1979. Back-arc opening and the mode of subduction. Journal of
Pindell, I.L., Cande, S.C., Pitman III, W.C., Rowley, D.B., Dewey, J.F., LaBreque, J., Haxby, W., Geophysical Research 84, 1049. https://doi.org/10.1029/JB084iB03p01049.
1998. A platekinematic framework for models of Caribbean evolution. Valencia, V., Ruiz, J., Barra, F., Gehrels, G., Ducea, M., Titley, S.R., Ochoa, L., 2005. U–Pb zir-
Tectonophysics 155. con and Re–Os molybdenite geochronology from La Caridad porphyry copper de-
Pindell, I.L., Kenan, L., Maresch, W.V., Stasnek, K.P., Draper, G., Higgs, R., 2005. Plate kine- posit: insights for the duration of magmatism and mineralization in the Nacozari
matic and crustal dynamics of circum-Caribbean arc-continent interactions: tectonic District, Sonora, Mexico. Mineral Deposits 40, 175–191.
controls on basin development in the Proto-Caribbean margins. Geological Society of Vallejo, C., Spikings, R., Winkler, W., Luxieux, L., Chew, D., 2006. The early interaction be-
America Bulletin 394, 7–52. tween the Caribbean Plateau and the NW South American plate. Terra Nova 18,
Quandt, D., Trumbull, R.B., Altenberger, U., Cardona, A., Romer, R.L., Bayona, G., ... Guzman, 264–269.
G., 2018. The geochemistry and geochronology of Early Jurassic igneous rocks from Van Der Lelij, R., Spikings, R.A., Kerr, A.C., Kounov, A., Cosca, M., Chew, D., Villagomez, D.,
the Sierra Nevada de Santa Marta, NW Colombia, and tectono-magmatic implications. 2010. Thermochronology and Tectonics of the Leeward Antilles: evolution of the
Journal of South American Earth Sciences 86, 216–230. Southern Caribbean Plate Boundary Zone. Tectonics 29, 1–30. https://doi.org/
Quiroz, L., 2005. Cartografía y análisis estratigráfico de las formaciones Valle Alto y 10.1029/2009TC002654.
Abejorral en la región de San Félix, departamento de caldas, Colombia. National Uni- Vasquez, M., Altenberger, U., 2005. Mid-Cretaceous extension-related magmatism in the
versity of Colombia. eastern Colombian Andes. Journal of South American Earth Sciences 20, 193–210.
Ramos, V.A., 1999. Plate tectonic setting of the Andean Cordillera. Episodes 22, 183–190. Vavra, G., Schmid, R., Gebauer, D., 1999. Internal morphology, habit and U-Th-Pb micro-
Restrepo, J.J., Toussaint, J.F., 1988. ). Terranes and continental accretion in the Colombian analysis of amphibolite-to-granulite facies zircons: geochronology of the Ivrea Zone
Andes. Episodes 11 (3), 189–193. (Southern Alps). Contributions to Mineralogy and Petrology 134, 380–404.
Restrepo, J.J., Toussaint, J.F., 1991. Terranes and continental Acretion in the Colombian Vervoort, J.D., Blichert-Toft, J., 1999. Evolution of the depleted mantle: Hf isotope evidence
Andes. Episodes 11 (3), 189–193. from juvenile rocks through time. Geochimica et Cosmochimica Acta 63, 533–556.
Restrepo, J.J., Ordóñez-carmona, O., Moreno-sánchez, M., 2009. A Comment On “The Villagómez, D., Spikings, R., 2013. Thermochronology and tectonics of the Central and
Quebradagrande Complex: a Lower Cretaceous ensialic marginal basin in the Central Western Cordilleras of Colombia: Early Cretaceous–Tertiary evolution of the North-
Cordillera of the Colombian Andes” by Nivia et al. Journal of South American Earth ern Andes. Lithos 160–161, 228–249. https://doi.org/10.1016/j.lithos.2012.12.008.
Sciences 28, 204–205. https://doi.org/10.1016/j.jsames.2009.03.004. Villagómez, D., Spikings, R., Magna, T., Kammer, A., Winkler, W., Beltrán, A., 2011. Geo-
Restrepo-Moreno, S.A., Foster, D.A., Stockli, D.F., Parra-Sánchez, L.N., 2009. Long-term ero- chronology, geochemistry and tectonic evolution of the Western and Central cordil-
sion and exhumation of the “Altiplano Antioqueño”, Northern Andes (Colombia) leras of Colombia. Lithos 125, 875–896. https://doi.org/10.1016/j.lithos.2011.05.003.
from apatite (U–Th)/He thermochronology. Earth and Planetary Science Letters Villamil, T., Pindell, J., 1999. Mesozoic paleogeographic evolution of northern South
278, 1–12. https://doi.org/10.1016/j.epsl.2008.09.037. America. Society for Sedimentary Geology 58, 283–318.
Rodriguez, G., Zapata, G., 2013. Comparative analysis of the Barroso Formation and Vinasco, C., Cordani, U., 2012. Reactivation episodes of the Romeral fault system in the
Quebradagrande Complex: a volcanic arc tholeiitic–calcoalcaline, segmented by the Northwestern part of Central Andes, Colombia, through 39AR-40AR and K-AR results.
fault system Romeral in Northern Andes. Boletín de Ciencias de la Tierra 33, 39–58. Boletín Ciencias la Tierra 111–124.
Rodríguez, G., Cetina, T., María, L., 2016. Caracterízación petrografíca y química de rocas Vinasco Vallejo, C., Cordani, U., 2012. Episodios de reactivación del sistema de fallas del
de corteza oceánica del Complejo Quebradagrande y comparación con rocas de la Romeral en la parte Nor-Occidental de los Andes Centrales de Colombia a través de
unidad Diabasas de San José de Urama. Boletín de Geología 38, 15–29. resultados 39AR-40AR y K-AR. Boletín Ciencias la Tierra 111–124.
Royden, L.H., 1993. The tectonic expression slab pull at continental convergent bound- Vinasco, C., Cordani, U.G., González, H., Weber, M., Pelaez, C., 2006. Geochronological, iso-
aries. Tectonics 12, 303–325. https://doi.org/10.1029/92TC02248. topic, and geochemical data from Permo-Triassic granitic gneisses and granitoids of
Rubatto, D., 2002. Zircon trace element geochemistry: partitioning with garnet and the the Colombian Central Andes. Journal of South American Earth Sciences 21, 355–371.
link between U–Pb ages and metamorphism. Chemical Geology 184, 123–138. Viramonte, J.G., Kay, S.M., Becchio, R., Escayola, M., Novitski, I., 1999. Cretaceous rift re-
Ruiz-Jimenez, E.C., Blanco-Quintero, I.F., Toro-Toro, L.U.Z., Moreno-Sánchez, M., Vinasco, lated magmatism in central-western South America. Journal of South American
C.J., Garcia-Casco, A., Morata, D., Gomez-Cruz, A., 2012. Geochemestry and petrology Earth Sciences 12, 109–121. https://doi.org/10.1016/S0895-9811(99)00009-7.
of the metabasites of the Arquia Complex (Santa Fe de Antioquia and Arquia Creek). Weber, M., Gómez-Tapias, J., Cardona, A., Duarte, E., Pardo, A., Valencia, V., 2015. Geo-
Boletín Ciencias la Tierra 65–80. chemistry of the Santa Fe Batholith and Buriticá Tonalite in NW Colombia - evidence
Salfity, J.A., Marquillas, R.A., 1994. Tectonic and sedimentary evolution of the Cretaceous- of subduction initiation beneath the Colombian Caribbean Plateau. Journal of South
Eocene Salta Group basin, Argentina. Cretaceous Tectonics of the Andes. Springer, American Earth Sciences https://doi.org/10.1016/j.jsames.2015.04.002.
pp. 266–315. Zapata, S., 2009. Aportes a la evolución tectónica Neogena del Caribe, mediante la
Sarmiento-Rojas, L.F., Van Wess, L.D., Cloetingh, S., Sarmiento-Rojas, L.F., Van Wess, J.D., procedencia de Sedimentos terciarios al noroccidente de la serranía de Jarara,
Cloetingh, S., 2006. Mesozoic transtensional basin history of the Eastern Cordillera, Guajira/Colombia.
Colombian Andes: inferences from tectonic models. Journal of South American Zuluaga, C.A., Pinilla, A., Mann, P., 2015. Jurassic Silicic Volcanism and Associated
Earth Sciences 21, 383–411. https://doi.org/10.1016/j.jsames.2006.07.003. Continental-arc Basin in Northwestern Colombia (Southern Boundary of the Carib-
Shervais, J.W., 1982. Ti-V plots and the petrogenesis of modern and ophiolitic lavas. Earth bean Plate).
and Planetary Science Letters 59, 101–118.
Spikings, R., Cochrane, R., Villagomez, D., der Lelij, R., Vallejo, C., Winkler, W., Beate, B., Van
Der Lelij, R., Vallejo, C., Winkler, W., Beate, B., 2015. The geological history of

You might also like