You are on page 1of 28

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/303834834

Provenance of middle to late Palaeozoic


sediments in the northeastern Colombian
Andes: implications for Pangea...

Article in International Geology Review · June 2016


DOI: 10.1080/00206814.2016.1190948

CITATION READS

1 511

5 authors, including:

Agustin Cardona Victor A. Valencia


National University of Colombia Washington State University
117 PUBLICATIONS 1,536 CITATIONS 182 PUBLICATIONS 3,176 CITATIONS

SEE PROFILE SEE PROFILE

Yohana Villafanez-Cardona Germán Bayona


University of São Paulo Corporación Geológica ARES
3 PUBLICATIONS 1 CITATION 96 PUBLICATIONS 1,292 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

construction of a balanced cross section at 43°S from the Pacific to the foreland zone and progression
of deformation since the Cretaceous onwards View project

genesis of porphyry copper deposits View project

All content following this page was uploaded by Agustin Cardona on 09 June 2016.

The user has requested enhancement of the downloaded file.


International Geology Review

ISSN: 0020-6814 (Print) 1938-2839 (Online) Journal homepage: http://www.tandfonline.com/loi/tigr20

Provenance of middle to late Palaeozoic


sediments in the northeastern Colombian Andes:
implications for Pangea reconstruction

A. Cardona, V.A. Valencia, A. Lotero, Y. Villafañez & G. Bayona

To cite this article: A. Cardona, V.A. Valencia, A. Lotero, Y. Villafañez & G. Bayona (2016):
Provenance of middle to late Palaeozoic sediments in the northeastern Colombian
Andes: implications for Pangea reconstruction, International Geology Review, DOI:
10.1080/00206814.2016.1190948

To link to this article: http://dx.doi.org/10.1080/00206814.2016.1190948

View supplementary material

Published online: 06 Jun 2016.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tigr20

Download by: [Agustin Cardona] Date: 06 June 2016, At: 07:48


INTERNATIONAL GEOLOGY REVIEW, 2016
http://dx.doi.org/10.1080/00206814.2016.1190948

Provenance of middle to late Palaeozoic sediments in the northeastern


Colombian Andes: implications for Pangea reconstruction
A. Cardonaa, V.A. Valenciab, A. Loteroc, Y. Villafañezd and G. Bayonad
a
Departamento de Procesos y Energía, Universidad Nacional de Colombia, Medellín, Colombia; bSchool of Environment, Washington State
University, Pullman, USA; cDepartamento de Geociéncias y Medio Ambiente, Universidad Nacional de Colombia, Medellín, Colombia;
d
Corporación Geológica Ares, Bogotá, Colombia

ABSTRACT ARTICLE HISTORY


Global-scale Palaeozoic plate tectonic reconstructions have suggested that Laurentia was obliquely Received 20 February 2016
approaching against the northwestern margin of Gondwana until the final agglutination of Pangea. In Accepted 14 May 2016
this contribution integrated petrographic analysis, heavy mineral analysis, and tourmaline geochemistry KEYWORDS
were done, and U–Pb detrital zircon geochronology was obtained, in late Palaeozoic sedimentary and provenance; Palaeozoic;
meta-sedimentary units from the Floresta and Santander Massifs in the Eastern Colombian Andes in U–Pb LA-ICP-MS
order to constrain their provenance and related it with the magmatic, sedimentary, and deformational geochronology; Pangea;
record of the Gondwana–Laurentia convergence until the late Carboniferous to Permian formation of Northern Andes
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Pangea. Late Devonian to early Carboniferous sandstones from the Floresta Massif changed from
sublithoarenites to lithoarenites, tracking the progressive uplift and unroofing of sedimentary and
metamorphic rocks, with associated volcanic activity. The U–Pb detrital zircon geochronology from
the sedimentary and metasedimentary of Floresta and Santander documents Mesoproterozoic and
Palaeoproterozoic sources, and younger Ordovician to Silurian age populations, that can be related to
the early to middle Palaeozoic plutonic rocks and the Amazon Craton. The limited Silurian to Early
Devonian detrital ages that contrast with the more significant Middle to Late Devonian zircons that
document the erosion of contemporaneous magmatic sources formed after a late Silurian to Early
Devonian reduction on the magmatic activity along the proto-Andean margin. These rocks were
apparently deformed and metamorphosed between the late Carboniferous and the early Permian. It
is suggested that the filling and deformation record of these rocks documented the changes in plate
convergence obliquity at the western margin of Gondwana associated with the migration of Laurentia
until its final position in Pangea. Between the late Carboniferous and the early Permian, peri-
Gondwanan continental terranes also collided with the continental margin. Over-imposed Mesozoic
tectonics have contributed to the final redistribution of these terranes to their current position.
Abbreviations:LA: laser ablation inductively couple mass spectrometer; CL: cathodoluminiscence

Introduction South America and Africa to form Pangea (Cawood 2005;


Nance et al. 2010; Torsvik and Cocks 2013). Farther south
Palaeozoic plate tectonic reconstructions of the north-
along the proto-Andean margin of Peru, convergence was
ern segment of the proto-Andean margin have sug-
still active and also punctuated by terrane-related accre-
gested the existence of an extensive proto-Andean
tionary events and/or the continuous subduction of the
Ordovician to Silurian magmatic continental arc linked
Pacific plate (Chew et al. 2008a; Cardona et al. 2009;
to the subduction of the Pacific plate under the South
Mišković et al. 2009; Ramos 2009; Romero et al. 2013).
American margin, that was interrupted by several ter-
Within the eastern segment of the Colombian Andes
rane-related collisional events and/or the switch of the
outcrop several discontinuous greenschist to amphibo-
subduction conditions from extension to compression
lite facies metamorphic units intruded by Ordovician to
(Ramos and Aleman 2000; Cawood 2005; Chew et al.
Silurian magmatic rocks, as well as Devonian to
2007; Cardona et al. 2009; Ramos 2009; Mantilla et al.
Carboniferous sedimentary sequences that locally over-
2012; Van der Lelij et al. 2016).
lie the crystalline rocks (Horton et al. 2010; Restrepo-
Between the Devonian and the Carboniferous,
Pace and Cediel 2010; Mantilla et al. 2012, 2013; Van der
Laurentia, which was dextrally moving with respect to
Lelij et al. 2016; Figure 1).
Gondwana, began its final collision with northeastern

CONTACT Agustin Cardona agcardonamo@unal.edu.co Departamento de Procesos y Energía, Universidad Nacional de Colombia, Medellín, Colombia
Supplemental data for this article can be accessed here.
© 2016 Informa UK Limited, trading as Taylor & Francis Group
2 A. CARDONA ET AL.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Figure 1. a. Map of Paleozoic units from the Andes of Colombia and Peru (Modified from Ramos, 2009). b. Tectonostratigraphic
terranes from the Colombian Andes including Palaeozoic sedimentary rocks (Modified from Restrepo and Toussaint, 1988).

Their early to middle Palaeozoic magmatic and meta- (Restrepo-Pace and Cediel 2010; Mantilla et al. 2012; Van
morphic record has been related to the aforementioned der Lelij et al. 2016).
convergent margin evolution (Ordoñez-Carmona et al., In this contribution we present new U–Pb LA-ICP-MS
2006; Restrepo-Pace and Cediel 2010; Mantilla et al. 2012; geochronology from Palaeozoic metasedimentary rocks of
Van der Lelij et al. 2016). Whereas palaeontological consid- the Santander Massif, together with sandstone petrogra-
erations of the Devonian and Carboniferous sequences phy, heavy mineral analysis, tourmaline geochemistry, and
have suggested the existence of the proximity between U–Pb detrital zircon geochronology from the Late
Gondwana and Laurentia at least in the Late Devonian Devonian to Carboniferous siliciclastic rocks exposed in
(Forero-Suarez 1990; Janvier and Villarroel 1998, 2000; the Floresta Massif, both within the Eastern Cordillera of
Berry et al. 2000; Young and Moody 2002; Burrow et al. the Colombian Andes (Figure 1). These results have been
2003; Moreno-Sanchez 2004). integrated to refine the timing and character of early to
In spite of these major palaeogeographic considera- middle Palaeozoic magmatism (Horton et al. 2010;
tions, the palaeotectonic setting of the Devonian to Restrepo-Pace and Cediel 2010; Mantilla et al. 2012; Van
Carboniferous basins and the deformational record der Lelij et al. 2016), and to evaluate the provenance and
within the framework of the late Palaeozoic proto- palaeotectonic implications of the Devonian–Carboniferous
Andean evolution and the agglutination of Pangea with record within the tectonic evolution of the proto-Andean
associated terrane accretionary events are still controver- margin.
sial (Ordóñez-Carmona et al. 2006; Vinasco et al. 2006; The new provenance results from Palaeozoic sedimen-
Ramos 2009; Horton et al. 2010; Nance et al. 2010; tary and meta-sedimentary rocks from the Eastern
Restrepo et al. 2011; Torsvik and Cocks 2013; Martens Colombian Andes document the transition from an
et al. 2014). Moreover, precise constraints on protolith Ordovician–Early Devonian magmatic dominated margin
ages and metamorphism of the Palaeozoic metamorphic towards a margin characterized by more limited mag-
rocks from eastern Colombia are not fully resolved and matic activity, the formation of several Devonian to
are commonly considered to be early Palaeozoic in age Carboniferous sedimentary basins, and the deformation
INTERNATIONAL GEOLOGY REVIEW 3

of the basins between 300 and 250 Ma as a consequence faults that reactivated the Mesozoic extensional struc-
of major changes in plate convergence relations as tures in the northern segment of the Eastern Cordillera
Laurentia approaches and collides against Gondwana to (Mojica and Villarroel 1984; Cooper et al. 1995; Kammer
form Pangea. and Sanchez 2006; Mora et al. 2010; Saylor et al. 2012).
The Santander Massif (Figure 2) records a transpres-
sional deformational style and is separated from the
Geological setting
Floresta Massif and the Eastern Cordillera by the dex-
The Santander and Floresta Massifs are part of the tral Santa Marta–Bucaramanga Fault (Jimenez et al.
northern segment of the Eastern Cordillera of the 2015). Pre-Mesozoic rocks within this massif include
Colombian Andes, and together with the Upper and amphibolite facies rocks of possible Grenvillian age
the Lower Magdalena Valley, and the eastern flank of (Ward et al. 1974; Restrepo-Pace et al. 1997; Cordani
the Central Cordillera have been included in the et al. 2005), and a well-defined Ordovician to Silurian
Chibcha Terrane (Figure 1; Restrepo and Toussaint plutonic belt (Ward et al. 1974; Restrepo-Pace and
1988). The Llanos and Amazon basins are separated Cediel 2010, Mantilla et al. 2012, 2013; Van der Lelij
from the Chibcha Terrane and the Eastern Cordillera et al. 2016). Metasedimentary rocks include greenschist
by the Guaicaramo Fault system, which is currently an to amphibolite facies rocks of the Silgará Formation
active fault with thrust and strike slip components that can be divided into at least three different groups.
(Restrepo y Toussaint, 1988; Veloza et al. 2013). Some (1) Micaceous schists that are tightly associated with
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

authors considered that this terrane was accreted near early Palaeozoic granitoids may include cordierite or
its actual position in the middle to late Palaeozoic andalusite and sillimanite (Trumpy 1943; Radelli 1967;
(Forero-Suarez 1990; Toussaint 1993), whereas palaeo- Forero 1970a, 1970b). (2) Pelitic to semi-pelitic and
magnetic data from younger Jurassic rocks have sug- calcareous schists with middle pressure Barrovian
gested that during the middle and late Mesozoic there metamorphism (Ríos et al. 2003; Castellanos et al.
have been significant latitudinal south to north trans- 2004, 2008). (3) Very low grade metamorphic rocks
fers of continental slivers, that placed parts of this that were initially related to Silurian or Devonian pro-
terrane near Peruvian and Ecuadorian latitudes toliths and referred to as the metamorphosed Floresta
(Bayona et al. 2006; 2010). Formation (Ward et al. 1974; Forero-Suarez 1990).
To the west, the Central Cordillera includes Ordovician, However, more calcareous lithofacies and their fossil
Permian, and Triassic magmatic and metamorphic rocks content were re-interpreted in terms of Carboniferous
grouped in a series of continental terranes whose accre- age (Moreno-Sanchez et al. 2007).
tionary history is still highly debatable (Restrepo and Less deformed Late Devonian to Carboniferous sedi-
Toussaint 1988; Ordóñez-Carmona et al. 2006; Vinasco mentary units have also been recognized in the
et al. 2006; Martens et al. 2014), together with a series of Santander Massif. They are predominantly siliciclastic
accreted Mesozoic oceanic terranes formed in the Pacific with minor calcareous beds that overlie the igneous
that also extends to the Western Cordillera and the Pacific and metamorphic basement considered to be early
region (Kerr et al. 1997; Villagómez et al. 2011, Figure 1). Palaeozoic in age (Trumpy 1943; Radelli 1967; Forero
The geological record of the Eastern Cordillera 1970a, 1970b; Moreno-Sanchez 2004).
includes Palaeozoic metamorphic, igneous, and sedimen- The Floresta Massif is a NNE–SSW elongated massif.
tary rocks covered by extensive Mesozoic and Cenozoic Major limits of the massif include the Soapaga Fault,
rocks that record the transition from marine to continen- which separates the Palaeozoic and Mesozoic rocks of
tal environments (Etayo-Serna et al. 1983; Cooper et al. the Floresta Massif from Cenozoic sedimentary rocks
1995; Moreno-Sanchez 2004; Gómez et al. 2005a,2005b; (Figure 3, Kimberley 1980; Bayona et al. 2008; Santos
Santos et al. 2008; Horton et al. 2010). The changes from et al. 2008; Saylor et al. 2012), and the Boyacá Fault,
Mesozoic extension to Cenozoic compression in the which juxtaposes Jurassic red beds against the
Andean orogeny (Colleta et al. 1990; Dengo and Covey Cretaceous sedimentary sequences (Mojica and
1993; Cooper et al. 1995; Bayona et al. 2008; Parra et al. Villarroel 1984; Kammer and Sanchez 2006).
2009; Mora et al. 2010) include several Palaeogene to The Floresta Massif includes an Ordovician igneous
Neogene orogenic phases that caused the diachronous and metamorphic basement covered unconformably by
uplift of different segments of the Colombian Andes Devonian to Carboniferous sedimentary rocks (Botero
(Mora et al. 2008a; Bayona et al. 2008; 2012; Parra et al. 1950; Cediel 1969; Mojica and Villarroel 1984; Horton
2009; Nie et al. 2010; Saylor et al. 2012). et al. 2010).
The Santander and Floresta Massifs (Figure 1) are Palaeomagnetic results have suggested that during
related to a series of Palaeogene to Neogene reverse the Jurassic, the Palaeozoic and Mesozoic rocks from
4 A. CARDONA ET AL.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Figure 2. Geological map of the Santander Massif (modified from Ward et al., 1974).

the Floresta Massif were located at southern latitudes, (Toussaint 1993, Figure 1). However, their relation to
probably between northern Ecuador and Southern other Palaeozoic metamorphic and plutonic rocks are
Colombia (Bayona et al. 2006). not well defined, as they are characterized by predomi-
nantly faulted relations. U–Pb geochronological con-
straints from plutonic rocks in the Santander, Floresta,
Middle to late Palaeozoic stratigraphy of the and Quetame Massifs from the Eastern Cordillera have
Floresta and Santander Massifs shown that Ordovician to Silurian magmatism is widely
Ordovician fossiliferous sedimentary rocks have been distributed, and record an apparent decrease in magmatic
found at the eastern flank of the Central Cordillera and activity at ca. 415 Ma (Horton et al. 2010; Restrepo-Pace
the Upper Magdalena Valley of Colombia, and are and Cediel 2010; Mantilla et al. 2012, 2013; Van der Lelij
included in the aforementioned Chibcha Terrane et al. 2016).
INTERNATIONAL GEOLOGY REVIEW 5

Middle to late Palaeozoic sedimentary rocks are also deep marine conditions to a deltaic front (Moreno-
discontinuously exposed at the Eastern Andes of Sanchez 2004).
Colombia (Figure 1) and include predominantly conti- The top of the Palaeozoic sequence includes the
nental to transitional siliciclastic sequences (Toussaint Cuche Formation with ca. 800 m of intercalated micac-
1993; Moreno-Sanchez 2004; Horton et al. 2010). eous mudstones, sandy mudstones, and amalgamated
Middle Palaeozoic rocks of the Floresta Massif sandstones (Cediel 1969; Mojica and Villarroel 1984;
include a Devonian siliciclastic sequence divided in the Moreno-Sanchez 2004). The Cuche Formation is charac-
El Tibet, Floresta, and Cuche Formations (Cediel 1969; terized by abundant flora remnants that suggest a Late
Mojica and Villarroel 1984; Moreno-Sanchez 2004; Devonian Frasnian to Famenian deposition within a
Figure 3). The lower El Tibet Formation rests over ca. delta plain environment (Berry et al. 2000; Janvier and
482 Ma granitoids and associated cordierite and anda- Villarroel 2000; Burrow et al. 2003; Moreno-Sanchez
lusite micaceous schists that were thermally metamor- 2004).
phosed during the intrusion of the Ordovician This Devonian sequence is discordantly covered by
granitoids (Jimenez 2000; Horton et al. 2010). This unit middle Mesozoic red beds and conglomerates from the
has a stratigraphic thickness from 30 to 700 m and Girón Formation (Cediel 1969).
includes a lower segment of claystones and muddy Palaeontological studies including invertebrates,
sandstones that grade to pebbly conglomerates and vertebrates, and plants have suggested a close affi-
sandstones. nity to the middle Palaeozoic sedimentary rocks from
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

The presence of miaspores and stratigraphic charac- the Floresta Massif with North American provinces,
teristics suggest an Emsian accumulation age and a although there are some endemic Gondwanan ele-
transgressive fluvial environment (Mojica and Villarroel ments that suggest some kind of connection between
1984; Janvier and Villarroel 2000; Moreno-Sanchez these two regions (Forero-Suarez 1990; Janvier and
2004). Villarroel 1998, 2000; Berry et al. 2000; Moreno-
This unit is conformably covered by the Floresta Sanchez 2004).
Formation, which includes 400–800 m of grey mud- In the Santander Massif (Figure 2) a series of sand-
stones and black shales that grade towards the top to stones and mudstones of Devonian to Carboniferous
quartz-rich lithic sandstones with interbedded muddy ages (Hubach in Trumpy 1943; Stibane and Forero
sandstones (Moreno-Sanchez 2004). Fossil abundance is 1969; Boinet et al. 1986; Royero 2001) have been corre-
remarkably high and includes brachiopods, corals, trilo- lated with the Floresta and Cuche Formations from the
bites, and fishes of Middle Devonian age (Caster 1939; Floresta Massif as well as other siliciclastic units farther
Royo and Gomez 1942; Morales 1965; Janvier and north in the Perija Range.
Villarroel 1998). Environmental interpretations suggest These units overlie micaceous schists with cordier-
a depositional environment that evolved from relatively ite and granitoids assigned to the Silgará Formation

Figure 3. Geological map of the Floresta Massif (modified from Ulloa et al. 1998).
6 A. CARDONA ET AL.

(Hubach and Trumpy 1943; Royero 2001), which also ±0.01 and 0.04 wt% (1σ). Analytical results are pre-
resemble the Busbanza Schists and Otenga granitoid sented in the supplementary data.
intrusive relations found in the Floresta Massif. Data processing was performed using the software
Recently Moreno-Sanchez et al. (2007) have sug- TOURMAL (Yavuz, 1997). This program is designed for
gested that some of the deformed low-grade meta- the analysis of tourmaline composition. Recalculated
morphic rocks with associated carbonaceous rocks of composition was plotted on the ternary diagrams Ca-
the Santander Massif formerly considered to be Fe (tot)-Mg and Al-Fe (tot)-Mg that discriminate tourma-
Silurian or Devonian in age and coined as the line formed in different rock environments (Henry and
Metamorphosed Floresta Formation, were accumu- Guidotti 1985).
lated in the Carboniferous.

U–Pb geochronology
Methods
Heavy mineral concentrates of the <350 µm fraction
Sandstones petrography and heavy minerals were separated using traditional techniques at
ZirChron LLC. Zircons from the non-magnetic fraction
Modal compositional data of 18 sandstones were obtained
were handpicked under the microscope and mounted
from standard petrographic analysis. Thin sections were
in a 1-inch diameter epoxy puck and slightly ground
stained for potassium and calcium feldspar and 350 frame-
and polished to expose the surface and keep as much
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

work grains were point counted following the Gazzi–


material as possible for laser ablation analyses. After CL
Dickinson method (Dickinson 1985). Petrographic cate-
imaging, the LA-ICP-MS U–Pb analyses were conducted
gories and recalculated modal data are listed in
using a New Wave Nd:YAG UV 213 nm laser coupled to
Supplementary Table 1.
a ThermoFinnigan Element 2 single collector, double-
For heavy mineral analysis, sandstone samples were
focusing, magnetic sector ICP-MS. Operating procedures
crushed, sieved, and hydraulically concentrated in the
and parameters are similar to those of Chang et al.
<400 μm fraction. Subsequently sodium polytungstate
(2006). Laser spot size and repetition rate were 30 µm
(LST®) was used to obtained the >2.89 g/cc fraction.
and 10 Hz, respectively. U and Th concentrations were
Grains were mounted using the Meltmount® resin with
monitored by comparing to NIST 610 trace element
a refraction index of 1.539. Mineral identification of
glass. Two zircon standards were used: Plesovice, with
around 300 translucent grains was carried out following
an age of 338 Ma (Slama et al. 2008) and FC-1, with an
the ribbon-counting method (Mange and Maurer 1992).
age of 1099 Ma (Paces and Miller 1993). Uranium–lead
Results are presented in Supplementary Table 2.
ages were calculated using Isoplot (Ludwig 2007).
For the detrital samples, probability plots were also
Tourmaline geochemistry obtained with the ISOPLOT 3.62. Representative age popu-
lations were considered when more than three grains
Tourmaline is a chemically complex boron-bearing overlapped in age (Gehrels et al. 2006). This statistical
mineral commonly found as an accessory mineral in assumption relies on the fact that individual grains may
granites, and metamorphic rocks formed over a wide sometimes represent lead-loss trajectories. Analytical
range of metamorphic grades (Henry and Guidotti results are presented in Supplementary Table 4.
1985). The complex chemical variability of tourmaline When possible, the youngest detrital zircon age was
and its relationship with its formation environment determined from the weight average age of at least
allows this mineral to be considered an effective indica- three grains at the 1σ level; in other cases we have
tor for sediment provenance analysis (Biernacka 2012). considered the youngest available zircon age, which,
Between 1 and 13 grains from each of seven sand- although not statistically robust and can be compro-
stone samples from the El Tibet Formation (4), Floresta mised by lead-loss trends (Gehrels et al. 2006;
(1), and Cuche Formation (2) were considered for che- Dickinson and Gehrels 2009), can be reasonably used
mical analysis (Supplementary Table 3). The samples with other biostratigraphic evidences.
were analysed using a CAMECA SX-50 electron microp-
robe at the Department of Lunar and Planetary Sciences
at the University of Arizona, USA. Analyses were per- Results
formed with a beam current of 20.0 nA and an accel-
Sampling
erating voltage of 15 kV. Counting time was 10 s for
sodium and 20 s for the rest of the elements. Samples for provenance analysis from El Tibet, Cuche,
Microprobe analytical error varies roughly between and Floresta Formations were obtained from two
INTERNATIONAL GEOLOGY REVIEW 7

different stratigraphic sections in the Floresta Massif of laminated grey to yellowish fossiliferous mudstones,
(Figure 3). We sampled the Potrero Rincón section, followed by 220 m that only adds up to 340 m of
which is widely known for its faunal and floral diversity laminated black shales to yellow quartz rich muddy
(Botero 1950; Mojica and Villarroel 1984; Moreno- sandstones.
Sanchez 2004). Description of this section can be The Cuche Formation is the uppermost unit, and is
found in Botero (1950) and Moreno-Sanchez (2004). characterized at the base by the appearance of reddish
In this section (Figures 3 and 4(a)), the lowest part of sediments (Botero 1950; Moreno-Sanchez 2004). The
El Tibet Formation is 40 m thick and includes 10 m of first 190 m includes micaceous reddish to yellowish
laminated yellowish claystones and muddy sandstones siltstones and muddy sandstones followed by 140 m
followed by 30 m of pebble conglomerates and pebbly of amalgamated quartz-rich sandstones and 430 m of
sandstones. The Floresta Formation includes 485 m of lithic sandstones and intercalated siltstones and
fine grain rocks that can be divided into around 120 m claystones.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Figure 4. Stratigraphic columns of the Potrero Rincón and Otenga sections within the Floresta Massif (new data and modified from
Moreno-Sanchez 2004).
8 A. CARDONA ET AL.

The other stratigraphic section that was reviewed carbonatic protoliths are common, and are character-
and measured is exposed along the road between ized by greenschist to amphibolite facies metamorphic
Otengá and Floresta (Figures 3 and 4(b)). This section conditions (Castellanos et al. 2004; Rios et al. 2008).
includes a basal sandy conglomerate of the El Tibet
Formation that disconformably rests over a deformed
K-feldspar biotite granite with hornblende and biotite Sandstone petrography Floresta Massif
pervasively altered to chlorite and epidote. This unit Sample locations from the El Tibet, Floresta, and Cuche
also includes 100 m of yellowish micaceous sandstones Formations as well as modal percentages are presented
and is followed by 220 m of grey shales and siltstones of in Figures 3 and 4 and Supplementary Table 1.
the Floresta Formation. Sandstones from the El Tibet Formation are medium to
Metasedimentary rocks from the Silgará Formation coarse-grained, with sub-angular to sub-rounded grains,
were sampled near the Pescadero Canyon in the good sorting, and very low matrix content (<2%). They are
Santander Massif (Figure 2), where metamorphic rocks also characterized by concave–convex to sutured grain
are mainly pelitic to semi-pelitic, with few amphibolites contacts, and include silica overgrowth in some quartz
(Ríos et al. 2003). The metamorphic grade increases grains as well as ferruginous cement (approximately
towards the northeast from the biotite to the sillimanite <1%). Compositionally the sandstones can be classified
zone. Other samples were collected in the Motiscua as sub-lithoarenites (Figure 5(a)). Quartz is predominantly
region (Figure 2), where pelites, semi-pelites, and monocrystalline, whereas polycrystalline quartz includes
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Figure 5. Petrographic characterization of sandstones. A. Sandstone classification after Folk (1974). B. Lithic diagram. C. Provenance
discrimination diagram after Dickinson (1985).
INTERNATIONAL GEOLOGY REVIEW 9

diffuse, sedimentary, and foliated varieties. Lithics include ultra-stable minerals including rutile and tourmaline are
micaceous metamorphic fragments and sedimentary frag- less abundant (1–6% and 1–10%). Unstable minerals
ments such as mudstones and sandstones (Figure 5(b)). include epidote (4–9%), chlorite (3%), and tremolite
Very low proportions of amphibole, zircon, opaque miner- (<3%) as well as andalusite and amphibole (<2% and
als, and muscovite are also common. 2–3%), which become more common at the top of the
The Floresta Formation includes coarse-grained sand- formation (Figure 6).
stones, with rounded to sub-rounded and angular The Floresta Formation also includes ultra-stable
clasts, very low matrix content (<1%), and good sorting. and stable heavy minerals (Figure 6) such as muscovite
They are also characterized by concave–convex to (11–48%), zircon (25–51%), and apatite (6–14%) with
sutured grain contacts and ferruginous cement (<1%). lower contents of tourmaline (7–8%) and rutile (1–3%).
Lithics include micaceous metamorphic fragments and Unstable minerals are still common including epidote
sedimentary fragments such as mudstones. They (7–12%), chlorite (3%), and tremolite (<2%).
include predominantly monocrystalline quartz and dif- Within the Cuche Formation, zircon is the most
fuse polycrystalline quartz, followed by sedimentary abundant heavy mineral (21–85%). Other stable and
quartz and foliated polycrystalline quartz. Very low pro- ultra-stable phases such as muscovite (2–61%), apatite
portions of rutile, amphibole, opaque minerals, and (2–19%), tourmaline (4–8%), and rutile (<3%) are also
muscovite are also common. Compositionally the present. Epidote (2–9 %), biotite (<2%), and corundum
Floresta Formation sandstones are also classified as (<2%) are common at the middle of the section
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

sub-lithoarenites (Figure 5(a)). (Figure 6).


Finally, sandstones from the Cuche Formation are
medium to fine grained, with sub-rounded to sub-angu-
Tourmaline geochemistry
lar grains, clay matrix (<2%) and good sorting. Grain
contacts include tangential, sutured, and longitudinal Previous studies on tourmaline geochemistry have
relations. Ferruginous cement is also common, but in shown that the chemical composition of tourmalines
low content (ca. <3%). Compositionally the sandstones does not have a clear relationship with the colour
can be classified as lithoarenites (Figure 5(a)). Lithics seen in the microscope (Henry and Dutrow 1992;
include micaceous schist fragments, mudstones, and Biernacka 2012), therefore we have randomly picked
volcanic clasts (<5%, Figure 5(b)), which are character- tourmalines from different sandstone samples of the
ized by small plagioclase crystals embedded in an apha- three formations.
nitic matrix. Quartz is predominantly monocrystalline, Tourmalines from the El Tibet Formation are green
whereas polycrystalline quartz includes diffuse, sedi- to brown with broken or angular and rounded crys-
mentary, and foliated types. Very low proportions of tals. They are characterized by high Al content
rutile, amphibole, chlorite, apatite, zircon, tourmaline, (Figure 7), high Al-Li-Elbaite (77-67%), and low
opaque minerals, and muscovite are also common, pla- Schorl-Fe (20-10%) and Dravite-Mg (18% and 8%).
gioclase and K-feldspars are present in very low con- The Fe/(Fe+Mg) ratio varies widely between 0.68 and
tent (<2%). 0.37. Calcium values are low varying between 0.25%
Up-section compositional changes are well recorded and 0.03%.
in the Dickinson (1985) provenance diagram, with the Tourmalines from the Floresta Formation are predo-
increase in the lithic proportions recorded by the minantly angular, with rounded to spherical forms and
change from transitional recycled towards the lithic mostly green to brown colours. They include high Al
recycled between El Tibet in the Cuche Formation values with Elbaite (78–56%), Dravite (28–1%), and
(Figure 5(c)). Schorl (22–3%) (Figure 7). Fe/(Fe+Mg) varies between
0.93 and 0.14. The concentrations of Ca are also
low (0.62%).
Heavy minerals
Tourmaline grains from the Cuche Formation are
A total of 15 samples from the El Tibet (5), Floresta (3), angular and prismatic with rounded edges. Analysed
and Cuche Formations (7) were analysed. The strati- tourmaline grains present a high Al content (Figure 7)
graphic distribution of the samples is presented in the with Elbaite (74–67%), Schorl (25–10%), and Dravite
stratigraphic columns and the geological map (Figures 3 (20–0.6%). Fe/(Fe+Mg) ratios vary between 0.89 and
and 4). Results are presented in Supplementary Table 2. 0.36 and Ca concentrations also vary between 0.32%
El Tibet Formation is characterized by a high content and 0.01%.
of ultra-stable to stable heavy minerals such as zircon Within the tourmaline chemical discrimination dia-
(19–53%), muscovite (11–64%), and apatite (19%). Other grams, grains from the three Devonian formations
10 A. CARDONA ET AL.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Figure 6. Distribution of heavy minerals in the Palaeozoic sandstones from the Floresta Massif.

can be related to tourmalines formed in Li-poor U–Pb detrital zircon geochronology


granitoids and Ca-poor and Al-saturated metapelites
Detrital zircons from four sandstone samples from the
(Figure 7). Some grains from the Floresta Formation
El Tibet, Floresta, and Cuche Formations were ana-
are also related to tourmalines derived from Ca-rich
lysed by LA-ICP-MS. Analytical results are presented
low-grade metamorphic rocks (Figure 7).
in Supplementary Table 4.
INTERNATIONAL GEOLOGY REVIEW 11

Figure 7. Tourmaline geochemical discrimination diagrams (Henry and Guidotti 1985). (a) Field 1: Li-rich granitoids, pegmatites and
aplites. Field 2: Li-poor granitoid pegmatite and aplites. Field 3: Hydrothermally altered granitic rocks. Field 4: aluminous metapelites
and metapsammites. 5: Al-poor metapelites and metapsammites. Field 6: Fe3+-rich quartz-tourmaline rocks, calcosilicates and
metapelites. Field 7: low-Ca ultramafic. Field 8: metacarbonates and metapyroxenites. (b) Field 1: Li-rich granitoids, pegmatites
and aplites. Field 2: Li-poor granitoids, pegmatites and aplites. Field 3: Ca-rich metapelites, metapsammites and calcsilicate. Field 4:
Ca-poor metapelites, metapsammites and quartz-tourmaline rocks. Field 5: metacarbonates. Field 6: metapyroxenites.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

One hundred and fourteen detrital zircon ages were 1211, 1006, 1300, and 1062 Ma (Figure 8). The young
analysed from the El Tibet Formation (Sample 640052). 409 Ma age peak constrained a maximum Early
Older ages included Mesoproterozoic detrital zircons Devonian age for accumulation.
that included 50% with age peaks at 1000 and
1502 Ma. Other Precambrian zircons were represented
by Palaeoproterozoic zircons (11%) that formed a ca. Santander Massif
1724 Ma age peak (Figure 8). Neoproterozoic zircons
Silgará Formation
represent 15% of the grains and include a 592 Ma age
peak, whereas Palaeozoic zircons were also common Metamorphic rocks from the Silgará Formation were
(15%) and characterized by two main age peaks of 414 sampled for detrital zircon geochronology in two local-
and 472 Ma (Figure 8). The 414 Ma age peak con- ities where petrological studies have characterized
strained a maximum early Devonian depositional age greenschist to amphibolite facies rocks with Barrovian-
for the El Tibet Formation. type metamorphism (Ríos et al. 2003; Castellanos et al.
One hundred and fourteen detrital zircons from 2004, 2008). Samples include three schists from the
the Floresta Formation (Sample 640038) ages were Motiscua region and four from the Pescadero Canyon
analysed. Detrital zircons included a Mesoproterozoic (Figure 2).
age population that include 36% of the analysed One hundred and fifteen detrital zircons were ana-
grains and was characterized by peaks at 1100 and lysed from a garnet + staurolite schist (Sample AMVO-
1508 Ma, whereas 10% of the zircon grains included 25) collected from the Motiscua region. The major det-
Palaeoproterozoic ages (Figure 8). Palaeozoic zircons rital age population includes 50% of the analysed grains
made up nearly half of the analysed zircons and and spans from the early Palaeozoic to the
represented 43% of the grains, including age peaks Neoproterozoic, with age peaks at 414, 540, 625, and
at 362, 398, and 448 Ma (Figure 8). The youngest 1185 Ma, whereas older Mesoproterozoic to
362 Ma peak suggested a maximum Late Devonian Palaeoproterozoic ages between ca. 1500 and 2257 Ma
depositional age. that represent 15% are also common (Figure 9). The
Two hundred and thirty-six detrital zircon ages were youngest and well-defined age population is Silurian
analysed in two samples from the Cuche Formation and is interpreted as a maximum 414 Ma depositional
(Samples 640095 and 640030). A prominent group of age. Two single grains have yielded 305.9 ± 7.1 Ma and
zircon grains (46%) have early to middle Palaeozoic ages 310.4 ± 6.4 Ma ages, which, although statistically insuf-
with defined peaks at 419, 470, and 452 Ma in the two ficient to defined a relatively robust detrital age popula-
samples (Figure 8). Around 15% of the older analysed tion (Gehrels et al. 2006), suggest that late
detrital grains were of Palaeoproterozoic age and Carboniferous sources may be exposed.
included peaks at 1600 and 1751 Ma (Figure 8), whereas One hundred and one zircons from an albite + mus-
Mesoproterozoic ages were also significant and covite schist (Sample AMVO-26) yielded Palaeozoic
included 30% of the analysed grains with peaks at (57%) and Proterozoic (43%) ages. Main age peaks
12 A. CARDONA ET AL.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Figure 8. U–Pb detrital zircons distribution diagrams from Palaeozoic rocks of the Floresta Massif.

included 368, 400, 481, 621, 1165, and 1500 Ma Zircons from another muscovite + albite sample
(Figure 9). (AMVO-27) were also analysed. From the 106 analyses,
The 368 Ma age peak and two single detrital zircon 36% of the grains yielded Silurian to Devonian ages with
ages of ca. 333 Ma suggest a maximum Late Devonian two main peaks at 362 and 409 Ma. Thirty-eight percent
and probably early Carboniferous age for protolith of the grains yielded Ordovician to Cambrian ages with
formation. an age peak of 478 Ma. The other (less representative)
INTERNATIONAL GEOLOGY REVIEW 13
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Figure 9. U–Pb detrital zircons distribution diagrams from metasedimentary rocks of the Silgará Formation in the Motiscua region.

zircon populations yielded Precambrian ages with some 1960 Ma. Six zircon grains with younger ages between
peaks at 1041 and 1152 Ma (Figure 10). 566 and 1090 Ma were also obtained. U–Pb results from
Six samples from the Pescadero Canyon were ana- 107 zircon grains from a biotite + staurolite schist (AMVO-
lysed for U–Pb detrital zircon geochronology. Ninety- 40) yielded similar Proterozoic ages with age peaks at 936,
eight detrital zircons from a sericitic schist (Sample 1025, and 1206 Ma and older 1348, 1564 Ma (Figure 10).
AMVO-33) exposed at the highest topographic level In contrast with the other samples, the 74 zircon
included detrital ages (94%) and age peaks of mainly grains from the chlorite schist (Sample AMVO-41) yielded
Mesoproterozoic with a less marked amount of early Palaeozoic to Neoproterozoic detrital zircons (11
Palaeoproterozoic zircon grains. The age peaks are grains) of ages between 489 and 634 Ma, and with a
932, 961, 1010, 1157, 1350, and 1543 Ma (Figure 10). defined age peak of 489 Ma and a single zircon grain
The 104 analysed detrital zircon grains from a garnet + with a late Ordovician age of 457 ± 9.5 Ma (Figure 10).
muscovite schist (Sample AMVO-37) yielded Proterozoic The rest of the analysed grains were Mesoproterozoic to
ages with well-defined peaks at 1358, 1557, 1712, and Palaeopreoterozoic with age peak at 1815 Ma.
14 A. CARDONA ET AL.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Figure 10. U–Pb detrital zircons distribution diagrams from metasedimentary rocks of the Silgará Formation in the Pescadero region.

The 98 zircon grains from a muscovite schist with two peaks at 1604 and 1755 Ma. Two single zircon
retrograde chlorite (Sample AMVO-43) yielded mainly grains also yielded younger ages of 512 and 550 Ma.
Mesoproterozoic zircons ages (76% of the analysed Finally 86 zircon grains were analysed from a coarse
grains) with two peaks at 1348 and 1535 Ma grain micaceous schist (Sample AMVO-45) that was
(Figure 10). Older Palaeoproterozoic zircons yielded intruded by granitic dikes. All but one of the grains
INTERNATIONAL GEOLOGY REVIEW 15

yielded Proterozoic ages with early Neoproterozoic Although a detailed discrimination of the different
peaks at 981 Ma and older Mesoproterozoic peaks at units that may include the Silgará Formation need a
1176, 1347, and 1637 Ma (Figure 10). detail cartographic analysis with associated geochrono-
logical results (Mantilla-Figueroa et al. 2016), the U–Pb
detrital geochronological results presented here and the
Discussion review of their geological characteristics, suggest the
existence of at least three different metamorphic units.
Sedimentary accumulation and protolith ages
(1) The micaceous schists that may include andalusite
Palaeontological constraints from the late Palaeozoic and cordierite (not sampled in this contribution), are
rocks of the Floresta Massif have suggested that sedi- discordantly overlain by Devonian rocks or are closely
mentary accumulation extended from the Emsian to the associated with the Ordovician to Silurian magmatic
Famennian. The detrital zircon age peaks obtained in rocks, probably record the early Palaeozoic (Ordovician
this contribution from El Tibet, Floresta, and Cuche to Silurian) metamorphism associated with the mag-
Formations also confirmed their post-Silurian accumula- matic evolution, and were also probably deposited in
tion. Whereas the youngest ca. 362 Ma age peak found the early Palaeozoic (Van der Lelij et al. 2016). (2) In the
in the Floresta Formation, formerly considered as Pescadero and Motiscua regions, the metamorphic
Middle Devonian suggests that accumulation of both rocks are mainly pelitic to semi-pelitic with carbonates,
the Floresta and Cuche Formation have a maximum with a Barrovian metamorphism (Ríos et al. 2003;
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Late Devonian age of accumulation. Castellanos et al. 2004, 2008), and are characterized by
Available age constraints for sedimentary protolith Ordovician, Silurian, and Devonian zircons (this contri-
formation and metamorphism of the Silgará Formation bution; Mantilla-Figueroa et al. 2016).
have been based on local discordance relationships of (3) The metasedimentary rocks formerly included as
some of the metamorphic units with the overlain sedi- the metamorphosed Floresta Formation within the
mentary rocks Middle Devonian in age (Ward et al. 1974; Santander Massif (Ward et al. 1974) and initially consid-
Boinet et al. 1986), and therefore sedimentation and ered to be Silurian in age, have been recently inter-
metamorphism have been considered as a pre-Middle preted based on regional lithofacies as Carboniferous
Devonian event. Some authors have used stable isoto- (Moreno-Sanchez et al. 2007).
pic techniques on metamorphosed carbonate rocks to
suggest that protolith formation is Neoproterozoic in
Provenance
age (Silva et al. 2004).
Recent geochronological constraints from Palaeozoic We will discuss the provenance of the Floresta Massif
magmatic rocks that have been considered as closely including the integrated petrography, heavy minerals,
associated with the metamorphic rocks of the Silgará and U–Pb geochronology, whereas for the Silgará
Formation, have suggested that metamorphism may Formation we will discuss from a U–Pb detrital zircon
extend between 479 and 439 Ma (Restrepo-Pace and geochronological perspective. Such an approach allows
Cediel 2010; Mantilla et al. 2012; Van der Lelij et al. tackling the age(s) of source areas and the identification
2016). of nearly contemporaneous magmatic activity.
The new detrital zircon geochronological results from
the studied localities in the Pescadero and Motiscua
Floresta Massif
regions suggest a more complex temporal framework
for the accumulation and metamorphism of the Silgará The presence of high quartz content and sedimentary
Formation. (1) In the Pescadero Creek one sample has to metasedimentary lithics (Figure 5), as well as the
yielded an Ordovician peak of 489 Ma and therefore heavy mineral content marked by the abundance of
suggests that accumulation post-date the early ultra-stable phases (Figure 6), suggests that the source
Ordovician (Figure 10). A statistically limited single region for the Early Devonian El Tibet Formation within
grain age of ca. 457 Ma may suggest that accumulation the Floresta Massif includes recycled sedimentary and
and metamorphism may have taken place after the metasedimentary rocks. Due to the limited presence of
middle Ordovician. (2) In the Motiscua region 368 and greenschist or amphibolites in the Palaeozoic meta-
414 Ma detrital zircon age peaks (Figure 9), and a single morphic sequence (Busbanza Schists, Jimenez 2000),
grain with ca. 315 Ma from the Silgará Formation sug- we related the presence of chlorite, tremolite, and epi-
gest that their accumulation and metamorphism took dote (Figure 6) to granitoid rocks modified by hydro-
place after the Early Devonian and therefore meta- thermal alteration, such as is found in the underlying
morphism is also post-Devonian in age. Otenga Granitoid along the Otenga–Floresta road.
16 A. CARDONA ET AL.

The sub-angular texture and the preservation of These results together with the compositional trend of
labile sedimentary lithics (Figure 5) such as mudstones sandstones suggest that basin filling in the Floresta
and the presence of chlorite, tremolite, and hornblende Massif included source areas in the western margin of
in the heavy mineral fraction suggest low residence the Amazon Craton or a derived Palaeozoic terrane, with
time in the source area and fast burial. a proximal source in the Ordovician to Silurian magmatic
When the compositional characteristics of the El and metamorphic basement of the Busbanza Schist and
Tibet Formation are compared with the overlying for- Otenga Granitoid. The apparent increase of sandstone
mations, both the Floresta and Cuche Formations are immaturity as documented by the increase in the pre-
more immature and characterized by an increase in the sence of sedimentary lithics and the appearance of vol-
sedimentary and metamorphic lithics, as well as char- canic sources in the Floresta and Cuche Formations, and
acterized by the presence of feldspar and volcanic the presence of a more significant Ordovician to Silurian
lithics (Figure 5). Their heavy minerals contents also sedimentary signature, as well as the Devonian to
include amphibole and epidote (Figure 6), which can Carboniferous detrital zircon basin filling recorded uplift
be related to the aforementioned plutonic source. in the source area with associated magmatic activity,
Tourmaline can resist multiple sedimentary cycles which is a common known feature of a basin fill during
(Mange and Morton 2007) and although it is not possi- orogenic growth (Dickinson 1985).
ble to discriminate whether it came from recycling or
from pristine crystalline rocks, the geochemical compo-
Silgará Formation Santander Massif
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

sitions of the analysed tourmalines are also compatible


with the already recognized granitoid and metapelitic Similar to the Palaeozoic units from the Floresta Massif,
provenance (Figure 7). the studied samples from the Silgará Complex are
U–Pb detrital zircon geochronology from the late characterized by abundant Mesoproterozoic and
Palaeozoic units of the Floresta Massif are characterized Palaeoproterozoic age populations, as well as more
by similar Palaeoproterozoic, Mesoproterozoic, and limited early Palaeozoic ages (Figures 10 and 11).
Ordovician to Silurian age peaks (Figure 8). Both the Silurian to Devonian aged zircons were found to the
Floresta and Cuche Formations show a relative increase east in schists from the Motiscua region (Figure 10).
in the proportion of early to middle Palaeozoic zircons, As mentioned, the Mesoproterozoic and
including Devonian zircons (Figure 8). Palaeoproterozoic detrital zircon ages are also similar
The Ordovician to Silurian zircons are characteristic of to source areas in the Amazon Craton and the Andean
the early Palaeozoic magmatic belt that characterized Grenvillian inliers (Ibañez-Mejía et al. 2011) therefore
the proto-Andean margin (Chew et al. 2007; Ramos suggesting that accumulation took place in the western
2009), and are also well represented by the early margin of the Amazon Craton or a Palaeozoic terrane or
Palaeozoic granitoid bodies found within the Floresta ribbon detached from the Amazon Craton.
Massif (Horton et al. 2010). The Late Devonian to early The Ordovician ages are similar to the nearby gran-
Carboniferous zircons are close to the palaeontological itoids and indicate that the adjacent basement was a
depositional ages (Janvier and Villarroel 1998, 2000; major source area during accumulation of the protoliths
Berry et al. 2000; Young and Moody 2002; Burrow (Restrepo-Pace and Cediel 2010; Mantilla et al. 2012,
et al. 2003; Moreno-Sanchez 2004) and therefore sug- 2013; Van der Lelij et al. 2016). In the Motiscua region
gest that basin filling was accompanied by magmatic the presence of Devonian zircons suggests that at least
activity. the Motiscua protoliths were formed in a basin source
Proterozoic 1750, 1600, 1300, and 1500 Ma together by magmatic rocks of late Palaeozoic ages, also similar
with Grenvillian age peaks (1000–1100 Ma) are similar to to those found in the Floresta Massif.
source ages from the western margin of the Amazon
Craton in the eastern Colombia sub-Andean region
Palaeotectonic considerations
(Ibañez-Mejia et al. 2011) and the Grenvillian-aged mas-
sifs that characterize the Northern Andes (Cordani et al. Permo-Triassic palaeogeographic reconstructions of the
2005; Cardona et al. 2010; Ibañez-Mejía et al., 2011). Northern Andes within Pangea supercontinent have
Zircon crystallization ages between 1500 and 1600 Ma suggested that the continental fragments of southern
are a major characteristic of the Amazon Craton and Mexico overlap with the tectonostratigraphic terranes
therefore link the filling of this basin to a source area in that formed the Northern Andes (Bullard et al. 1965;
South America (Tassinari and Macambira 1999; Cordani Pindell and Dewey 1982; Pindell et al. 1988). This has
et al. 2005; Martens et al. 2014), precluding a derivation left a major ‘room’ problem commonly solved by the
of this domain from Laurentia. redistribution of the Northern Andean and Mexican
INTERNATIONAL GEOLOGY REVIEW 17

Figure 11. Palaeogeographic reconstruction of the Santander and Floresta Massifs and other segments of the Colombian Andes
during the Jura–Cretaceous based on palaeomagnetic results (Bayona et al. 2006). F/B, Floresta Massif; SM, Santander Massif; SMM,
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Santa Marta Massif; SL, San Lucas Range; ECC, Eastern Central Cordillera.

continental terranes along the margin between Peru Massifs is associated with sinestral strike slip displace-
and Venezuela (Restrepo-Pace et al. 1997; Martens ment of the Santa Marta–Bucaramanga Fault (Jimenez
et al. 2014). et al. 2015).
Palaeomagnetic results from the Mesozoic cover of In the next section we will discuss the Palaeozoic
the Floresta Massif, and other segments of the palaeotectonic setting of these units and their relation
Colombian Andes have suggested that these crustal with the growth of the proto-Andean margin from Peru
segments were probably located in a southern latitude, to Venezuela as they probably represent a common
between northern Peru and southern Colombia (Bayona continental margin (Figure 12).
et al. 2006, 2010), and therefore offered alternatives to
the spatial discrepancy.
The U–Pb detrital zircon geochronology from the sedi-
Ordovician to Silurian magmatism and
mentary and metasedimentary rocks of both the Floresta
metamorphism
and Santander Massif have shown their affinity to the
western margin of the Amazon Craton, with major simi- Geochronological constraints from plutonic rocks in the
larities to Proterozoic zircon crystallization ages found in Santander, Floresta, and Quetame Massifs within the
the Putumayo Orogen in southern Colombia. These, Eastern Cordillera of Colombia (Horton et al. 2010;
together with their common Ordovician magmatic Restrepo-Pace y Cediel, 2010; Mantilla et al. 2012; Van
record (Horton et al. 2010; Restrepo-Pace and Cediel der Lelij et al. 2016), the Merida Andes of Venezuela
2010; Mantilla et al. 2012; Van der Lelij et al. 2016) and (Van der Lelij et al. 2016), and the Eastern Cordillera of
correlation of middle Palaeozoic sedimentary rocks, place the Peruvian Andes (Chew et al. 2007; Cardona et al.
their Palaeozoic record near the margin of the north- 2009) confirmed the wide distribution of Ordovician to
western Amazon Craton. Silurian magmatism in the Northern Andes, and the
However, the current spatial distribution has been importance of Ordovician magmatism along the wes-
modified by a series of Meso-Cenozoic over-imposed tern margin of Gondwana (Cawood 2005, Figure 13(a)).
tectonic events. Following the discussion of palaeomag- The geochemical characters of the magmatic rocks
netic data by Bayona et al. (2006), it turns out that these within the Santander Massif have been related to an arc
units may be placed in southernmost Colombia and setting characterized by continuous magmatism until
northern Ecuador during the Middle Jurassic, and that ca. 430 Ma (Mantilla et al. 2012; Van der Lelij et al.
they have been transported along the margin to the 2016). This arc-related magmatism is well recognized
north as a consequence of oblique subduction at least along most of the Andean chain including Argentina,
until the Early Cretaceous (Figure 11). Subsequent Bolivia, and Peru and is called the Famatinian arc due to
Cenozoic disruption of the Santander and Floresta the well-constrained record of this arc in the Andes of
18 A. CARDONA ET AL.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Figure 12. Correlation of Palaeozoic terranes from the Colombian, Venezuela, and Peruvian Andes and the southern Mexican
Terranes (names of units and major events after Toussaint 1993; Moreno-Sanchez 2004; Ortega-Obregón et al., 2008; Laya and
Tucker, 2012; Van der Lelij et al. 2016).

Argentina, and has been linked to the subduction of the Cordierite-bearing Busbanza Schist from the Floresta
Pacific Plate under the South American continent Massif that is discordantly covered by the late
(Pankhurst et al. 1998; Miller and Söllner 2005; Chew Palaeozoic sedimentary units is included within this
et al. 2008a; Ramos 2009; Mišković et al. 2009; Mantilla early Palaeozoic metamorphism.
et al. 2012; Van der Lelij et al. 2016). Magmatic activity seems to have continued until the
The early Palaeozoic metamorphic record of the late Silurian as indicated by plutonic crystallization and
Silgará Formation within the Santander Massif have detrital zircon age populations on the Devonian rocks.
also been associated with the continuous evolution of Plutonic and metamorphic rocks were exposed after the
the convergent margin, where the alternation of exten- Silurian as they were covered by Devonian sedimentary
sional and compressional phases, that allow the open- rocks, and therefore document a major exhumation
ing and closing the back-arc or intra-arc basin is event at the end of the Silurian. Such an event has
responsible for the formation of metamorphic belts been related either to the collision and accretion of a
associated with the magmatic activity (Collins 2002; continental terrane (Ordóñez-Carmona et al. 2006) or to
Van der Lelij et al. 2016). This metamorphism is tempo- active subduction. Such a Silurian event has also been
rally constrained between 471 and 439 Ma (Mantilla identified in the Eastern Cordillera of the Peruvian
et al. 2012, 2013; Van der Lelij et al. 2016). Andes (Cardona 2006).
We suggest that this metamorphic evolution may be Although results from this contribution do not allow
restricted to the metamorphic rocks that are spatially us to discriminate between collisional and subduction-
related to the granitoids and/or have closed stratigraphic related models, global Palaeogeographic reconstruc-
relations with overlying late Palaeozoic sediments. tions suggest that Laurentia was moving closer to
INTERNATIONAL GEOLOGY REVIEW 19
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Figure 13. Palaeozoic palaeogeographic reconstructions modified from Torsvik and Cocks (2013). MT, Mexican Terranes; V,
Venezuelan Andes; S, Santander Massif; F, Floresta Massif; P, Peruvian Andes. (a) Silurian (440 Ma) reconstruction, (b) Middle
Devonian (370 Ma) reconstruction, (c) Mississippian (340 Ma) reconstruction, (d) Pennsylvanian (310 Ma) reconstruction.

Western Gondwana, and therefore microcontinents or associated with a high-grade metamorphism, Pre-Silurian
continental blocks formed either during the break-up of sedimentation, and a ca. 420 Ma metamorphic event
Rodina or during the Ordovician to Silurian extensional (Cardona 2006; Chew et al. 2007, 2008a; Cardona et al.
phases of the convergent margin could have been re- 2009; Mišković et al. 2009).
incorporated into the continental margin (Chew et al. Mexican terranes with relatively well studied
2008b; Torsvik and Cocks 2013; Shao et al. 2015) form- Palaeozoic rocks include the Oaxaquia, Mixteca, and
ing the middle Palaeozoic orogenic events. Maya terranes (see reviews in Keppie et al. 2008;
Within the Eastern Cordillera of the Peruvian Andes Weber et al. 2008; Nance et al. 2009; Martens et al.
outcrop, the Marañon Complex is considered a composite 2010). Such terranes have been placed near the
tectonostratigraphic unit with a protracted Palaeozoic Amazon Craton and western Gondwana, based on a
record. It includes a major Ordovician magmatism characteristic Mesoproterozoic Amazon Craton zircon
20 A. CARDONA ET AL.

signature and early Palaeozoic Gondwanan faunas and which may reflect the collision with Laurentia that
are also characterized by Ordovician magmatism. start the final agglutination of Pangea (Figure 13(d)).
(Keppie et al. 2008; Weber et al. 2008; Nance et al. In the Peruvian Andes, a Devonian to Carboniferous arc
2009; Martens et al. 2010; Ibañez-Mejía et al. 2015). related setting extends between ca. 360 and 310 Ma and
was followed by a significant magmatic hiatus, whereas in
the Oaxaquia and Mixteco terranes, late Palaeozoic rocks
include Devonian to early Carboniferous detrital zircons
Devonian to early Carboniferous basin and
(340–390 Ma) that have been related to a poorly exposed
magmatic evolution
magmatic arc of such an age, and a high-pressure meta-
It has long been suggested that after the Silurian a morphic unit of ca. 352 Ma in the Acatlan complex is
period of magmatic quiescence characterized the con- interpreted as related to a subduction setting (Review in
tinental margin of South America, north of Northern Estrada-Carmona et al. in press).
Chile (Bahlburg and Hervé 1997; Chew et al. 2008a; Although varying in some details, the active margin
Bahlburg et al. 2009; Cardona et al. 2009; Mišković setting is similar to the one recorded in Colombia.
et al. 2009; Van der Lelij et al. 2016). These authors
have suggested that the hiatus lasted from ca. 420
Late Carboniferous to Permian metamorphism and
until 335 Ma in the Peruvian Andes, and until ca.
deformation
300 Ma in the Northern Andes of Colombia.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

The detrital zircon ages from the Devonian rocks of Stratigraphic relations in the Floresta Massif indicate
the Floresta Formation, and those of the Silgará that late Palaeozoic units are in angular unconformity
Formation rocks near Motiscua, suggest that magmatic with Early Jurassic red beds (Cediel 1969; Mojica and
activity was common in the Devonian, although the Villarroel 1984). This relation suggests that a major
magmatic volumes showed a decrease when compared deformational event took place between the Late
with the older record. Ibañez-Mejía et al. (2013) have Devonian and the Early Jurassic.
also suggested that Silurian to Devonian zircons have In some segments of the Santander Massif, and to
been found in the Llanos Basin of Colombia and there- the southeast of the Eastern Cordillera, fossiliferous
fore confirmed the importance of this event. Permian rocks are exposed and characterized by a pro-
We suggest that the change in the magmatic volumes minent calcareous sequences with intercalated mud-
between the Ordovician–Silurian and the Devonian, may stones that follow a clastic sandstone segment at the
be a response to the oblique convergence of the Pacific base (Stibane and Forero 1969). These units overlie in
and Rheic Ocean, as Laurentia was moving to the north- discordance the Devonian to Carboniferous sequences
east and the ocean that separated the two continental similar to rocks from the Floresta Massif (Toussaint
masses was closed (Cocks and Torsvik 2002; Torsvik and 1993). Such stratigraphic relationships, and the appar-
Cocks 2013; Domeier and Torsvik 2014; Figure 13(b, c)). ent absence of late Carboniferous sedimentation, also
Palaeontological constraints from the Devonian sedimen- confirmed that a major deformational event may have
tary rocks of the Floresta Massif show evidence for a taken place between the late Carboniferous and the
significant proximity with Laurentia as expected by the Permian.
approach of this continent to the South American Margin In the Motiscua region, detrital age peaks for the
to finally close the Rheic Ocean and form Pangea (Forero- Silgará Formation also suggest that accumulation may
Suarez 1990; Janvier and Villarroel 1998, 2000; Berry et al. have taken place after ca. 365 Ma and may be as young
2000; Young and Moody 2002; Burrow et al. 2003; as 311 Ma, whereas the presence of intrusive relations
Moreno-Sanchez 2004). with ca. 210 Ma undeformed granitoids suggest that
It is therefore suggested that the transgressive basin- metamorphism took place in this time interval
filling recorded by the El Tibet Formation may be part of (Mantilla et al. 2012; Van der Lelij et al. 2016).
this scenario of oblique convergence and migration of Metamorphic constraints from the Silgará Formation
Laurentia, which again became more compressive in the in the Motiscua region are characterized by middle-
Late Devonian and the early Carboniferous as sug- pressure, Barrovian-type metamorphism, which can be
gested by the provenance change in the Floresta and related to a collisional event (Castellanos et al. 2008).
Cuche Formations, the increase in sedimentary and In the Llanos Basin of Eastern Colombia, recent seis-
metamorphic lithics and the reappearance of a mag- mic and well analyses have suggested that a major late
matic input. Palaeozoic, post-Devonian event was recorded in the
Within the Merida Andes of Venezuela, magmatic subsurface (Suárez Díaz and Solano 2012; Moreno-
activity seems to have stopped until the Mesozoic, López and Escalona 2015), and therefore confirmed
INTERNATIONAL GEOLOGY REVIEW 21

the importance of the late Palaeozoic event along the supported A. Lotero and Y. Villafañez. COLCIENCIAS and
Andean margin. Corporación Geológica ARES also provided a short term fellow-
Devonian to Lower Pennsylvanian slates and quart- ship support to Y. Villafañez, G. Cañizales and W. Echavarria for
their help with sample preparation. We also thank M. Ibañez
zites of the Venezuelan Andes were apparently metamor-
and O. Talavera for discussions during field work and for pro-
phosed in the Late Pennsylvanian and subsequently viding access to some petrographic samples. We thank S.
covered by a Middle Pennsylvanian fluvial succession Zapata and K. Domanick for their help in LA-ICP-MS and
made of sandstones and conglomerates, which recorded microprobe data acquisition. We thank J. S. Jaramillo, J. C.
a regressive tectonic event (reviews in Laya and Tucker, Piedrahita and A. M. Patiño for their help with the editing of
the manuscript. We thank J. Vanegas for discussions and final
2012). In the Peruvian Andes and the Mexican Terranes a
organization of the manuscript. We are enormously grateful to
late Carboniferous to early Permian deformational and/or two anonymous reviewers and the editorial care of R. Stern for
metamorphic event has been recorded (Cardona 2006; their comments and suggestions, which improved the
Chew et al. 2007; Grodzicki et al. 2008; Cardona et al. manuscript.
2009; Mišković et al., 2009). Some of these terranes were
subsequently affected by a middle to late Permian arc
Disclosure statement
magmatism that placed them in a Pacific position out-
side the core of the Pangean supercontinent and there- No potential conflict of interest was reported by the authors.
fore suggest that this event may be related to more local
terrane related accretionary events.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

It is therefore suggested that this late Palaeozoic Funding


event is related to the late stages of Pangea agglutina- We thank the Asociación Colombiana de Geólogos y
tion, which included the collision between Laurentia Geofísicos del Petróleo (ACGGP), for the ARES-CORRIGAN
and Gondwana, and the accretion of several continental grants, which supported A. Lotero and Y. Villafañez.
terranes and ribbons located both in the upper plate, or COLCIENCIAS and Corporación Geológica ARES also provided
a short-term fellowship support to Y. Villafañez.
transferred from the Pacific plate after the closure of the
Rheic Ocean (Figure 13(d)).
Although more precise temporal constraints are neces- References
sary to understand these deformational and metamorphic
Bahlburg, H., and Hervé, F., 1997, Geodynamic evolution and
events, the Colombian, Mexican, and Venezuelan terranes
tectonostratigraphic terranes of NW-Argentina and N-Chile:
are covered by correlatable Permian Carbonates that also Geological Society of America Bulletin, v. 109, p. 869–884.
share faunal affinities with North America (see review in doi:10.1130/0016-7606(1997)109<0869:GEATTO>2.3.CO;2
Laya and Tucker, 2012), and therefore reinforce the spatial Bahlburg, H., Vervoort, J.D., Du Frane, A., Bock, B., Augustsson,
proximity of all these terranes. C., and Reimann, C., 2009, Timing of crust formation and
recycling in accretionary orogens: Insights learned from the
western margin of South America: Earth-Science Reviews, v.
Conclusions 97, p. 215–241. doi:10.1016/j.earscirev.2009.10.006
Bayona, G., Cortés, M., Jaramillo, C., Ojeda, G., Aristizabál, J.J.,
New provenance constraints from the late Palaeozoic and Reyes-Harker, A., 2008, An integrated analysis of an
sedimentary rocks from the Floresta Massif and metase- orogen sedimentary basin pair: Latest Cretaceous-Cenozoic
dimentary rocks from the Santander Massif record evolution of the linked Eastern Cordillera orogen and the
Llanos foreland basin of Colombia: Geological Society of
Devonian to Permian basin filling, magmatism and
America Bulletin, v. 120, no. 9–10, p. 1171–1197.
deformation associated with the evolution of the con- doi:10.1130/B26187.1
tinental margin before and during the final agglutina- Bayona, G., Rapalini, A., and Costanzo-Alvarez, V., 2006,
tion of Pangea. The tectonostratigraphic record within Paleomagnetism in Mesozoic rocks of the northern Andes
these and other limitedly studied late Palaeozoic units and its implications in Mesozoic tectonics of northwestern
South America: Earth, Planets and Space, v. 58, p. 1255–
of the northern Andes offer a major picture on the
1272. doi:10.1186/BF03352621
continuous palaeogeographic changes that affected Bayona, G.A., Cardona, A., Jaramillo, C.A., Mora, A., Montes, C.,
the continental margins and the ocean as Gondwana Valencia, V., Ayala, C., Montenegro, O., and Ibañez, M., 2012,
and Laurentia achieved their positions within the Early Paleogene magmatism in the northern Andes: Insights
Pangea supercontinent. on the effects of Oceanic Plateau-continent convergence:
Earth and Planetary Science Letters, v. 331–332, p. 97–111.
doi:10.1016/j.epsl.2012.03.015
Acknowledgements Bayona, G.A., Jimenez, G., Silva, C., Cardona, A., Montes, C.,
Roncancio, J., and Cordani, U., 2010, Paleomagnetic data
We thank the Asociación Colombiana de Geólogos y Geofísicos and K-Ar ages from Mesozoic units of the Santa Marta
del Petróleo (ACGGP), for the ARES-CORRIGAN grants that Massif: A preliminary interpretation for block rotations and
22 A. CARDONA ET AL.

translations: Journal of South American Earth Sciences, v. Cawood, P.A., 2005, Terra Australis orogen: Rodinia breakup
29, p. 817–831. doi:10.1016/j.jsames.2009.10.005 and development of the Pacific and Iapetus margins of
Berry, C.M., Morel, E., Mojica, J., and Villarroel, C., 2000, Devonian Gondwana during the Neoproterozoic and Paleozoic:
plants from Colombia, with discussion of their geological and Earth-Science Reviews, v. 69, p. 249–279. doi:10.1016/j.
palaeogeographical context: Geological Magazine, v. 137, p. earscirev.2004.09.001
257–268. doi:10.1017/S0016756800003964 Cediel, F., 1969, Geologia del Macizo de Floresta: Memoria,
Biernacka, J., 2012, Provenance of upper Cretaceous quartz- Primer Congreso Colombiano de Geologia, v. 1, p. 17–29.
rich sandstones from the North Sudetic Synclinorium, SW Chang, Z., Vervoort, J.D., McClelland, W.C., and Knaack., C.,
Poland: Constraints from detrital tourmaline: Geological 2006, U-Pb dating of zircon by LA-ICP-MS: Geochemistry,
Quarterly, v. 56, p. 315–332. doi:10.7306/gq.1024 Geophysics, Geosystems, v. 7, p. 1–14. doi:10.1029/
Boinet, T., Babin, C., Bourgois, J., Brutin, J., Lardeux, H., Ponds, 2005GC001100
D., and Racheboeuf, P., 1986, Les grandes étapes de Chew, D., Schaltegger, U., Kosler, J., Whitehouse, M.J., Gutjahr,
l’évolution paléozoïque du massif de Santander (Andes de M., Spikings, R.A., and Miskovic, A., 2007, U–Pb geochrono-
Colombie): Signification de la discordance du Dévonien logic evidence for the evolution of the Gondwanan margin
moyen: Comptes Rendus de l’Académie des Sciences. of the north-central Andes: Geological Society of America
Série 2, Mécanique, Physique, Chimie, Sciences de l’univers, Bulletin, v. 119, p. 697–711. doi:10.1130/B26080.1
Sciences de la Terre, v. 303, p. 707–712. Chew, D.M., Flowerdew, M.J., Page, L.M., Crowley, Q.G., Daly, J.
Botero, C., 1950. Reconocimiento geológico del área compren- S., Cooper, M.J., and Whitehouse, M.J., 2008a, The tecto-
dida entre los municipios de Belén, Cerinza, Floresta, Nobsa nothermal evolution and provenance of the Tyrone
y Santa Rosa de Viterbo, Depto. de Boyacá. Ingeominas, Central Inlier, Ireland: Grampian imbrication of an outboard
Informe 534: CEGOC, v. 8, p. 244–311. Laurentian microcontinent?: Journal of the Geological
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Bullard, E., Everett, J.E., and Smith, A.G., 1965, The fit of the Society, London, v. 165, p. 675–685. doi:10.1144/0016-
continents around the Atlantic: Philosophical Transactions 76492007-120
of the Royal Society A: Mathematical, Physical and Chew, D.M., Magna, T., Kirkland, C.L., Miskovic, A., Cardona, A.,
Engineering Sciences, v. 258, p. 41–51. doi:10.1098/ Spikings, R., and Schaltegger, U., 2008b, Detrital zircon fin-
rsta.1965.0020 gerprint of the Proto-Andes: Evidence for a Neoproterozoic
Burrow, C.J., Janvier, P., and Villarroel, C., 2003, Late Devonian active margin?: Precambrian Research, v. 167, p. 186–200.
acanthodians from Colombia: Journal of South American doi:10.1016/j.precamres.2008.08.002
Earth Sciences, v. 16, p. 155–161. doi:10.1016/S0895-9811 Cocks, L.R.M., and Torsvik, T.H., 2002, Earth geography from
(03)00026-9 500 to 400 million years ago: A faunal and palaeomagnetic
Cardona, A., 2006, Reconhecimento da evolução tectônica da review: Journal of the Geological Society, v. 159, p. 631–644.
proto-margem Andina do Centro-Norte Peruano, baseada doi:10.1144/0016-764901-118
em dados geoquímicos e isotópicos do embasamento da Colleta, B., Hébrard, F., Letouzey, J., Werner, P., and
Cordilheira Oriental na região de Huánuco-La Unión [Tese Rudkiewikz, J.L., 1990, Tectonic style and crustal structure
de Doutoramento]. São Paulo, Universidade de São Paulo. of the Eastern Cordillera (Colombia), from a balanced cross
247 p. section, in Letouzey, J., ed., Petroleum and tectonics in
Cardona, A., Chew, D., Valencia, V.A., Bayona, G., Mišković, A., mobile belts: Paris, Editions Technip, Paris 11990, p. 81–100.
and Ibañez-Mejía, M., 2010, Grenvillian remnants in the Collins, W.J., 2002, Hot orogens, tectonic switching, and crea-
northern Andes: Rodinian and Phanerozoic paleogeo- tion of continental crust: Geology, v. 30, p. 535–538.
graphic perspectives: Journal of South American Earth doi:10.1130/0091-7613(2002)030<0535:HOTSAC>2.0.CO;2
Sciences, v. 29, p. 92–104. doi:10.1016/j.jsames.2009.07.011 Cooper, M.A., Addison, F.T., Alvarez, R., Coral, M., Graham, R.H.,
Cardona, A., Cordani, U., Ruiz, J., Valencia, V.A., Armstrong, R., Hayward, A.B., Howe, S., Martinez, J., Naar, J., Peñas, R.,
Nutman, A., and Sanchez, A., 2009, U/Pb zircon and Nd Pulham, A.J., and Taborda, A., 1995, Basin development and
isotopic signatures of the pre-Mesozoic metamorphic base- tectonic history of the Llanos Basin, Eastern Cordillera, and
ment of the Eastern Peruvian Andes: Growth and prove- Middle Magdalena Valley, Colombia: American Association of
nance of a late Neoproterozoic to Carboniferous Petroleum Geologists Bulletin, v. 79, p. 1421–1443.
accretionary orogen on the Northwest margin of Cordani, U.G., Cardona, A., Jimenez, D.M., Liu, D., and Nutman,
Gondwana: Journal of Geology, v. 117, p. 285–305. A.P., 2005, Geochronology of Proterozoic basement inliers
doi:10.1086/597472 in the Colombian Andes: Tectonic history of remnants of a
Castellanos, A., Óscar, M., Ríos, R., Carlos, A., and Takasu, A., fragmented grenville belt, in Vaughan, A.P.M., Leat, P.T., and
2008, A new approach on the tectonometamorphic Pankhurst, R.J., eds., Terrane processes at the margins of
mechanisms associated with PT paths of the Barrovian- Gondwana: The Geological Society of London Special
type Silgará Formation at the Central Santander Massif, Publication, v. 246, p. 329–346.doi:10.1144/GSL.
Colombian Andes: Earth Sciences Research Journal, v. 12, SP.2005.246.01.13
no. 2, p. 125–155. Dengo, C.A., and Covey, M.C., 1993, Structure of the Eastern
Castellanos, O., Rios, C., and Takasu, A., 2004, Chemically sec- Cordillera of Colombia: Implications for trap styles and
tor-zoned garnets in the metapelitic rocks of the Silgará regional tectonics: American Association of Petroleum
Formation in the Central Santander Massif, Colombian Geologists Bulletin, v. 77, p. 1315–1337.
Andes: Occurrence and growth history: Boletin de Dickinson, W., 1985, Interpreting provenance relations from
Geología, v. 26, p. 9–18. detrital modes of sandstones, in Zuffa, G.G., ed.,
Caster, K.E., 1939, A Devonian fauna from Colombia: Bulletin of Dordrechet, D. Reidel Publishing Company, NATO ASI
American Paleontology, v. 24, p. 101–318. Series, v. 48, p. 333–362.
INTERNATIONAL GEOLOGY REVIEW 23

Dickinson, W.R., and Gehrels, G.E., 2009, Use of U–Pb ages of Ibañez-Mejía, M., Pullen, A., Arenstein, J., Gehrels, G.E., Valley,
detrital zircons to infer maximum depositional ages of J., Ducea, M.N., Mora, A.R., Pecha, M., and Ruiz, J., 2015,
strata: A test against a Colorado Plateau Mesozoic database: Unraveling crustal growth and reworking processes in com-
Earth and Planetary Science Letters, v. 288, p. 115–125. plex zircons from orogenic lower-crust: The Proterozoic
doi:10.1016/j.epsl.2009.09.013 Putumayo Orogen of Amazonia: Precambrian Research, v.
Domeier, M., and Torsvik, T.H., 2014, Plate tectonics in the late 267, p. 285–310. doi:10.1016/j.precamres.2015.06.014
Paleozoic: Geoscience Frontiers, v. 5, p. 303–350. Ibañez-Mejía, M., Ruiz, J., De Freita, M.G., Mora, A., and Mora, A.,
doi:10.1016/j.gsf.2014.01.002 2013, Paleozoic Tectonics and basin evolution along the
Estrada-Carmona, J., Weber, B., Scherer, E.E., Martens, U., and northwestern South American margin. Insights from detrital-
Elías-Herrera, M., in press. Lu-Hf geochronology of zircon U-Pb geochronology. in AAPG Search and Discovery
Mississippian high-pressure metamorphism in the Acatlán Article #90166©2013 AAPG International Conference &
Complex, southern México: Gondwana Research, In Press. Exhibition, Cartagena, Colombia, 8–11 September.
doi:10.1016/j.gr.2015.02.016 Ibañez-Mejía, M., Ruiz, J., Valencia, V.A., Cardona, A., Gehrels, G.E.,
Etayo-Serna, F., Barrero, D., and Others, 1983, Mapa De and Mora, A., 2011, The Putumayo Orogen of Amazonia and
Terrenos Geológicos De Colombia: Bogotá, Ingeominas. its implications for Rodinia reconstructions: New U-Pb geo-
Folk, R.L., 1974, Petrology of sedimentary rocks: Austin, TX, chronological insights into the Proterozoic tectonic evolution
Hemphill Publishing Co., 182p. of northwestern South America: Precambrian Research, v. 191,
Forero, A., 1970a, El Paleozoico Superior del flanco oriental de p. 58–77. doi:10.1016/j.precamres.2011.09.005
la Cordillera Central: Geología Colombiana, v. 7, p. 139–145. Janvier, P., and Villarroel, C., 1998, Los peces Devónicos del
Forero, A., 1970b, Estratigrafía del pre-Cretáceo en el flanco Macizo de Floresta (Boyacá, Colombia). Consideraciones
occidental de la Serranía de Perijá: Geología Colombiana, taxonómicas, bioestratigráficas, biogeográficas y ambien-
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

v. 7, p. 7–77. tales: Geología Colombiana, v. 23, p. 3–18.


Forero-Suarez, A., 1990, The basement of the Eastern Cordillera, Janvier, P., and Villarroel, C., 2000, Devonian vertebrates from
Colombia: An allochthonous terrane in northwestern South Colombia: Palaeontology, v. 43, p. 729–763. doi:10.1111/
America: Journal of South American Earth Sciences, v. 3, p. 1475-4983.00147
141–151. doi:10.1016/0895-9811(90)90026-W Jimenez, D.M., 2000, Catalogo de las unidades
Gehrels, G., Valencia, V., and Pullen, A., 2006, Detrital zircon litoestratigráficas de Colombia, in Filitas y Esquistos de
geochronology by laser- ablation multicollector ICPMS at Busbanzá: Bogotá, Ingeominas, 19 p.
the Arizona LaserChron Center: Paleontological Society Jimenez, G., Speranza, F., Faccenna, C., Bayona, G., and Mora, A.,
Papers, v. 12, p. 67–76. 2015, Magnetic stratigraphy of the Bucaramanga alluvial fan:
Gómez, E., Jordan, T., Allmendinger, R.W., and Cardozo, N., Evidence for a ≤3 mm/yr slip rate for the Bucaramanga-Santa
2005a, Development of the Colombian foreland-basin sys- Marta Fault, Colombia: Journal of South American Earth
tem as a consequence of diachronous exhumation of the Sciences, v. 57, p. 12–22. doi:10.1016/j.jsames.2014.11.001
Northern Andes: Geological Society of America Bulletin, v. Jiménez Mejía, D.M., Juliani, C., and Cordani, U.G., 2006, P-T-t
117, p. 1272–1292. doi:10.1130/B25456.1 conditions of high-grade metamorphic rocks of the Garzon
Gómez, E., Jordan, T., Allmendinger, R.W., Hegarty, K., and massif, Andean basement, SE Colombia: Journal of South
Kelley, S., 2005b, Syntectonic cenozoic sedimentation in American Earth Sciences, v. 21, p. 322–336. doi:10.1016/j.
the Northern Middle Magdalena Valley basin of Colombia jsames.2006.07.001
and implications for exhumation of the Northern Andes: Kammer, A., and Sanchez, J., 2006, Early Jurassic rift struc-
Geological Society of America Bulletin, v. 117, p. 547–569. tures associated with the Soapaga and Boyacá Faults of
doi:10.1130/B25454.1 the Eastern Cordillera, Colombia: Sedimentological infer-
Grodzicki, K.R., Nance, R.D., Keppie, J.D., Dostal, J., and Murphy, ences and regional implications: Journal of South
J.B., 2008, Structural, geochemical and geochronological American Earth Sciences, v. 21, p. 412–422. doi:10.1016/j.
analysis of metasedimentary and metavolcanic rocks of jsames.2006.07.006
the Coatlaco area, Acatlán Complex, southern Mexico: Keppie, J.D., Dostal, J., Murphy, J.B., and Nance, R.D., 2008,
Tectonophysics, v. 461, p. 311–323. doi:10.1016/j. Synthesis and tectonic interpretation of the westernmost
tecto.2008.01.016 Paleozoic Variscan orogen in southern Mexico: From rifted
Henry, D.J., and Dutrow, B.L., 1992, Tourmaline in a low grade Rheic margin to active Pacific margin: Tectonophysics, v.
clastic metasedimentary rock: An example of the petroge- 461, p. 277–290. doi:10.1016/j.tecto.2008.01.012
netic potential of tourmaline: Contributions to Mineralogy Kerr, A.C., Marriner, G.F., Tarney, J., Nivia, A., Saunders, A.D.,
and Petrology, v. 112, p. 203–218. doi:10.1007/BF00310455 Thirlwall, M.F., and Sinton, C.W., 1997, Cretaceous basaltic
Henry, D.J., and Guidotti, C.V., 1985, Tourmaline in the staur- terranes in western Colombia: Elemental chronological and
olite grade metapelites of NW Maine: A petrogenetic indi- Sr-Nd constraints on petrogenesis: Journal of Petrology, v.
cator mineral: American Mineralogist, v. 70, p. 1–15. 38, p. 677–705. doi:10.1093/petroj/38.6.677
Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M., Reyes- Kimberley, M.M., 1980, Paz-de-Rio oolitic inland- sea iron
Harker, A., and Stockli, D.F., 2010, Linking sedimentation in Formation: Economic Geology, v. 75, p. 97–106.
the northern Andes to basement configuration, Mesozoic doi:10.2113/gsecongeo.75.1.97
extension, and Cenozoic shortening: Evidence from detrital Laya, J.C., and Tucker, M.E., 2012. Facies analysis and deposi-
zircon U-Pb ages, Eastern Cordillera, Colombia: Geological tional environments of Permian carbonates in the
Society of America Bulletin, v. 122, p. 1423–1442. Venezuelan Andes. Palaeogeographic implication to
doi:10.1130/B30118.1 Northern Gondwana: Palaeogeography, Palaeoclimatology,
24 A. CARDONA ET AL.

Palaeoecology, v. 332–332, p. 1–26. doi:10.1016/j. extensional fault segmentation and linkage in contractional
palaeo.2012.02.011 orogenesis: A reconstruction of lower cretaceous inverted rift
Ludwig, K.C., 2007, User´s manual for Isoplot 3.7: Berkley, basins in the Eastern Cordillera of Colombia: Basin Research,
Berkley Geochronology Center, 70 p. v. 21, p. 111–137. doi:10.1111/j.1365-2117.2008.00367.x
Mange, A.M., and Maurer, W.H., 1992, Heavy minerals in col- Mora, A., Horton, B.K., Mesa, A., Rubiano, J., Ketchman, R.,
our: Springer, 147p. doi:10.1007/978-94-011-2308-2 Parra, M., Blanco, V., Garcia, D., and Stockli, D., 2010,
Mange, M.A., and Morton, A.C., 2007, Geochemistry of heavy Migration of Cenozoic deformation in the Eastern
minerals, in Mange, M., and Wright, D.T., eds., Heavy miner- Cordillera of Colombia interpreted from fission track results
als in use: Amsterdam, Elsevier, Developments in and structural relationships: Implications for petroleum sys-
Sedimentology, Vol. 58, p. 345–391. tems: AAPG Bulletin, v. 94, p. 1543–1580. doi:10.1306/
Mantilla, L.C., Bissig, T., Cottle, J.M., and Hart, C.J.R., 2012, 01051009111
Remains of early Ordovician mantle-derived magmatism in Morales, P.A., 1965, A contribution to the knowledge
the Santander Massif (Colombian Eastern Cordillera): Devonian faunas of Colombia: Boletín de Geología UIS, v.
Journal of South American Earth Sciences, v. 38, p. 1–12. 19, p. 27–60.
doi:10.1016/j.jsames.2012.03.001 Moreno-López, M.C., and Escalona, A., 2015, Precambrian–
Mantilla, L.C., Bissig, T., Valencia, V., and Hart, C.J., 2013, The Pleistocene tectono-stratigraphic evolution of the southern
magmatic history of the Vetas California mining district; Llanos basin, Colombia: AAPG Bulletin, v. 99, p. 1473–1501.
Santander massif, eastern Cordillera, Colombia: Journal of doi:10.1306/11111413138
South American Earth Sciences, v. 45, p. 235–249. Moreno-Sanchez, M., 2004. Devonian plants from Colombia:
doi:10.1016/j.jsames.2013.03.006 Geologic framework and paleogeographic implications [Ph.
Mantilla-Figueroa, L.C., García-Ramirez, C.A., and Valencia, V., D. Thesis]. Liége, Université de Liége, p. 183.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

2016, Propuesta de escisión de la denominada ‘Formación Moreno-Sanchez, M., Gomez, A., and Castillo, H., 2007, La
Silgará’ (Macizo de Santander, Colombia), a partir de edades “Formación Floresta metamorfoseada” (sensu Ward.,
U-Pb en circones detríticos: Boletín De Geología, v. 38, p. 1973) no es la formación floresta sin metamorfosear, in
33–50. doi:10.18273/revbol.v38n1-2016002 X Congreso Colombiano de Geología: Paipa, Resúmenes,
Martens, U., Restrepo, J.J., Ordoñez-Carmona, O., and Correa- 7 p.
Martinez, A., 2014, The Tahamí and Anacona terranes of the Nance, R.D., Gutiérrez-Alonso, G., Keppie, J.D., Linnemann, U.,
Colombian Andes: Missing links between the South Murphy, J.B., Quesada, C., Strachan, R.A., and Woodcock, N.
American and Mexican Gondwanan margins: The Journal H., 2010, Evolution of the Rheic Ocean: Gondwana Research,
of Geology, v. 122, p. 507–530. doi:10.1086/677177 v. 17, p. 194–222. doi:10.1016/j.gr.2009.08.001
Martens, U., Weber, B., and Valencia, V.A., 2010, U/Pb geochro- Nance, R.D., Keppie, J.D., Miller, B.V., Murphy, J.B., and Dostal, J.,
nology of Devonian and older Paleozoic beds in the south- 2009, Palaeozoic palaeogeography of Mexico: Constraints
eastern Maya block, Central America: Its affinity with peri- from detrital zircon age data: Geological Society, London,
Gondwanan terranes: Geological Society of America Special Publications, v. 327, p. 239–269. doi:10.1144/SP327.12
Bulletin, v. 122, p. 815–829. doi:10.1130/B26405.1 Nie, J., Horton, B.K., Mora, A., Saylor, J.E., Housh, T.B., Rubiano,
Miller, H., and Söllner, F., 2005, The famatina complex (NW- J., and Naranjo, J., 2010, Tracking exhumation of Andean
Argentina): Back-docking of an island arc or terrane accre- ranges bounding the Middle Magdalena Valley basin,
tion? Early Palaeozoic geodynamics at the western Colombia: Geology, v. 38, p. 451–454. doi:10.1130/G30775.1
Gondwana margin, in Vaughan, A.P.M., Leat, P.T., and Ordóñez-Carmona, O., Restrepo Álvarez, J.J., and Pimentel, M.
Pankhurst, R.J., eds., Terrane processes at the margins of M., 2006, Geochronological and isotopical review of pre-
Gondwana: Geological Society of London, Special Devonian crustal basement of the Colombian Andes:
Publications, v. 246, p. 241–256. doi:10.1144/GSL. Journal of South American Earth Sciences, v. 21, p. 372–
SP.2005.246.01.08 382. doi:10.1016/j.jsames.2006.07.005
Miskovic, A., Schaltegger, U., Spikings, R. A., Chew, D.M., Košler, Ortega-Obregón, C., Solari L.A., Keppie, J.D., Ortega-Gutiérrez,
J., 2009. Tectono-magmatic evolution of Western Amazonia: F., Solé, J., and Morán-Ical, S., 2008, Middle – Late
geochemical characterization and zircon U-Pb geochrono- Ordovician magmatism and Late Cretaceous collision in
logic constraints from the Peruvian Eastern Cordilleran the southern Maya block, Rabinal - Salamá area, central
granitoids. Geological Society of America Bulletin, v. 121, Guatemala: implications for North America-Caribbean
p.1289–1324. plate tectonics: Geological Society of America Bulletin, v.
Mišković, A., Spikings, R.A., Chew, D.M., Košler, J., Ulianov, A., 120, p. 556–570. doi:10.1130/B26238.1
and Schaltegger, U., 2009, Tectonomagmatic evolution of Paces, J., and Miller, J., 1993, Precise U-Pb ages of Duluth
Western Amazonia: Geochemical characterization and zircon complex and related mafic intrusions, northeastern
U–Pb geochronologic constraints from the Peruvian Eastern Minnesota; geochronological insights to physical, petroge-
Cordilleran granitoids: Geological Society of America netic, paleomagnetic, and tectonomagmatic processes
Bulletin, v. 121, p. 1298–1324. doi:10.1130/B26488.1 associated with the 1.1 Ga midcontinent rift system:
Mojica, J., and Villarroel, C., 1984, Contribución al conoci- Journal of Geophysical Research, v. 98, no. B8, p. 13997–
miento de las unidades paleozoicas del área de Floresta 14013. doi:10.1029/93JB01159
(Cordillera Oriental Colombiana, Departamento de Boyacá) Pankhurst, R.J., Rapela, C.W., Saavedra, J., Baldo, E.G.,
y en especial de la Formación Cuche: Geología Colombiana, Dahlquist, J.A., Pascua, I., and Fanning, C.M., 1998, The
v. 13, p. 55–80. Famatinian arc in the central Sierras Pampeanas: An early
Mora, A., Gaona, T., Kley, J., Montoya, D., Parra, M., Quiroz, L., to mid-ordovician continental arc on the Gondwana mar-
Reyes, G., and Strecker, M.R., 2008a, the role of inherited gin, in Pankhurst, R.J., and Rapela, C.W., eds., The proto-
INTERNATIONAL GEOLOGY REVIEW 25

andean margin of Gondwana: Geological Society of Romero, D., Valencia, K., Alarcón, P., Peña, D., and Ramos, V.A.,
London, Special Publications, v. 142, p. 343–367. 2013, The offshore basement of Perú: Evidence for different
doi:10.1144/GSL.SP.1998.142.01.17 igneous and metamorphic domains in the forearc: Journal
Parra, M., Mora, A., Jaramillo, C., Strecker, M.R., Sobel, E.R., of South American Earth Sciences, v. 42, p. 47–60.
Quiroz, L.I., Rueda, M., and Torres, V., 2009, Orogenic doi:10.1016/j.jsames.2012.11.003
wedge advance in the Northern Andes: Evidence from the Royero, J.M., 2001, Memoria explicativa: Geología y
oligocene-miocene sedimentary record of the Medina geoquímica de la plancha 111 Toledo – Norte de
basin, Eastern Cordillera, Colombia: Geological Society of Santander: Bucaramanga, Ingeominas, 92 p.
America Bulletin, v. 121, p. 780–800. doi:10.1130/B26257.1 Royo, Y., and Gomez, J., 1942, Fósiles Devónicos de Floresta
Pindell, J., and Dewey, J.F., 1982, Permo-triassic reconstruction (Departamento de Boyacá)in Compilación de los Estudios
of western Pangea and the evolution of the Gulf of Mexico/ Geológicos Oficiales en Colombia: Bogotá, Ministerio de Minas
Caribbean region: Tectonics, v. 1, no. 2, p. 179–211. y Petróleos, Servicio Geológico Nacional, v. 5, p. 389–396.
doi:10.1029/TC001i002p00179 Santos, C., Jaramillo, C., Bayona, G., Rueda, M., and Torres, V.,
Pindell, J.L., Cande, S., Pitman III, W.C., Rowley, D.B., Dewey, J. 2008, Late eocene marine incursion in Northwestern South
F., Labrecque, J., and Haxby, W., 1988, A plate-kinematic America: Palaeogeography, Palaeoclimatology, Palaeoecology,
framework for models of Caribbean evolution: v. 264, p. 140–146. doi:10.1016/j.palaeo.2008.04.010
Tectonophysics, v. 155, p. 121–138. doi:10.1016/0040-1951 Saylor, J.E., Stockli, D., Horton, B., Nie, J., and Mora, A., 2012,
(88)90262-4 Discriminating rapid exhumation from syndepositional vol-
Radelli, L., 1967, Geologic des Andes Colombiennes: Travaux canism using detrital zircon double dating: Implications for
de laboratoire de geologic de la faculte des sciences de the tectonic history of the Eastern Cordillera, Colombia:
Grenoble: Memoirs, v. 6, p. 464. Geological Society of America Bulletin, v. 124, p. 762–779.
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

Ramos, V.A., 2009, Anatomy and global context of the Andes: doi:10.1130/B30534.1
Main geologic features and the Andean orogenic cycle: Shao, W.-Y., Chung, S.-L., Chen, W.S., Lee, H.-Y., and Xie, L.-W.,
Backbone of the Americas: Shallow subduction, plateau 2015, Old continental zircons from a young oceanic arc,
uplift, and ridge and terrane collision: Geological Society eastern Taiwan: Implications for Luzon subduction initiation
of America Memoirs, v. 204, p. 31–65. doi:10.1130/ and Asian accretionary orogeny: Geology, v. 43, p. 479–482.
2009.1204(02) doi:10.1130/G36499.1
Ramos, V.A., and Aleman, A., 2000, Tectonic evolution of the Silva, J.C., Sial, A.N., Ferreira, V.P., and Estrada, J.J., 2004, C-
Andes, in Cordani, U.J., Milani, E.J., Thomaz Filho, A., and isotope stratigraphy of a Vendian carbonate succession in
Campos, D.A., eds., Tectonic, evolution of South America: northwestern Andes: Implications for the NW Andes, in IV
Río de Janeiro, IGC, p. 635–685. Reunión Ciencias de la Tierra: Querétaro, México, Abstracts,
Restrepo, J.J., Ordoñez-Carmona, O., Armstrong, R., and Vol. 198.
Pimentel, M.M., 2011, Triassic metamorphism in the north- Sláma, J., Kosler, J., Condon, D.J., Crowley, J.L., Gerdes, A.,
ern part of the Tahamí Terrane of the central cordillera of Hanchar, J.M., Horstwood, M.S.A., Morris, G.A., Nasdala, L.,
Colombia: Journal of South American Earth Sciences, v. 32, Norberg, N., Schaltegger, U., Schoene, B., Tubrett, M.N., and
p. 497–507. doi:10.1016/j.jsames.2011.04.009 Whitehouse, M.J., 2008, Plesovice zircon - a new natural
Restrepo, J.J., and Toussaint, J.F., 1988, Terranes and conti- reference material for U-Pb and Hf isotopic microanalysis:
nental accretion in the Colombian Andes: Episodes, v. 7, p. Chemical Geology, v. 249, no. 1–2, p. 1–35. doi:10.1016/j.
189–193. chemgeo.2007.11.005
Restrepo-Pace, P.A., and Cediel, F., 2010, Northern South Stibane, F., and Forero, A., 1969, Los afloramientos del
America basement tectonics and implications for paleocon- Paleozoico en la Jagua (Huila) y Río Nevado (Santander):
tinental reconstructions of the Americas: Journal of South Geología Colombiana, v. 6, p. 31–66.
American Earth Sciences, v. 29, p. 764–771. doi:10.1016/j. Suárez Díaz, H.G., and Solano, Y.P., 2012, El Paleozoico en los
jsames.2010.06.002 Llanos Orientales de Colombia: Una nueva ventana en la
Restrepo-Pace, P.A., Ruiz, J., Gehrels, G., and Cosca, M., 1997, búsqueda de fuentes de hidrocarburos: Revista de la
Geochronology and Nd isotopic data of Grenville-age Asociación Colombiana de Geólogos y Geofisicos del
rocks in the Colombian Andes: New constraints for Late Petróleo, v. 14, p. 8–11.
Proterozoic–Early Paleozoic paleo continental reconstruc- Tassinari, C., and Macambira, M., 1999, Geochronological
tions of the Americas: Earth and Planetary Science provinces of the Amazonian Craton: Episodes, v. 22, p.
Letters, v. 150, p. 427–441. doi:10.1016/S0012-821X(97) 174–182.
00091-5 Torsvik, T.H., and Cocks, R.L., 2013, Gondwana from top to
Ríos, C., García, C., and Takasu, A., 2003, Tectono-metamorphic base in space and time: Gondwana Research, v. 24, no. 3–
evolution of the Silgará Formation metamorphic rocks in 4, p. 999–1030. doi:10.1016/j.gr.2013.06.012
the southwestern Santander Massif, Colombian Andes: Toussaint, J.F., 1993, Evolución geológica de Colombia,
Journal of South American Earth Sciences, v. 16, p. 133– Precámbrico–Paleozóico: Medellín, Universidad Nacional
154. doi:10.1016/S0895-9811(03)00025-7 de Colombia, 129 p.
Rios, C.A., Castellanos, O.M., Gomez, S.I.E., and Avila, G.A., 2008, Trumpy, D., 1943, Pre-Cretaceous of Colombia: Geological
Petrogenesis of the metacarbonate and related rocks of the Society of America Bulletin, v. 54, p. 1281–1304.
Silgará Formation, Central Santander Massif, Colombian doi:10.1130/GSAB-54-1281
Andes: An overview of a “Reaction Calcic Exoscarn”: Earth Ulloa, C., Guerra, A., and Escovar, R., 1998, Geología Plancha
Sciences Researcher Journal, v. 12, no. 1, p. 72–106. 172 - Paz de Río. Escala 1:100.000: Bogotá, Ingeominas.
26 A. CARDONA ET AL.

Van der Lelij, R., Spikings, R., Ulianov, A., Chiaradia, M., and Mora, American Earth Sciences, v. 21, p. 355–371. doi:10.1016/j.
A., 2016, Palaeozoic to Early Jurassic history of the north- jsames.2006.07.007
western corner of Gondwana, and implications for the evolu- Ward, D.E., Goldsmith, R., Jaime, B., and Restrepo, H.A., 1974,
tion of the Iapetus, Rheic and Pacific Oceans: Gondwana Geology of quadrangles H-12, H-13, and parts of I-12 and
Research, v. 31, p. 271–294. doi:10.1016/j.gr.2015.01.011 I-13, (zone III) in northeastern Santander Department,
Veloza, G., Styron, R., Taylor, M., and Mora, A., 2013, Open Colombia: U.S. Geological Survey.
Soure archive of active faults for northwestern South Weber, B., Valencia, V.A., Schaaf, P., Pompa-Mera, V., and
America: GSA Today, v. 22, no. 10. doi:GSATG156A/ Ruiz, J., 2008, Significance of provenance ages from the
GSATG156A.1 Chiapas Massif complex (SE México): Redefining the
Villagómez, D., Spikings, R., Magna, T., Kammer, A., Winkler, W., Paleozoic basement of the Maya block and its evolution
and Beltrán, A., 2011, Geochronology, geochemistry and in a peri-Gondwanan realm: The Journal of Geology, v.
tectonic evolution of the Western and Central cordilleras 116, p. 619–639. doi:10.1086/591994
of Colombia: Lithos, v. 125, p. 875–896. doi:10.1016/j. Yavuz, F., 1997. TOURMAL: Software package for tourmaline,
lithos.2011.05.003 tourmaline-rich rocks and related ore deposits: Computers
Vinasco, C.J., Cordani, U.G., González, H., Weber, M., and and Geosciences, v. 23, p. 947–959.
Pelaez, C., 2006, Geochronological, isotopic and geochem- Young, G.C., and Moody, J.M., 2002, A Middle-late devonian fish
ical data from Permo-Triassic granitic gneisses and grani- fauna from the Sierra de Perijá, western Venezuela, South
toids of the Colombian Central Andes: Journal of South America: Fossil Record, v. 5, p. 155–206. doi:10.5194/fr-5-155-
2002
Downloaded by [Agustin Cardona] at 07:48 06 June 2016

View publication stats

You might also like