You are on page 1of 12

Journal of South American Earth Sciences xxx (xxxx) xxx

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


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

Jambaló blueschist and greenschist protoliths in the Central Cordillera of


the Colombian Andes and their tectonic implications for Late Cretaceous
Caribbean-South American interactions
Andres Bustamante a, *, Camilo Bustamante b, Agustín Cardona c, Caetano Juliani d,
Salviano Pereira da Silva a
a
Departamento de Geologia, Universidade Federal de Pernambuco, Av. da Arquitetura s/nº, CEP 50740-550, Recife, PE, Brazil
b
Departamento de Ciencias de la Tierra, Universidad EAFIT, Carrera 49 N◦ 7 Sur-50, Medellín, Colombia
c
Departamento de Procesos y Energía, Universidad Nacional de Colombia, Carrera 80 # 65-223, Medellín, Colombia
d
Instituto de Geociências, Universidade de São Paulo, Rua do Lago 562, CEP 05508-080, São Paulo, SP, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: Whole-rock geochemistry and Nd isotopes were used to constrain the nature of the protoliths and provide clues
Jambaló blueschists on the metamorphic conditions of the Jambaló blueschists in the Central Cordillera (Colombian Andes).
High-pressure metamorphism Blueschist-facies rocks and related greenschists belong to a Cretaceous–Paleocene subduction complex. The
Northern Andes
blueschist occurs as a lens surrounded by greenschist-facies rocks, suggesting that the latter resembles retrograde
Caribbean tectonics
products of high-pressure rocks. The geochemical composition of blueschists and greenschists indicates similar
protoliths, implying that the Jambaló rocks were probably part of a unique and coherent block exposed to
different degrees of retrograde metamorphism. Th/La ratios above 0.19 and 143Nd/144Nd compositions of
0.51272 and 0.51259 indicate subduction zone sedimentary input or some crustal contamination, interpreted as
the interaction of arc-derived sediments with basalts formed in a supra-subduction zone. The Jambaló schists
may represent the youngest exposure of high-pressure metamorphic rocks along the Andes and records of Late
Cretaceous Caribbean and South American convergence.

1. Introduction coexistence in the transition zone between both rock facies (Maruyama
et al., 1986; Evans, 1990; Baziotis et al., 2009). In contrast, other
The blueschist-facies rocks are one of the significant relics of sub­ high-pressure terrains spatially related to greenschist-facies rocks are
duction zones. Their exhumation may reflect serpentinization of the slab characterized by medium-to low-pressure metamorphic overprinting
mantle that enhances the mechanical coupling, facilitating the relationships (Yokohama et al., 1986; Bröcker, 1990).
buoyancy-driven exhumation with associated erosion. Also, they may High-pressure metamorphic records in the Northern Andes include
document changes in the rate of underthrusting, slab rollback, or very Mesozoic blueschists, eclogites, and amphibolite-facies rocks, which
oblique convergence (Ernst, 1988; Maruyama et al., 1996; Agard et al., appear as discontinuous exposures in Ecuador, Venezuela, and
2009; Guillot et al., 2009). Colombia. They include the Mesozoic Raspas Metamorphic Complex in
In several localities where blueschist-facies rocks occur, they are southwestern Ecuador and carry signatures of OIB and mid-ocean ridge
usually found associated with greenschists-facies rocks, and their basalt (MORB) (Arculus et al., 1999; Bosch et al., 2002; John et al.,
petrogenetic relations were commonly the object of significant discus­ 2010). The Villa de Cura blueschist belt is an E-W trend belt that in­
sions (Dungan et al., 1983; Brown, 1986; Oberhänsli, 1986; Bröcker, cludes the Caribbean Mountain system of northern Venezuela, charac­
1990; Bustamante, 2008; Baziotis et al., 2009). Examples include the terized by arc-related protoliths (Smith et al., 1999). Finally, in the
presence of lenticular blueschists within large bodies of Central Cordillera of the Colombian Andes, the high-pressure meta­
greenschist-facies rocks or as alternating layers. Their association is morphic records are included as part of the single Arquía Complex
related to bulk compositional variations that can represent their (Feininger, 1980; McCourt and Feininger, 1984; Aspden and McCourt,

* Corresponding author.
E-mail address: andresbl@aim.com (A. Bustamante).

https://doi.org/10.1016/j.jsames.2020.102977
Received 7 March 2020; Received in revised form 9 September 2020; Accepted 19 October 2020
Available online 12 November 2020
0895-9811/© 2020 Elsevier Ltd. All rights reserved.

Please cite this article as: Andres Bustamante, Journal of South American Earth Sciences, https://doi.org/10.1016/j.jsames.2020.102977
A. Bustamante et al. Journal of South American Earth Sciences xxx (xxxx) xxx

1986; Bourgois et al., 1987; Aspden et al., 1995; Maya and González, representing an Early Cretaceous subduction/accretion complex. In
1995; Bustamante, 2008), but with at least two different Early and Late contrast, the Ar–Ar age of approximately 62 Ma in Jambaló blueschists
Cretaceous units. The approximately 120-Ma (Ar–Ar in muscovite) (Bustamante et al., 2011) and the different supra-subduction settings
blueschists with features resembling normal mid-ocean ridge basalt (Bustamante and Bustamante, 2019) from geochemical evaluations of
(N-MORB), along with retrograded eclogites of the Barragán region blueschist samples indicate various metamorphic records.
(Bustamante et al., 2012), may have been formed by the subduction of This study introduces the whole-rock geochemistry and petrography
the Farallon Plate beneath the northwestern margin of the South of less-studied greenschist-facies rocks associated with blueschist-facies
American Plate (Bustamante et al., 2011, 2012; Bustamante and Bus­ rocks in the Jambaló region. These results, together with the blueschist
tamante, 2019). The second occurrence includes blueschists of approx­ geochemistry previously studied by Bustamante and Bustamante (2019)
imately 62 Ma (Ar–Ar in paragonite) in the Jambaló region, whose and the Nd isotopes from blueschists, shed light on the nature of pro­
protolith was formed in a plume-influenced intra-oceanic arc. It records toliths and tectonic implications of the occurrence of the youngest
interactions between the Caribbean oceanic crust and the Northern blueschists in the Andes.
Andes (Bustamante and Bustamante, 2019).
Controversy about the origin of protoliths and the meaning of 2. Geological setting
Jambaló blueschists remains in the incomplete geochemical database
due to the lack of samples available from this locality. One model pro­ The Colombian Andes consists of three mountain ranges or Cordil­
poses that both the Jambaló and Barragán high-pressure rocks can be leras, separated by two major rivers (Cauca and Magdalena) that flow
correlated with those of the Raspas Metamorphic belt in Ecuador (Fei­ into the Caribbean (Fig. 1). The Western Cordillera is characterized by
ninger, 1982; Villagómez et al., 2011; Spikings et al., 2015, 2019). This Cretaceous allochthonous volcanic and plutonic rocks formed in intra-
indicates fragments of a contemporaneous metamorphic belt oceanic plateau and arc settings (Kerr et al., 1997; Villagómez et al.,

Fig. 1. (A) Metamorphic units of the Central Cordillera of the Colombian Andes were measured according to the methods of Maya and González (1995). Lithological
relationships between high-pressure rocks from the Jambaló region and other metamorphic units were modified from Maya (2001) and Maya and Vásquez (2001).
On the map, the high variability range of the eras is due to the lack of coherent and systematic geochronological data available from the Central Cordillera of
Colombia, (B) Schematic map (right) of the Jambaló region showing geological/structural relationships with other units in the area, after Maya (2001).

2
A. Bustamante et al. Journal of South American Earth Sciences xxx (xxxx) xxx

2011). The Central Cordillera, which includes Permo-Triassic syn-tec­ These lithologies exist as lenses of several meters in size and are sur­
tonic granitoids and amphibolites to the east (Vinasco et al., 2006; Vil­ rounded by a “matrix” of greenschist.
lagómez et al., 2011), are exposed primarily as roof pendants within Marbles exist north of Jambaló and are composed of calcite and
Jurassic batholiths (Cochrane et al., 2014; Bustamante et al., 2016). The possibly aragonite, and to a lesser extent, dolomite. Accessory minerals
Otú-Pericos Fault separates these rocks from western Triassic to Jurassic are epidote and/or clinozoisite, quartz, talc, apatite, chlorite, opaque
greenschist, amphibolite, and quartzofeldspathic gneiss belts (Restrepo minerals, rutile, and titanite. These rocks are predominantly grano­
et al., 2011; Blanco-Quintero et al., 2014; Bustamante et al., 2017). blastic with carbonate composition and show lepido-nematoblastic and
These are limited to Early to Late Cretaceous volcano-sedimentary rocks nemato-lepidoblastic textures in areas where talc and epidote-
of the Quebradagrande Complex by the San Jerónimo Fault (Toussaint, clinozoisite are concentrated.
1996; Nivia et al., 2006; Villagómez et al., 2011; Cochrane et al., 2014). Metamarls are present near the town of Jambaló and are essentially
The Cretaceous, which mainly consists of amphibolite-facies rocks, are composed of epidote and/or clinozoisite (50–63%), carbonates (up to
grouped into the Arquía Complex (Maya and González, 1995). They are 11%), opaque minerals (up to 20%), amphibole (up to 10%), plagioclase
exposed as a series of discontinuous belts in the western flank of the (10%), chlorite (7%), quartz (3%), pumpellyite (~1–5%), mica (2%),
Central Cordillera, with the Silvia-Pijao Fault limiting them by the Late and zircon (accessory). This lithology has a granofelsic structure and a
Cretaceous volcano-sedimentary rocks of the Quebradagrande Complex granoblastic texture.
(Fig. 1). The Eastern Cordillera is composed of Precambrian to Paleozoic Quartzites have limited occurrences north of Jambaló, near the San
amphibolite-facies and plutonic rocks and is covered with thick Paleo­ Francisco region. Accessory phases are white mica (16–25%) and opa­
zoic to Cenozoic sedimentary deposits (Horton et al., 2010; Sarmien­ que minerals (trace to 4%). They also present traces (~1%) of plagio­
to-Rojas et al., 2006). clase, clinozoisite, chlorite, glaucophane, zoisite, stilpnomelane, and
garnet. Texturally, they exhibit mylonitic foliation with lepidoblastic
2.1. High-pressure metamorphic rocks in the Central Cordillera of and lepido-granoblastic habits. Structurally, it is possible to identify a
Colombia wide range of foliation (S2) created by the orientation of quartz and
mica. In addition, one can see crenulations (pos-S2) formed primarily by
As mentioned earlier, high-pressure metamorphic rocks in the the orientation of white mica crystals. S1 is identified only in garnets and
Arquía Complex are made up of several types of rocks (Fig. 1) such as exhibits a pattern of quartz inclusions. An opaque mineral with a
amphibolite, garnet amphibolite, greenschist, amphibole schist, and different orientation from that observed in S2 and pos-S2.
ultramafic rocks (Maya and González, 1995). They also include Keratophyres can only be found near the Toribío area, north of
low-grade metamorphic rocks such as quartz-muscovite schist, actinolite Jambaló. They are composed of plagioclase An5 (>90%) and traces of
schist, and quartzite (Orrego et al., 1993; Maya and González, 1995). amphibole, stilpnomelane, opaque minerals, chlorite, epidote, and
The age of this complex is still a matter of debate at Early Cretaceous to quartz. Texturally, the rocks exhibit evidence of recrystallization,
Paleogene ages, certainly reflecting the complex sequence of events plagioclase subgrains, and folded twinning. However, this lithotype
experienced by the continental margin during the Cretaceous (Busta­ lacks metamorphic foliation.
mante et al., 2011; Bustamante and Bustamante, 2019; Avellaneda-­ Serpentinites are always associated with the occurrence of blues­
Jiménez et al., 2019). These rocks are also mixed with Pre-Cretaceous chists, especially around Jambaló. Accessory phases include opaque
metamorphic and igneous remnants (Zapata et al., 2019), which may minerals, chlorite, amphibole (possibly actinolite), and olivine and py­
represent fragments of the Pre-Cretaceous continental margin forearc roxene relics, presenting a mesh texture.
basin basement.
Two well-defined blueschist occurrences have been recognized in the 3. Analytical methods
western part of the Central Cordillera in the Barragán and Jambaló re­
gions (Bustamante et al., 2011, 2012) (Fig. 1). These two occurrences 3.1. Whole-rock geochemistry
have proven to be different in age and protolith content. The Barragán
rocks are characterized by epidote-glaucophane and chlorite-lawsonite Sample types and mineral contents are presented in Table 1. Based
schists, exhibiting distinct MORB-type protoliths and metamorphic on the petrographic analysis, ten greenschist and eight blueschist sam­
Ar–Ar ages obtained from white mica in the associated metapelites of ples were chosen for whole-rock geochemistry, avoiding veins and
approximately 120 Ma (Bustamante et al., 2012). On the other hand, weathered surfaces before crushing with a jaw crusher. We analyzed the
Jambaló blueschists with an Ar–Ar metamorphic age of about 62 Ma whole-rock geochemistry of 18 representative samples to constrain the
have been interpreted as plume-influenced intra-oceanic arc-like pro­ tectonic setting of protoliths. The results are shown in Table 2 and
toliths (Bustamante et al., 2011; Bustamante and Bustamante, 2019). include whole-rock geochemistry reported for blueschists in Bustamante
Fig. 1 shows the geological and stratigraphic relationships of and Bustamante (2019) and new geochemical data for greenschists.
Jambaló schists. Jambaló blueschists are exposed in structural contacts Rock powders (90% < 200 mesh) were prepared using a tungsten
with different pelitic and amphibole-bearing schists and granitoids in carbide ring mill at the Institute of Geosciences, University of São Paulo,
the Jurassic to Triassic Cajamarca Complex to the east, as multiple lenses Brazil. The analytical procedure included acid digestion of 0.25 g sam­
in a body of elongated greenschist covering an area of approximately 25 ple. Although tungsten carbide is a potential contaminant (Nb and Co),
km2 (e.g., Bustamante, 2008; Bustamante et al., 2017). To the west, they experiments performed by Johnson et al. (1999) suggest that the extent
are in contact with a variety of volcanic and gabbroic bodies associated of contamination is on the order of magnitude (2%) of the detection
with the Cretaceous Quebradagrande Complex (Maya and González, limit (one standard deviation equals 1.0 ppm). As with many other
1995). However, the complex structural relationships and lack of scientists, Yamasaki (2018) reported that major elemental analysis
geochemical constraints of the latter unit hinder its real characteristics, showed no evidence of sample contamination using different types of
as it may also be correlated with other oceanic units formed on its mills. Trace element analysis was affected by W and Co contamination
western margin. In this study, we only focused on greenschists and from tungsten carbide ring mills, while other trends were easily
blueschists. discernible.
Samples were analyzed using an Inductively Coupled Plasma-
2.1.1. The Jambaló schists Emission Spectrometer (ICP-ES) for major elements and an Inductively
Field observations and sampling were performed near the town of Coupled Plasma-Mass Spectrometer (ICP-MS) for trace and Rare Earth
Jambaló. The region includes blueschists with various marbles, meta­ Elements (REE) at the Activation Laboratories Ltd. (Actlabs), Ancaster,
marls, quartzites, keratophyres (albitized dacites), and serpentinites. Ontario, Canada. The detection limit for SiO2, Al2O3, Fe2O3, MgO, CaO,

3
A. Bustamante et al. Journal of South American Earth Sciences xxx (xxxx) xxx

Na2O, K2O, and P2O5 was 0.01%, MnO and TiO2 was 0.001%, and the

129A
loss on ignition (LOI) was 0.01%. For trace elements, the detection limits
BS

+
+

+
+
+

+
+

+
varied between 0.005 and 10.000 ppm depending on the element
analyzed. The results of the geochemical analysis were processed using
GCDKit4.1 software (Janoušek et al., 2006).
125M
BS

+
+

+
+
+
3.2. Whole-rock Sm and Nd isotopes
125K

Two blueschist and one greenschist samples were analyzed at the


BS

+
+
+
+
+

+
+
+
+
+
Geochronology Laboratory, University of Brasília, Brazil. Sm and Nd
isotopic compositions were analyzed by Thermal Ionization Mass
Spectrometry (TIMS) using a Finnigan MAT 262 mass spectrometer
125I
BS

+
+

+
+

+
+
+

+
+

+
+
(provided with seven collectors) in static mode. Prior to isotopic mass
spectrometer analysis, the chemical extraction of Sm and Nd was con­
ducted using the conventional cation-exchange technique with
124J

approximately 100 mg of the whole-rock powder. A detailed description


BS

+
+
+
+
+

+
+
+
of the analytical method can be found in Gioia and Pimentel (2000).
124G

3.3. Mineral analysis


BS

+
+
+
+
+

+
+
+

+
+
+

X-ray mapping was performed using a CAMECA SX-100 Electron


Microprobe at the Centro de Instrumentación Científica, Universidad de
123A
BS

+
+

+
+

+
+
+
+
+

Granada, Spain. Elemental X-ray maps of Al, Ba, Ca, Cl, Fe, K, Mg, Mn,
Na, S, Si, Ti, and Zr were obtained for sample #118 under the following
operating conditions: 15 keV and 300 nA, focused beam, 4–8 μm step
121B

(pixel) size depending on grain size, and 30 ms/pixel count time. Images
BS

+
+

+
+

+
+
+

+
+
+

+
+

were processed using XMapTools software (Lanari et al., 2014, 2019).


128A

4. Results
GS

+
+

4.1. Petrography
127B
GS

Stratigraphic relationships are rarely exposed and faulted contacts


+
+
+
+

+
+

between geological units are common. However, on a regional scale,


blueschists are embedded as lenses in the greenschist-facies matrix.
127A
GS

+
+

4.1.1. Blueschists
Blueschists are fine-to medium-grained rocks characterized by
Rock-forming minerals of the used samples in this study. Greenschist (GS) and blueschist (BS).

124C

metamorphic foliation. The prograde metamorphic mineral assemblage


GS

+
+
+

consists of glaucophane (30–50%) + white mica (including paragonite


and minor phengite) (10–36%) + quartz (5–12%) + epidote/clinozoisite
(1–30%) ± chlorite (1–12%) ± carbonate (up to 3%). Accessory and
121C
GS

retrograded minerals include albite, opaque minerals (including pyrite,


+
+
+
+

+
+

ilmenite, and magnetite), titanite, zircon, garnet, rutile, apatite, and


stilpnomelane. Glaucophane occurs with a partial replacement on the
114
GS

rims of barroisite and actinolite (Fig. 2A), and on the cleavage planes,
+
+
+
+
+
+

fractures, and rims of chlorite (Fig. 2B). Inclusions of epidote and/or


clinozoisite, opaque minerals, zircon, and rarely rutile are present.
110
GS

Although glaucophane orientation defines the main foliation (S2), it also


+
+
+
+

exists in a decussate arrangement (Fig. 2C). White mica also defines the
major foliation (Fig. 2D). Quartz is typically scattered throughout the
107B

sample and occurs as inclusions in carbonates, chlorite, and white mica,


GS

+
+
+
+
+
+

or as a pressure shadow around glaucophane. Epidote and clinozoisite


are also common. They are predominantly xenoblastic, and in some
107A

cases, subidioblastic. Some idoblastic crystals are rarely disseminated


GS

+
+
+
+

+
+

throughout the rock samples. Carbonates exist as xenoblastic grains or as


inclusions in glaucophane. In samples containing garnet, amphiboles are
gulf-shaped, have boundaries of corrosion, and may be associated with
107
GS

+
+

+
+

pressure shadows.

4.1.2. Greenschists
Epidote-clinozoiste

Greenschists are also fine-to medium-grained rocks characterized by


Stilpnomelane
Glaucophane

metamorphic foliation, which are locally folded. They include albite


Pumpellyite
White mica
Amphibole

Magnetite
Rock type

Sample #

(44–68%), chlorite (8–35%), quartz (up to 17%), white mica (2–15%),


Ilmenite
Chlorite

Apatite
Sphene
Calcite
Quartz
Table 1

Garnet
Zoisite

Zircon
Albite

Rutile
Pyrite

opaque minerals (up to 13%), epidote and/or clinozoisite (1–8%),


zoisite (up to 8%), biotite (up to 4%), titanite (1–3%), carbonates (up to

4
A. Bustamante et al.
Table 2
Major and trace elements of the Jambaló greenschist (GS) and blueschist (BS). *Analyses after Bustamante and Bustamante (2019).
Rock type Unit Symbol Detection Limit Std GS GS GS GS GS GS GS GS GS GS BS* BS* BS* BS* BS* BS* BS* BS*

Sample # W-2a 107 107A 107B 110 114 121C 124C 127A 127B 128A 121B 123A 124G 124J 125I 125K 125M 129A

SiO2 % 0.01 52.58 59.27 49.06 58.12 48.53 55.16 50.94 49.53 48.4 49.77 49.37 51.74 50.42 56.9 52.01 53.28 53.06 56.33 51.42
Al2O3 % 0.01 15.34 15.58 16.55 16.99 16.02 16.48 17.67 15.25 16.94 18.59 16.38 15.59 17.76 15.62 18.06 17.15 19.77 17.77 14.41
Fe2O3 % 0.01 10.72 7.08 11.52 7.78 11.45 9.12 6.96 11.21 11.39 9.43 10.84 9.66 9.75 9.5 8.84 9.12 10.87 8.14 9.53
MnO % 0.001 0.168 0.156 0.18 0.096 0.16 0.174 0.118 0.16 0.16 0.09 0.16 0.441 0.139 0.123 0.145 0.24 0.08 0.103 0.129
MgO % 0.01 6.37 3.18 7.29 3.31 6.65 5.91 6.13 6.8 8.25 5.39 5.33 3.49 3.27 4.02 4.48 5.41 2.59 4.68 5.33
CaO % 0.01 10.92 2.99 6.65 4.7 7 4.6 8.87 7.84 4.31 6.57 7.59 5.76 11.6 2.94 3.48 3.67 1.74 1.64 3.95
Na2O % 0.01 2.22 6.67 3.86 3.56 2.79 1.86 2.33 4.48 2.25 4.85 3.91 4.78 2.62 4.49 5.57 5.07 3.36 5.65 5.18
K2O % 0.01 0.63 1.59 0.51 1.19 2.42 1.36 0.13 0.09 1.76 0.41 0.25 0.31 0.35 0.83 0.69 0.5 2.91 0.65 0.16
TiO2 % 0.001 1.088 1.02 1.196 0.821 1.04 0.778 0.68 1.97 1.3 1.15 2.06 1.428 1.462 1.71 1.478 1.53 2.08 1.167 1.838
P2O5 % 0.01 0.13 0.34 0.13 0.14 0.17 0.16 0.09 0.18 0.15 0.29 0.51 0.29 0.41 0.29 0.23 0.21 0.15 0.13 0.23
LOI % 0.01 – 1.97 3.57 3.52 3.4 4.77 4.69 2.2 4.8 3.2 3.3 6.6 2.58 3.8 4.18 3.5 3.2 3.74 7.84
Total % 0.01 – 99.85 100.5 100.2 99.65 100.4 98.61 99.76 99.69 99.72 99.69 100.1 100.4 100.2 99.15 99.74 99.78 99.99 100
CaO/Na2O – – – 0.45 1.72 1.32 2.51 2.47 3.81 1.75 1.92 1.35 1.94 1.21 4.43 0.65 0.62 0.72 0.52 0.29 0.76
CaO/Fe2O3 – – – 0.42 0.58 0.60 0.61 0.50 1.27 0.70 0.38 0.70 0.70 0.60 1.19 0.31 0.39 0.40 0.16 0.20 0.41
Sc ppm 1 36 15 41 27 36 34 33 40 33 26 34 29 20 27 25 30 31 28 37
V ppm 5 276 107 384 239 315 315 184 282 298 268 262 214 184 238 185 252 170 194 278
Co ppm 1 42 21 29 26 51.3 22 32 34.6 34.5 45.5 27 33 36 32 51 34.4 39.5 34 35
Ni ppm 20 70 20 30 – 57 – 70 143.3 41.3 49.2 59.3 30 70 – 30 59.7 3.5 30 30
Cu ppm 10 100 60 70 40 21.5 50 20 2.8 40.6 42.9 7.5 – 50 – – 3.3 1.1 50 30
Zn ppm 30 90 70 100 70 39 160 50 37 48 58 96 70 80 130 130 125 41 110 130
Ga ppm 1 18 15 19 18 12.5 21 13 17.2 16.1 18.4 23.9 17 25 20 22 18 24.7 19 20
5

Rb ppm 1 20 26 9 21 46.5 18 2 0.6 16.6 5.3 5.9 5 6 20 16 11.1 91.1 14 3


Sr ppm 2 194 57 249 324 147.5 414 151 142.8 196.4 266.9 381.4 166 574 68 80 113.2 69.5 83 163
Y ppm 0.5 22.2 42.4 18.6 18.7 17.8 20.1 13.8 36.2 24 23.7 50.2 33 23.5 35.8 48.6 39.7 32.7 23.8 38.4
Zr ppm 4 91 164 54 78 33.9 67 38 129.3 64.4 64.4 308.7 126 224 186 264 139 230.2 147 175
Nb ppm 0.2 6.2 5.6 2 2.8 1.5 2.4 1.6 6.9 2.7 2.8 27 5.3 15.4 7.2 9.5 6.2 9.1 4.9 7.7
Cs ppm 0.1 0.9 3 0.6 1.1 1 0.3 – – 0.8 0.3 0.3 1.2 0.3 0.4 0.4 0.5 1.2 0.5 0.3
Ba ppm 3 176 988 458 921 1040 440 154 20 369 382 174 66 71 119 108 222 259 191 25
La ppm 0.05 11.1 16.8 6.29 10.5 4.2 12.9 3.59 5.6 6.5 7.4 22.4 10.4 22.3 15.6 14.8 11.5 17.4 12.4 9.22
Ce ppm 0.05 24.1 39 14.3 22.5 10.8 26.8 8.99 16.9 16.2 18.5 52.2 25.9 46.6 35.4 40.4 24.2 36.6 28.1 24
Pr ppm 0.01 – 5.45 2.12 2.9 1.71 3.42 1.24 2.66 2.51 2.74 6.98 3.62 5.86 4.71 5.65 3.81 4.52 3.36 3.38
Nd ppm 0.05 12.8 25.1 10.4 13.1 8.8 14.5 6.09 14.2 12.3 13.6 33 17.5 24.4 20.4 24.3 17.8 19.8 14.2 16.2
Sm ppm 0.01 3.2 6.42 2.71 3.19 2.5 3.3 1.72 4.25 3.23 3.47 7.62 4.59 5.31 5.02 6.04 4.52 4.42 3.51 4.74
Eu ppm 0.005 1.18 2.15 1.13 1.13 0.99 1.21 0.817 1.55 1.33 1.21 2.55 1.83 1.98 1.7 1.95 1.53 1.07 1.18 1.58

Journal of South American Earth Sciences xxx (xxxx) xxx


Gd ppm 0.01 – 6.82 2.96 3.14 3.07 3.28 2.05 5.65 3.95 4.02 8.79 5.27 5.17 5.45 6.76 5.61 4.88 3.81 5.32
Tb ppm 0.01 0.66 1.18 0.53 0.55 0.53 0.56 0.39 1.02 0.69 0.71 1.5 0.94 0.84 0.99 1.27 1.02 0.91 0.71 1.02
Dy ppm 0.01 4 7.34 3.25 3.4 3.09 3.41 2.38 6.05 4.08 3.97 8.53 5.92 4.74 6.23 8.01 6.19 5.48 4.55 6.74
Ho ppm 0.01 0.77 1.48 0.65 0.68 0.66 0.71 0.48 1.32 0.85 0.83 1.8 1.18 0.86 1.28 1.63 1.31 1.16 0.89 1.4
Er ppm 0.01 2.27 4.36 1.86 2.06 1.86 2.15 1.41 3.94 2.4 2.49 5.31 3.51 2.35 3.74 4.88 3.96 3.58 2.66 4.16
Tm ppm 0.005 0.331 0.659 0.269 0.331 0.28 0.329 0.207 0.56 0.36 0.36 0.78 0.532 0.327 0.542 0.721 0.55 0.55 0.396 0.631
Yb ppm 0.01 2.1 4.27 1.66 2.38 1.6 2.15 1.33 3.46 2.17 2.17 4.73 3.45 2.03 3.48 4.53 3.33 3.52 2.55 4.14
Lu ppm 0.002 0.317 0.689 0.247 0.397 0.26 0.316 0.197 0.55 0.33 0.33 0.73 0.516 0.29 0.55 0.683 0.52 0.57 0.377 0.635
Hf ppm 0.1 2.4 4.2 1.6 2.2 1.2 1.9 1.1 3.7 2.1 1.9 7.6 3.3 4.9 4.5 6.4 3.9 6.6 3.7 4.3
Ta ppm 0.01 0.52 0.56 0.35 0.88 0.3 0.72 0.67 0.5 0.2 0.6 2.1 1.12 2.21 1.46 1.39 0.6 1.2 1.01 1.29
W ppm 0.5 <0.5 76.5 71.8 228 84.3 189 219 0.6 – 169.7 47.2 258 333 322 230 107.4 220.1 171 107
Th ppm 0.05 2.07 2.79 0.93 1.83 0.3 2.09 0.28 0.4 1.2 0.9 2 1 1.48 3.64 3.74 1.7 5.1 3.09 1.2
U ppm 0.01 0.49 1.38 0.36 0.81 0.2 0.79 0.13 – 0.4 0.3 0.8 0.39 0.34 0.93 1.38 0.6 1.1 0.75 0.52
A. Bustamante et al. Journal of South American Earth Sciences xxx (xxxx) xxx

Fig. 2. Petrographic characteristics of the Jambaló region samples: (A) barroisite (Brs) and actinolite (Act) replacing glaucophane (Gln) rims in blueschists, (B)
blueschist with glaucophane (Gln) crystals replaced by chlorite (Chl) in fractures and rims, (C) decussate arrangement of glaucophane at the macroscopic scale, (D)
white mica displaying main foliations of S1 and S2. This image corresponds to an EMPA X-ray map and shows mineral mask contents in a blueschist sample, (E)
presence of glaucophane relicts in a greenschist-facies association. The red line indicates blueschist (BS) to greenschist (GS) facies transition, (F) greenschist
exhibiting metamorphic association between chlorite (Chl) and plagioclase (Pl) with glaucophane (Gln) relicts, (G) greenschist displaying inclusions of opaque
minerals in plagioclase, and (H) pseudomorphic chlorite after garnet in a greenschist-facies association. Photomicrographs were obtained with plane polarized light.
(For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

6
A. Bustamante et al. Journal of South American Earth Sciences xxx (xxxx) xxx

3%), actinolite (up to 3%), and apatite and zircon as accessory phases. A 48.40–59.27, 14.41–19.77, 1.74–11.60, and 0.68–2.08 wt%, respec­
remarkable feature of these rocks is the presence of glaucophane relics tively. During prograde and retrograde metamorphism, major elements
replaced by chlorite and light green calcic amphibole. Albite crystals are have the property of mobility, causing problems when the protolith
associated with chlorite (Fig. 2E) and opaque mineral inclusions, needs to be restrained (Rollinson, 1993). This may be reflected in weak
epidote (Fig. 2F), and occasionally quartz and white mica. The albite correlations observed in the MgO vs. CaO, MnO, and TiO2 variation
grains are deformed, indicating that the deformation is post-S1 or syn-S2 diagrams (not shown). Therefore, instead, trace elements were used for
(Fig. 2F). Pseudomorphic chlorite is also found after garnet. The opaque comparison and classification. Good positive correlations were observed
mineral inclusions observed in chlorite show continuity with plagioclase when comparing Nb and immobile elements with Hf, Sm, La, and Yb
inclusions. Quartz is concentrated in thin bands or lenses, oriented (Fig. 3). This means that these elements were unaffected by the alter­
parallel to the foliation, exhibiting strong undulatory extinction. Para­ ation, despite the LOI values of 3.54 and 4.43 wt% recorded for
gonite is also detected in the petrographic analysis. It is found in small greenschists and blueschists, respectively.
quantities, generally hosts small grains, and sometimes grows with The Nb/Y vs. Zr/Ti diagrams of Winchester and Floyd (1977) and
chlorite. Some grains are lightly folded and are accompanied by S2 those modified by Pearce (1996) use immobile elements, showing a
foliation. Epidote and/or clinozoisite also follow the rock foliation. basaltic to basaltic andesite composition for greenschists and blueschists
Zoisite is observed in limited quantities and is always found as inclusion with a sub-alkaline affinity (Fig. 4).
in plagioclase. Actinolite is prismatic and acicular-shaped. The chondrite-normalized REE patterns of blueschists and greens­
Texturally, greenschists are well foliated with superimposed de­ chists show a slight enrichment of light REEs (LREEs) when compared to
formations, resulting in the undulatory extinction of amphiboles. There heavy REEs (HREEs) (Fig. 5), as reflected in the (La/Yb)N values of
are randomly distributed glaucophane crystals, some of which exhibit 1.08–4.00 (2.42 on average) for greenschists and 1.48–3.30 (3.12 on
asymmetric pressure shadows. average) for blueschists, although one value reaches 7.32. These (La/
Yb)N values indicate that blueschists are slightly richer in LREEs than
4.2. Whole-rock geochemistry and Nd isotopes greenschists. The Eu anomaly is zero to slightly positive for greenschists
(Eu/Eu* from 1.0 to 1.3) and zero to slightly negative for blueschists
The blueschist- (8 samples) and greenschist- (10 samples) facies (Eu/Eu* from 0.7 to 1.2). Within the MORB-normalized trace element
rocks in the Jambaló region have a mafic to intermediate composition. variation plots (Pearce, 1983), both blueschist- and greenschist-facies
The ranges of MgO, SiO2, Al2O3, CaO, and TiO2 are 2.59–8.25, samples were characterized by negative Nb and Ti anomalies. Also, Ti,

Fig. 3. Nb vs. Hf (A), Sm (B), La (C), and Yb (D) diagrams showing different distribution patterns and evidencing the unaffected character of protoliths in blueschist-
and greenschist-facies rocks.

7
A. Bustamante et al. Journal of South American Earth Sciences xxx (xxxx) xxx

Fig. 4. (A) Nb/Y vs. Zr/Ti diagram of Winchester and Floyd (1977), modified by Pearce (1996), using immobile elements and showing basaltic to basaltic andesite
composition for greenschists and blueschists, with sub-alkaline affinity in the Jambaló region, (B) comparison of Nd isotope data from Western Colombia with the Nd
isotopic composition of Jambaló blueschists. Triangles correspond to Jambaló analyzed samples, and black circles correspond to samples plotted in Kerr et al. (1997).
Pentagons correspond to the Cretaceous igneous rocks from the Central Cordillera (Zapata et al., 2019), (C) TiO2/Yb vs. Nb/Yb diagram showing that most of the
samples indicate a MORB array, whereas in the Th/Yb vs. Nb/Yb diagram (D) the analysis features a relative enrichment in N-MORB signatures approaching the arc
trend that corroborates arc-related signatures. Diagrams C and D after Pearce (2008).

Y, and Yb of greenschist (mainly) and some blueschist samples plot basaltic andesite protoliths (Fig. 4). In the field, blueschists appear as
below the MORB normalizing factor (Fig. 5). Depletion of High Field lenses embedded in greenschists. In contrast, petrography suggests that
Strength Elements (HFSE) is also found in both greenschists and blues­ greenschists seem to be associated with the retrogression of blueschists,
chists (Fig. 5). Initial 143Nd/144Nd and εNd values were calculated based as evidenced by relics of glaucophane replaced by chlorite and light
on 62 Ma, the age at which white mica recrystallized or re-equilibrated. green calcic amphiboles. From a geochemical point of view, greenschists
However, the εNd unit remained unchanged. The initial values for εNd are generally characterized by higher CaO/Na2O and CaO/Fe2O3 ratios
of 0.31 and 3.04 and 143Nd/144Nd 0.512794 and 0.512654values of the lower total iron content (here considered as Fe2O3) than blueschists
two blueschists were weakly radiogenic. These values are lower than in (Dungan et al., 1983). This is consistent with the Jambaló region results,
the MORB-depleted mantle reservoirs (e.g., Faure, 2001; Dickin, 2005). with CaO/Na2O values (Table 2) ranging from 1.32 to 3.81 for greens­
By comparison, Nd isotope data of western Colombia, including Creta­ chists and 0.29 to 1.21 for blueschists. Linear arrays shown in trace
ceous rocks of the Quebradgrande Complex (Zapata et al., 2019), Nd element variation diagrams (Fig. 5) suggest that andesites to basalts,
isotope data of the Caribbean-Colombian Plateau (Kerr et al., 1997), and defined as protoliths, are consanguineous and can be associated with a
Nd isotopic composition of Jambaló blueschists are markedly different similar trend of fractional crystallization. Therefore, it can be suggested
and characterized by low values (Fig. 4). that the Jambaló blueschists and greenschists are part of a coherent
block that underwent similar metamorphic pathways with different
5. Discussion pressure and temperature intensities. Variations in the chemical
composition of protoliths may have played a significant role in the
5.1. Supra-subduction components in Jambaló protoliths preservation of blueschists vs. greenschists (Dungan et al., 1983; Bar­
rientos and Selverstone, 1993; Baziotis et al., 2009).
From the Nb/Y vs. Zr/Ti classification diagram (Fig. 4), it is clear that Bustamante and Bustamante (2019) have recently proposed that
the Jambaló blueschists and greenschists have mafic to intermediate different exposures of HP-LT rocks from the Central Cordillera of
protoliths, and various evidences suggest that these rocks are basaltic to Colombia may have different protoliths, and the timing of

8
A. Bustamante et al. Journal of South American Earth Sciences xxx (xxxx) xxx

Fig. 6. Th-Hf-Nb diagram after Wood (1980), comparing Jambaló blueschists


and greenschists with metamorphic rocks from several Northern Andes units
and the volcano-sedimentary Quebradagrande Complex. The arrow shows the
SSZ mantle enrichment trend of the Jambaló rocks. Diagram according to
Pearce et al. (1984).
Fig. 5. (A) Chondrite-normalized REE patterns of blueschists and greenschists
showing slight enrichment in light REEs (LREEs) compared to heavy REEs Jambaló rocks are part of a segmented belt that includes other occur­
(HREEs). Values are similar to a typical N-MORB, but some variances are rences of HP-LT rocks such as Barragán eclogites, which also includes
observed in light REE (LREE) values. Ocean Island Basalt (OIB) and N-MORB amphibolite-facies rocks forming the Arquía Complex along the Central
data are after Sun and McDonough (1989), modified from Bustamante and
Cordillera. They extend from Southern Ecuador to the Raspas Complex
Bustamante (2019), and (B) trace element variation plots normalized to
(Spikings et al., 2015, 2019), with Lu–Hf ages of 133.4 ± 2.1 Ma and
N-MORB (Sun and McDonough, 1989). Both blueschist- and greenschist-facies
samples are characterized by negative Nb, Zr, Hf, Ti, and Ta anomalies, 126.4 ± 4.0 Ma reported for eclogites and blueschists, respectively
modified from Bustamante and Bustamante (2019). (John et al., 2010). These authors suggest that the protoliths of this belt
have MORB-related characters and represent a subduction channel that
records the Early Cretaceous convergence in the northwestern margin of
metamorphism includes Early and Late Cretaceous events, as evidenced
South America. However, Ar–Ar geochronological data for phengite and
by the difference in 40Ar–39Ar metamorphic ages (Bustamante et al.,
paragonite from the Jambaló schist, published in Bustamante et al.
2011, 2012). The relatively enriched LREE ratios ([La/Yb]N ~3 on
(2011), yield metamorphic ages of 67–60 Ma, suggesting the existence
average) and the fact that most samples plot above the MORB array
of younger Late Cretaceous metamorphic belts associated with different
(Fig. 4) in the TiO2/Yb vs. Ta/Yb plot of Pearce (2008), the presence of
tectonic scenarios (Fig. 7).
Ta, Nb, and Ti negative anomalies (Fig. 4), and the relative enrichment
Geochemical results presented in this contribution suggest that
of Jambaló rocks, when compared to NMORB in the Nb/Yb vs. Th/Yb
protoliths from the Jambaló schists were formed in an arc-related
relationships after Pearce (2008) (Fig. 4), suggest that the Jambaló
environment (Fig. 7). Both intra-oceanic and continental arcs are rich
schists may have protoliths of supra-subduction zone (SSZ) origin. The
in LILE and LREE and depleted in HREE and Ti compared to MORB
Jambaló blueschists and greenschists, shown in the Th-Hf-Nb diagram of
(Kelemen et al., 2014). However, intra-oceanic arc lavas have high
Wood (1980), suggest an initial subtle increase in Hf content and a clear
Th/La ratios (Plank, 2005; Kelemen et al., 2014), and in contrast to La,
enrichment in Th (Fig. 6), which are characteristic of SSZ basalts.
the concentrations of Ta and Nb tend to be equal to or below MORBs
Sediment involvement during magma formation can be traced by
(Kelemen et al., 2014). Conversely, continental arc lavas display higher
Th/La ratios higher than 0.2, indicating terrigenous deposits (Plank,
Nb, La, and Ta concentrations than MORB (Kelemen et al., 2014). Th/La
2005). Hence, the subduction zone sedimentary input is also indicated
higher than 0.19, and Ta values (0.2–2.2 ppm) are similar to most ocean
by the Th/La ratios greater than 0.19 found in most samples. Besides, the
arcs (Kelemen et al., 2014). For normal arcs with melting above 15%,
Nd isotopic composition (Table 3) also shows some degree of crustal
the Zr/Nb ratio is a valuable tracer for mantle sources. Specifically, it has
contamination, which can be interpreted as an interaction between
been found that plume-influenced arc rocks generally have Zr/Nb ratios
arc-derived sediments and basalts formed in the SSZ. This setting con­
of less than 32 when compared to the higher values of normal MORB
trasts with earlier models that proposed an Early Cretaceous MORB-like
modified arcs (reviews in Neill et al., 2011).
setting for the origin of protoliths and high-pressure metamorphism
Bustamante and Bustamante (2019) proposed that the arc signature
during the convergence of the Farallon plate with NW South America
was modified after the mantle plume volcanism based on the blueschist
(Spikings et al., 2015, 2019), which involves upper plate subduction
geochemistry. We suggest that this interpretation may apply to both
erosion during convergence.
blueschists and greenschists and indicate that the arc signature may be
more similar to an intra-oceanic than a continental-like setting such as
5.2. Regional tectonic implications Quebradagrande (Fig. 7; Jaramillo et al., 2017). Hence, this arc was built
over an old Cretaceous plateau that comprises the Caribbean plate. It is
Several tectonic models of the Early Cretaceous tectonic evolution of believed that it was formed in the Southwestern Pacific as an oceanic
the Northern Andes suggest that protoliths and metamorphism of the plateau that evolved toward an intra-oceanic arc in the Early to Late

9
A. Bustamante et al. Journal of South American Earth Sciences xxx (xxxx) xxx

Table 3
Sm–Nd isotope data from the Jambaló rocks.
147
Sample # Sm(ppm) Nd(ppm) Sm/144Nd 143
Nd/144Nd εNd (0) εNd (62Ma) 143
Nd/144Nd (62)

121B 4.449 16.762 0.1605 0.512794 3.04 3.33 0.512721


121C 2.691 9.228 0.1763 0.512794 3.04 3.20 0.512374
123A 5.703 26.111 0.1321 0.512654 0.31 0.82 0.512594

Fig. 7. Cartoon (not to scale) of the suggested model


for the generation of high-pressure metamorphic
rocks at Jambaló. The figure shows rocks in the
Jambaló region at two different stages: before the
Caribbean Plateau collision (~100–70 Ma), repre­
senting the generation and exhumation of blueschist-
facies rocks and after the Caribbean Plateau collision
(~70–60 Ma), displaying the retrograde metamorphic
event in greenschist facies and the preservation of
blueschist lenses. The figure also indicates the rela­
tionship between oblique convergence and the exhu­
mation of high-pressure metamorphic units in
Jambaló. The Caribbean Plateau subducts beneath
the western margin of Colombia, forming Jambaló
rocks, with continued oblique convergence (After
Bustamante and Bustamante, 2019).

Cretaceous. During the Late Cretaceous-Early Paleocene, it collided with 2010). Another interpretation of the isotope signature is that protoliths
the South American margin, causing the onset of the Andean orogeny and metamorphism took place at the margin of the continent, not in the
(Fig. 7) (Kerr et al., 1997; Luzieux et al., 2006; Pindell and Kennan, oceanic realm, in which the old basement and sediments contributed to
2009; Villagómez et al., 2011; Hincapié-Gómez et al., 2018). magma genesis. However, such an environment does not clearly explain
In the Cauca Valley and the Western Cordillera, right west of the the mentioned Zr/Nb values and the minor enrichment of LREE.
Jambaló schists, outcrops of Cretaceous basalts, believed to be remnants Therefore, the metamorphic record of the Jambaló schists, including
of the Caribbean plate (Kerr et al., 1997), are intruded by intermediate its protoliths, documents the approach of the Caribbean intra-oceanic
granitoids with arc signature and an approximate crystallization age of arc to the continental margin. Its magmatic history may be modified
90 Ma (Villagómez et al., 2011). These rocks, along with other plutonic as South American sediments enter the oceanic trench. It also documents
bodies in the Western Cordillera, are considered part of an oceanic arc metamorphism during subduction and retrogression due to the final
(Villagómez et al., 2011; Weber et al., 2015; Zapata-Villada et al., 2017) collision from the Late Cretaceous to Paleocene. In order to test this
with an approximate age of 86 Ma in the southern position (Hinca­ model and evaluate the complex structural mixing of volcanic, meta­
pié-Gómez et al., 2018). Protoliths from the Jambaló schists were morphic, and igneous units, additional structural, geochronological, and
probably formed in this magmatic arc (Fig. 7) and were incorporated geochemical data are necessary for other volcanic and metamorphic
into the subduction channel as coherent fragments of a wide upper plate units that outcrop in this region. Such units characterize the region,
(Keppie et al., 2009). They underwent high-pressure metamorphism reflecting the Late Cretaceous Caribbean-South American arc-
during the subduction, probably exhuming at about 67 Ma. continental collision and subsequent marginal strike-slip segmentation.
Our Nd isotopes, though still limited, also indicate less radiogenic
character when compared to other Cretaceous Caribbean arcs (Fig. 4; 6. Conclusions
Hastie et al., 2009; Neill et al., 2011). This may be paradoxical in the
proposed intra-oceanic arc model. However, this signature may reflect Compositional and temporal constraints of the Jambaló schists sug­
the input of continental sediments into the trench as the intra-oceanic gest that their protolith and metamorphic evolution are independent of
arc was approaching the South American margin (Elburg et al., 2005; other Early Cretaceous high-pressure metamorphic rocks exposed in
Bouilhol et al., 2013). This is opposed to the more juvenile character of Southern Ecuador (Raspas Complex) and the Barragán region of the
pure intra-oceanic arcs that grow far away from the continent (Stern, Colombian Central Cordillera. This suggests that exposed metamorphic

10
A. Bustamante et al. Journal of South American Earth Sciences xxx (xxxx) xxx

units in western Colombia have a more extensive Cretaceous tectonic Bustamante, A., Juliani, C., Hall, C.M., Essene, E.J., 2011. 40Ar/39Ar ages from
blueschists of the Jambaló region, Central Cordillera of Colombia: Implications on
record than previously estimated.
the styles of accretion in the Northern Andes. Geol. Acta 9, 351–362.
Whole-rock geochemistry and Nd isotopes from the Late Cretaceous Bustamante, C., Archanjo, C.J., Cardona, A., Vervoort, J.D., 2016. Late Jurassic to Early
blueschist- and greenschist-facies rocks of the Colombian Central Cretaceous plutonism in the Colombian Andes: A record of long-term arc maturity.
Cordillera indicate that these rocks share common basaltic protoliths Geol. Soc. Am. Bull. 128, 1762–1779.
Bustamante, C., Archanjo, C.J., Cardona, A., Bustamante, A., Valencia, V., 2017. U-Pb
that were formed in an arc setting with plume-modified lithosphere. ages and Hf isotopes in zircons from parautochthonous Mesozoic terranes in the
Unmetamorphosed volcanic and plutonic rock units of this Cretaceous western margin of Pangea: Implications for the terrane configurations in the
intra-oceanic arc and associated oceanic plateau are exposed in the Northern Andes. J. Geol. 125, 487–500.
Bustamante, C., Bustamante, A., 2019. Two Cretaceous subduction events in the Central
Cauca Valley and the Western Cordillera. This suggests that the Jambaló Cordillera: Insights from the high P–low T metamorphism. In: Gómez, J.,
schists represent the associated subduction zone records. Pinilla–Pachon, A.O. (Eds.), The Geology of Colombia, Volume 2 Mesozoic, vol. 36.
Servicio Geológico Colombiano, Publicaciones Geológicas Especiales, Bogotá,
pp. 517–559. https://doi.org/10.32685/pub.esp.36.2019.14.
Declaration of competing interest Cochrane, R., Spikings, R., Gerdes, A., Ulianov, A., Mora, A., Villagómez, D., Putlitz, B.,
Chiaradia, M., 2014. Permo-Triassic anatexis, continental rifting and the disassembly
of western Pangaea. Lithos 190–191, 383–402.
The authors declare that they have no known competing financial Dickin, A.P., 2005. Radiogenic Isotope Geology, second ed. Cambridge University Press,
interests or personal relationships that could have appeared to influence Cambridge, p. 492.
the work reported in this paper. Dungan, M.A., Vance, J.A., Blanchard, D.P., 1983. Geochemistry of the Shuksan
greenschists and blueschists, North Cascades, Washington: Variably fractionated and
altered metabasalts of oceanic affinity. Contrib. Mineral. Petrol. 82, 131–146.
Acknowledgments Elburg, M.A., Foden, J.D., van Bergen, M.J., Zulkarnain, I., 2005. Australia and Indonesia
in collision: geochemical sources of magmatism. J. Volcanol. Geoth. Res. 140, 25–47.
Ernst, W.G., 1988. Tectonic history of subduction zones inferred from retrograde
We would like to thank Thaís Hyppolito for her major discussions blueschist P-T paths. Geology 16, 1081–1084.
during the development of this project and Antonio Garcia Casco for Evans, B.W., 1990. Phase relations of epidote-blueschists. Lithos 25, 3–23.
support on EMPA analyses. Ralf Halama, Samuele Papeschi and an Faure, G., 2001. Origin of Igneous Rocks: the Isotopic Evidence. Springer Verlag, Berlin,
p. 496.
anonymous reviewer are acknowledged by their important comments Feininger, T., 1980. Eclogite and related high-pressure regional metamorphic rocks from
and improvements. The authors also thanks to the Colombian Fundación the Andes of Ecuador. J. Petrol. 21, 107–140.
para la Promoción de la Investigación y la Tecnología (grant 1819) and, Feininger, T., 1982. Glaucophane schist in the Andes at Jambaló, Colombia. Can.
Mineral. 20, 41–47.
from Brazil, the Fundação de Amparo à Pesquisa do Estado de São Paulo Gioia, S.M.C.L., Pimentel, M.M., 2000. The Sm-Nd isotopic method in the geochronology
(FAPESP, grants 2004/10203-7 and 2009/17380-5), and the National laboratory of the University of Brasília. An Acad. Bras Ciências 72, 219–245.
Council for Scientific and Technological Development (CNPq, grant Guillot, S., Hattori, K., Agard, P., Schwartz, S., Vidal, O., 2009. Exhumation Processes in
Oceanic and Continental Subduction Contexts: A Review. In: Lallemand, S.,
141946/2005-9) for their financial support.
Funiciello, F. (Eds.), Subduction Zone Geodynamics. Springer, pp. 175–205.
Hastie, A.R., Kerr, A.C., Mitchell, S.F., Millar, I.L., 2009. Geochemistry and
References tectonomagmatic significance of lower Cretaceous island arc lavas from the Devil’s
Racecourse Formation, eastern Jamaica. In: James, K.H., Lorente, M.A., Pindell, J.
(Eds.), The Origin and Evolution of the Caribbean Plate, vol. 328. Geological Society
Agard, P., Yamato, P., Jolivet, L., Burov, E., 2009. Exhumation of oceanic blueschists and
of London Special Publication, pp. 337–359.
eclogites in subduction zones: timing and mechanisms. Earth Sci. Rev. 92, 53–79.
Hincapié-Gómez, S., Cardona, A., Jiménez, G., Monsalve, G., Ramírez-Hoyos, L.,
Arculus, R.J., Lapierre, H., Jaillard, É., 1999. Geochemical window into subduction and
Bayona, G., 2018. Paleomagnetic and gravimetrical reconnaissance of Cretaceous
accretion processes: Raspas metamorphic complex, Ecuador. Geology 27, 547–550.
volcanic rocks from the Western Colombian Andes: Paleogeographic connections
Aspden, J.A., Bonilla, W., Duque, P., 1995. The El Oro metamorphic complex, Ecuador:
with the Caribbean Plate. Studia Geophys. Geod. 62, 485–511.
geology and economic mineral deposits. In: Overseas Geology and Mineral
Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M., Reyes-Harker, A., Stockli, D.F.,
Resources, vol. 67. British Geological Survey Publication, Nottingham, p. 63.
2010. Linking sedimentation in the northern Andes to basement configuration,
Aspden, J.A., McCourt, W.J., 1986. Mesozoic oceanic terrane in the central Andes of
Mesozoic extension, and Cenozoic shortening: Evidence from detrital zircon U-Pb
Colombia. Geology 14, 415–418.
ages, Eastern Cordillera, Colombia. Geol. Soc. Am. Bull. 122, 1423–1442.
Avellaneda-Jiménez, D.S., Cardona, A., Valencia, V., Barbosa, J.S., Jaramillo, J.S.,
Jaramillo, J.S., Cardona, A., León, S., Valencia, V., Vinasco, C., 2017. Geochemistry and
Monsalve, G., Ramírez-Hoyos, L., 2019. Erosion and regional exhumation of an Early
geochronology from Cretaceous magmatic and sedimentary rocks at 6◦ 35’N, western
Cretaceous subduction/accretion complex in the Northern Andes. Int. Geol. Rev. 62,
flank of the Central Cordillera (Colombian Andes): magmatic record of arc-growth
186–209.
and collision. J. S. Am. Earth Sci. 76, 460–481.
Barrientos, X., Selverstone, J., 1993. Infiltration vs. thermal overprint of epidote
Janoušek, V., Farrow, C.M., Erban, V., 2006. Interpretation of whole-rock geochemical
blueschists. Ile de Groix, France. Geology 21, 69–72.
data in igneous geochemistry: introducing Geochemical Data Toolkit (GCDkit).
Baziotis, I., Proyer, A., Mposkos, E., 2009. High-pressure/low-temperature
J. Petrol. 47, 1255–1259.
metamorphism of basalts in Lavrion (Greece): implications for the preservation of
John, T., Scherer, E.E., Schenk, V., Herms, P., Halama, R., Garbe-Schönberg, D., 2010.
peak metamorphic assemblage in blueschist and greenschist. Eur. J. Mineral 21,
Subducted seamounts in an eclogite-facies ophiolite sequence: the Andean Raspas
133–148.
Complex, SW Ecuador. Contrib. Mineral. Petrol. 159, 265–284.
Blanco-Quintero, I.F., García-Casco, A., Toro, L.M., Moreno, M., Ruiz, E.C., Vinasco, C.J.,
Johnson, D.M., Hooper, P.R., Conrey, R.M., 1999. XRF Analysis of Rocks and Minerals for
Cardona, A., Lázaro, C., Morata, D., 2014. Late jurassic terrane collision in the
Major and Trace Elements on a Single Low Dilution Li-Tetraborate Fused Bead.
northwestern margin of gondwana (Cajamarca complex, eastern flank of the central
JCPDS-International Centre for Diffraction Data, pp. 843–867.
cordillera, Colombia). Int. Geol. Rev. 56, 1852–1872.
Kelemen, P.B., Hanghøj, K., Greene, A.R., 2014. One View of the Geochemistry of
Bosch, D., Gabriele, P., Lapierre, H., Malfere, J.-L., Jaillard, E., 2002. Geodynamic
Subduction-Related Magmatic Arcs, with an Emphasis on Primitive Andesite and
significance of the Raspas Metamorphic Complex (SW Ecuador): geochemical and
Lower Crust. Treat. Geochem. 749–806.
isotopic constraints. Tectonophysics 345, 83–102.
Keppie, D.F., Currie, C.A., Warren, C., 2009. Subduction erosion modes: Comparing finite
Bouilhol, P., Jagoutz, O., Hancharb, J.M., Dudasa, F.O., 2013. Dating the India-Eurasia
element numerical models with the geological record. Earth Planet Sci. Lett. 287,
collision through arc magmatic records. Earth Planet Sci. Lett. 366, 163–175.
241–254.
Bourgois, J., Toussaint, J.F., Gonzalez, H., Azema, J., Calle, B., Desmet, A., Murcia, L.A.,
Kerr, A.C., Marriner, G.F., Tarney, J., Nivia, A., Saunders, A.D., Thirlwall, M.F., Sinton, C.
Acevedo, A.P., Parra, E., Tournon, J., 1987. Geological history of the Cretaceous
W., 1997. Cretaceous basaltic terranes in western Colombia: elemental,
ophiolitic complexes of northwestern South America (Colombian Andes).
chronological and Sr-Nd isotopic constraints on petrogenesis. J. Petrol. 38, 677–702.
Tectonophysics 143, 307–327.
Lanari, P., Vho, A., Bovay, T., Airaghi, L., Centrella, S., 2019. Quantitative compositional
Bröcker, M., 1990. Blueschist-to-greenschist transition in metabasite from Tinos Island
mapping of mineral phases by electron probe micro-analyser. Geol. Soc. Londn.
(Cyclades, Greece): compositional control or fluid infiltration? Lithos 25, 25–39.
Special Publ. 478, 39–63.
Brown, E.H., 1986. Geology of the Shuksan Suite, North Cascades, Washington, U.S.A.
Lanari, P., Vidal, O., De Andrade, V., Dubacq, B., Lewin, E., Grosch, E., Schwartz, S.,
Geol. Soc. Am. Bull. 164, 143–154.
2014. XMapTools: a MATLAB©-based program for electron microprobe X-ray image
Bustamante, A., 2008. Geotermobarometria, geoquímica, geocronologia e evolução
processing and geothermobarometry. Comput. Geosci. 62, 227–240.
tectônica das rochas da fácies xisto azul nas áreas de Jambaló (Cauca) e Barragán
Luzieux, L.D.A., Heller, F., Spikings, R., Vallejo, C.F., Winkler, W., 2006. Origin and
(Valle del Cauca), Colômbia. PhD Thesis. Universidade de São Paulo, p. 178. p.
Cretaceous tectonic history of the coastal Ecuadorian forearc between 1◦ S-4◦ S:
Available online since 08-25-2008 at. www.teses.usp.br. Last month checked
paleomagnetic, radiometric and fossil evidence. Earth Planet Sci. Lett. 249, 400–414.
September of 2020.
Maruyama, S., Liou, J.G., Terabayashi, M., 1996. Blueschists and eclogites of the world
Bustamante, A., Juliani, C., Essene, E.J., Hall, C.M., Hyppolito, T., 2012. Geochemical
and their exhumation. Int. Geol. Rev. 38, 485–594.
constraints on blueschist- and amphibolite-facies rocks of the Central Cordillera of
Colombia: The Andean Barragán Region. Int. Geol. Rev. 54, 1013–1030.

11
A. Bustamante et al. Journal of South American Earth Sciences xxx (xxxx) xxx

Maruyama, S., Cho, M., Liou, J.G., 1986. Experimental investigations of blueschist- Possible evidence for Late Cretaceous initiation of subduction in the Caribbean. Geol.
greenschist transition equilibria: pressure dependence of Al2O3 contents in sodic Soc. Am. Bull. 111, 831–848.
amphiboles - a new geobarometer. Geol. Soc. Am. Spec. Pap. 164, 1–16. Spikings, R., Cochrane, R., Vallejo, C., Villagomez, D., Van der Lelij, R., Paul, A.,
Maya, M., 2001. Distribución, facies y edad de las rocas metamórficas en Colombia, Winkler, W., 2019. Chapter 7 - Latest Triassic to Early Cretaceous tectonics of the
memoria explicativa: Mapa Metamórfico de Colombia. Instituto de Investigación e Northern Andes: Geochronology, geochemistry, isotopic tracing, and
Información Geocientífica, Minero - Ambiental y Nuclear, Ingeominas, Bogotá, p. 59. thermochronology. In: Horton, B.K., Folguera, A. (Eds.), Andean Tectonics. Elsevier,
Maya, M., González, H., 1995. Unidades litodémicas en la Cordillera Central de pp. 173–208.
Colombia. Bol. Geol. - Ingeominas 35, 43–57. Spikings, R., Cochrane, R., Villagómez, D., Van der Lelij, R., Vallejo, C., Winkler, W.,
Maya, M., Vásquez, E., 2001. Mapa Metamórfico de Colombia, escala 1:1’000.000. Beate, B., 2015. The geological history of northwestern South America: from
Ingeominas, Bogotá. Pangaea to the early collision of the Caribbean Large Igneous Province (290–75 Ma).
McCourt, W.J., Feininger, T., 1984. High pressure metamorphic rocks in the Central Gondwana Res. 27, 95–139.
Cordillera of Colombia. Br. Geol. Surv. Reprint Ser. 84, 28–35. Stern, R.J., 2010. The anatomy and ontogeny of modern intra-oceanic arc systems. In:
Neill, I., Kerr, A.C., Hastie, A.R., Stanek, K.P., Millar, I.L., 2011. Origin of the Aves Ridge Kusky, T.M., Zhai, M.-G., Xiao, W. (Eds.), The Evolving Continents: Understanding
and Dutch-Venezuelan Antilles: interaction of the Cretaceous ‘Great Arc’ and Processes of Continental Growth, vol. 338. Geological Society, London, Special
Caribbean-Colombian Oceanic Plateau? J. Geol. Soc. Londn. 168, 333–348. Publications, pp. 7–34.
Nivia, A., Marriner, G.F., Kerr, A.C., Tarney, J., 2006. The Quebradagrande Complex: A Sun, S., McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts:
Lower Cretaceous ensialic marginal basin in the Central Cordillera of the Colombian implications for mantle composition and processes. In: Saunder, A.D., Norry, M.J.
Andes. J. S. Am. Earth Sci. 21, 423–436. (Eds.), Magmatism in the Ocean Basins, vol. 42. Geological Society Special
Oberhänsli, E., 1986. Blue amphiboles in metamorphosed Mesozoic mafic rocks from the Publication, pp. 313–345.
central Alps, 164. Geological Society of America Memoir, pp. 239–248. Toussaint, J.F., 1996. Evolución Geológica de Colombia. Cretácico. Universidad Nacional
Orrego, A., Leon, L., Padilla, L., Acevedo, A.P., Marulanda, N., 1993. Mapa geológico de de Colombia, Medellín, p. 277.
la Plancha 364 - Timbío. Escala 1:100.000. INGEOMINAS. Villagómez, D., Spikings, R., Magna, T., Kammer, A., Winkler, W., Beltrán, A., 2011.
Pearce, J.A., 1983. Role of the Sub-Continental Lithosphere in Magma Genesis at Active Geochronology, geochemistry and tectonic evolution of the Western and Central
Continental Margins. In: Hawkesworth, C.J., Norry, M.J. (Eds.), Continental Basalts Cordilleras of Colombia. Lithos 125, 875–896.
and Mantle Xenoliths. Shiva, Nantwich, pp. 230–249. Vinasco, C.J., Cordani, U.G., González, H., Weber, M., Peláez, C., 2006.
Pearce, J.A., 1996. A User’s Guide to Basalt Discrimination Diagrams. Geol. Assoc. Geochronological, isotopic, and geochemical data from Permo-Triassic granitic
Canada Short Course Notes 12, 79–113. gneisses and granitoids of the Colombian Central Andes. J. S. Am. Earth Sci. 21,
Pearce, J.A., 2008. Geochemical fingerprinting of oceanic basalts with applications to 355–371.
ophiolite classification and the search for Archean oceanic crust. Lithos 100, 14–48. Weber, M.B.I., Cardona, A., Valencia, V., Gómez, J., Gómez, E., Pardo, A., 2015.
Pearce, J.A., Lippard, S.J., Roberts, S., 1984. Characteristics and tectonic significance of Geochemistry of the Santa Fé Batholith and Buriticá Tonalite in NW Colombia –
supra-subduction zone ophiolites. In: Kokelaar, B.P., Howells, M.F. (Eds.), Marginal Evidence of subduction initiation beneath the Colombian Caribbean Plateau. J. S.
Basin Geology: Volcanic and Associated Sedimentary and Tectonic Processes in Am. Earth Sci. 62, 257–274.
Modern and Ancient Marginal Basins, vol. 16. Geological Society Special Winchester, J.A., Floyd, P.A., 1977. Geochemical discrimination of different magma
Publications, pp. 77–94. series and their differentiation products using immobile element. Chem. Geol. 20,
Pindell, J.L., Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and 325–343.
northern South America in the mantle reference frame: an update. In: James, K., Wood, D.A., 1980. The application of a Th–Hf–Ta diagram to problems of
Lorente, M.A., Pindell, J. (Eds.), The Geology and Evolution of the Region between tectonomagmatic classification and to establishing the nature of crustal
North and South America, vol. 328. Geological Society of London Special contamination of basaltic lavas of the British Tertiary volcanic province. Earth
Publication, pp. 1–55. Planet Sci. Lett. 50, 11–30.
Plank, T., 2005. Constraints from Thorium/Lanthanum on Sediment Recycling at Yamasaki, T., 2018. Contamination from mortars and mills during laboratory crushing
Subduction Zones and the Evolution of the Continents. J. Petrol. 46, 921–944. and pulverizing. Bull. Geol. Surv. Jpn. 69, 201–210.
Restrepo, J.J., Ordóñez-Carmona, O., Armstrong, R., Pimentel, M.M., 2011. Triassic Yokohama, K., Brothers, R.N., Black, P.M., 1986. Regional eclogite facies in the high-
metamorphism in the northern part of the Tahamí Terrane of the central cordillera of pressure metamorphic belt of New Caledonia. Geol. Soc. Am. Mem. 164, 407–423.
Colombia. J. S. Am. Earth Sci. 32, 497–507. Zapata, S., Cardona, A., Jaramillo, J.S., Patiño, A., Valencia, V., León, S., Mejía, D.,
Rollinson, H., 1993. Using Geochemical Data. Longman Scientific and Technical, UK, Pardo-Trujillo, A., Castañeda, J.P., 2019. Cretaceous extensional and compressional
p. 352. tectonics in the Northwestern Andes, prior to the collision with the Caribbean
Sarmiento-Rojas, L.F., Van Wess, J.D., Cloetingh, S., 2006. Mesozoic transtensional basin oceanic plateau. Gondwana Res. 66, 207–226.
history of the Eastern Cordillera, Colombian Andes: inferences from tectonic models. Zapata-Villada, J.P., Restrepo, J.J., Cardona, A., Martens, U., 2017. Geoquímica y
J. S. Am. Earth Sci. 21, 383–411. geocronología de las rocas volcánicas básicas y el Gabro de Altamira, Cordillera
Smith, C.A., Sisson, V.B., Avé Lallemant, H.G., Copeland, P., 1999. Two contrasting Occidental (Colombia): Registro de ambientes de Plateau y arco oceánico
pressure-temperature-time paths in the Villa de Cura blueschist belt,Venezuela: superpuestos durante el cretácico. Bol. Geol. 39, 13–30.

12

You might also like