You are on page 1of 15

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

net/publication/223259129

The Quebradagrande Complex: A Lower Cretaceous ensialic marginal basin in


the Central Cordillera of the Colombian Andes

Article  in  Journal of South American Earth Sciences · January 2006


DOI: 10.1016/j.jsames.2006.07.002

CITATIONS READS

99 769

4 authors, including:

Jose Alvaro Nivia Andrew C. Kerr


Servicio Geológico Colombiano Cardiff University
17 PUBLICATIONS   1,322 CITATIONS    173 PUBLICATIONS   8,950 CITATIONS   

SEE PROFILE SEE PROFILE

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

Petrogenesis of plagiogranites in the Muslim Bagh Ophiolite, Pakistan (project completed as part of Masters degree) View project

Precambrian Greenstone Belts: their genesis and role in crustal evolution View project

All content following this page was uploaded by Andrew C. Kerr on 09 November 2017.

The user has requested enhancement of the downloaded file.


Journal of South American Earth Sciences 21 (2006) 423–436
www.elsevier.com/locate/jsames

The Quebradagrande Complex: A Lower Cretaceous ensialic


marginal basin in the Central Cordillera of the Colombian Andes
a,*
Alvaro Nivia , Giselle F. Marriner b, Andrew C. Kerr c, John Tarney d

a
Instituto Colombiano de Geologı́a y Minerı́a – INGEOMINAS, Unidad Operativa Cali, A.A. 9724, Cali, Colombia
b
Department of Geology, Royal Halloway University of London, Egham, Surrey TW20 0EX, UK
c
School of Earth, Ocean and Planetary Sciences, Cardiff University, Main Building, Park Place, Cardiff CF1 3YE, UK
d
University of Leicester, Department of Geology, Leicester LE1 7RH, UK

Received 1 October 2004; accepted 1 January 2006

Abstract

The Quebradagrande Complex of Western Colombia consists of volcanic and Albian–Aptian sedimentary rocks of oceanic affinity
and outcrops in a highly deformed zone where spatial relationships are difficult to unravel. Berriasian–Aptian sediments that display
continental to shallow marine sedimentary facies and mafic and ultramafic plutonic rocks are associated with the Quebradagrande Com-
plex. Geochemically, the basalts and andesites of the Quebradagrande Complex mostly display calc-alkaline affinities, are enriched in
large-ion lithophile elements relative to high field strength elements, and thus are typical of volcanic rocks generated in supra-subduction
zone mantle wedges. The Quebradagrande Complex parallels the western margin of the Colombian Andes’ Central Cordillera, forming a
narrow, discontinuous strip fault-bounded on both sides by metamorphic rocks. The age of the metamorphic rocks east of the Quebra-
dagrande Complex is well established as Neoproterozoic. However, the age of the metamorphics to the west – the Arquı́a Complex – is
poorly constrained; they may have formed during either the Neoproterozoic or Lower Cretaceous. A Neoproterozoic age for the Arquı́a
Complex is favored by both its close proximity to sedimentary rocks mapped as Paleozoic and its intrusion by Triassic plutons. Thus, the
Quebradagrande Complex could represent an intracratonic marginal basin produced by spreading-subsidence, where the progressive
thinning of the lithosphere generated gradually deeper sedimentary environments, eventually resulting in the generation of oceanic crust.
This phenomenon was common in the Peruvian and Chilean Andes during the Uppermost Jurassic and Lower Cretaceous. The marginal
basin was trapped during the collision of the Caribbean–Colombian Cretaceous oceanic plateau, which accreted west of the Arquı́a Com-
plex in the Early Eocene. Differences in the geochemical characteristics of basalts of the oceanic plateau and those of the Quebradagrande
Complex indicate these units were generated in very different tectonic settings.
Ó 2006 Elsevier Ltd. All rights reserved.

Keywords: Continental active margin; Back arc basin; Extensional tectonics; Ophiolitic complexes

1. Introduction based magmatic arcs. Throughout the Mesozoic and prob-


ably much of the Paleozoic, the western side of South
A characteristic feature of many convergent plate mar- America and the Antarctic Peninsula formed a semicontin-
gins, especially those affected by the subduction of old, uous magmatic arc along the margin of Gondwana, where
dense, oceanic lithosphere, is the development of backarc the formation of volcanic arcs and marginal basins clearly
basins resulting from extensional tectonics (Molnar and played an important evolutionary role (Dalziel et al., 1974;
Atwater, 1978; Tarney et al., 1981; Saunders and Tarney, Dalziel, 1981; Atherton et al., 1983; Saunders and Tarney,
1982). Such basins commonly occur in intra-oceanic set- 1984; Miller et al., 1994). During the Uppermost Jurassic–
tings, but examples also are associated with continental- Lower Cretaceous, the continental margin was on the brink
of splitting from the continent, and basin formation with
*
Corresponding author. the eruption of mantle-derived basalts was an outstanding
E-mail address: anivia@ingeominas.gov.co (A. Nivia). feature. A rosary of N–S-elongated, ensialic marginal

0895-9811/$ - see front matter Ó 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jsames.2006.07.002
424 A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436

basins of Tithonian to Albian age from Cape Horn and basins along the Northern Andes, according to the geo-
South Georgia Island to southwestern Mexico – the latter chemical results we present herein, is represented by the
continuous with South America in most Triassic–Jurassic Quebradagrande Complex that outcrops east of the accret-
Pangea reconstructions (Coney and Evenchick, 1994) – ed plateau in a more ensialic position and has been errone-
remain as scars of a generalized extensional episode. ously considered part of the latter (Bourgois et al., 1982,
In Colombia, a sequence of pillow lavas, diabases, and 1985, 1987; Toussaint and Restrepo, 1993; Kammer,
associated volcaniclastic materials, the Diabase and Dagua 1995; Kammer and Mojica, 1996).
groups (Nelson, 1957; Barrero, 1979) is postulated to con-
tinue this belt of marginal basins northward (Åberg et al., 2. Regional geology and stratigraphic nomenclature
1984; Aguirre, 1987; Aguirre and Atherton, 1987). Howev-
er, these materials correspond to younger (Upper Creta- On the western flank of the Central Cordillera of the
ceous) accreted terranes of oceanic plateau affinity Colombian Andes, a sequence of intermediate to basic vol-
(Millward et al., 1984; Nivia, 1987; Kerr et al., 1996, canic and sedimentary rocks of lower Cretaceous age
1997a,b, 2001). The missing link in the chain of marginal appears. In northern Colombia, this sequence was originally

78 oW 76 o W
.

Atlantic
Ocean
PANAMA 0º

SOUTH AMERICA

Pacific 20ºS
Ocean

AN1416 40ºS

AN1417
80ºW 60ºW 40ºW

Medellín
6 oN

AN1419

QBG95-2
QBG95-3

QBG95-7
3 QBG95-10
Manizales QBG95-11
2
1

PACIFIC Armenia
OCEAN AN1412
AN1410

4 oN
AN1410A LEGEND

Cenozoic rocks and deposits

Cali Western Lithospheric Oceanic


Cretaceous Province

3
Quebradagrande Complex - QGC

Popayán Triassic plutons

Arquía Complex
2 oN 2

Cajamarca Complex
AN1425
AN1426 MAIN STRUCTURAL ELEMENTS
1
1 San Jerónimo Fault
Pasto 2 Silvia-Pijao Fault
0 50 100 Km 3 Cauca-Almaguer Fault
ECUADOR

Fig. 1. Geological sketch map of western Colombia to show the distribution, spatial relationships of the Lower Cretaceous Quebradagrande Complex,
and location of samples listed in Table 2.
A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436 425

described as the Quebradagrande Formation (Botero, sandstones with clasts of cobbles and pebbles of both vol-
1963). Although regionally exposed (Fig. 1), it is covered canic rocks and chert (Gómez et al., 1995). The presence of
in places by recent volcanic and volcanoclastic deposits that these rocks suggests underwater volcanoclastic sedimenta-
make correlations difficult. Parts of this sequence have been tion produced by mass movements. The metasedimentary
mapped as the Aranzazu-Manizales sedimentary Complex horizons also contain lithic sandstones and volcanoclastic
(Gómez et al., 1995) and the Aranzazu-Manizales metasedi- arkoses. In the lithic sandstones, González (1980a) reports
mentary volcanic Complex (Mosquera, 1978); in other loca- basic volcanic rock fragments as the main components,
tions, it was mapped (Paris and Marı́n, 1979) within the with smaller quantities of mudstones and chert fragments,
Diabase Group (Nelson, 1957, 1962). The Quebradagrande whereas the clastic arkoses are dominated by plagioclase.
rocks consist of imbricated slices of strongly deformed Lozano et al. (1984b) report black and grey graphitic meta-
dynamometamorphic rocks, with crenulation cleavage and greywackes. Milonite slices, up to 1 km thick, formed from
Andean milonitic foliation that bears NNE and dips 50– clay-rich carbonaceous mudstones, intercalated with thin
70° to the east (Lozano et al., 1984a; Kammer, 1995; Gómez beds of limestone and cherts (González, 1980a).
et al., 1995). Deformation in these rocks has caused them to Marine fossils found in these metasedimentary rocks
be described erroneously as schists and included within the include ammonites, gastropods, bivalves, radiolarians, bra-
Cajamarca belt of metamorphic rocks (Nelson, 1957, 1962; chiopods, and residues of plants (Gómez et al., 1995).
Mosquera, 1978). Moreover, deformation has prevented According to González (1980a), faunas within the Quebra-
the identification of sedimentary sequences within the dagrande Complex would have lived in epineritic to brackish
Quebradagrande rocks (Rodrı́guez and Rojas, 1985); waters. However, González (1980a) interprets these rocks as
though they were originally defined as a formation with dis- part of a turbiditic sequence, whereas Lozano et al. (1984a),
tinct sedimentary and volcanic members, these lithostrati- on the basis of the lack of maturity of the sedimentary com-
graphic units lack precisely defined limits. ponents, suggest they accumulated in deep trenches. The fos-
To solve the problem of stratigraphic nomenclature sils range in age from Valanginian to Albian (140–97 Ma)
caused by the deformation of the Quebradagrande rocks, (González, 1980a; Gómez et al., 1995). Toussaint and Rest-
Maya and Gonzalez (1996) propose a stratigraphic scheme repo (1978) report a K–Ar (whole-rock) age of 105 ± 10 Ma
based on lithodemic units (North American Commission on from a basalt of the Quebradagrande Complex. Although
Stratigraphic Nomenclature, 1983). They assign the rank of K–Ar dating is notoriously unreliable in volcanic rocks as
structural complexes to the former Cajamarca, Quebrada- altered as those of the Quebradagrande Complex, the age
grande, and Arquı́a units (Table 1). Because the names of reported by Toussaint and Restrepo (1978) nonetheless
the regional faults that bound these complexes also vary agrees well with paleontological ages.
along their length, new names for the faults have been pro- Areno-rudaceous clastic sequences also are associated
posed. Thus, the fault separating the Cajamarca Complex with the Quebradagrande Complex. In northern Colombia,
to the east and the Quebradagrande Complex to the west these rocks are known as the Abejorral (Bürgl and Radelli,
has become known as the San Jerónimo fault, whereas that 1962), Valle Alto (González, 1980a), and La Soledad (Hall
which separates the Quebradagrande Complex to the east et al., 1972) formations; to the south, they are known as the
and the Arquı́a Complex to the west is called the Silvia– San Francisco (Orrego et al., 1976) and Rojiza (Orrego,
Pijao fault (Fig. 1, Table 1). Another important fault is 1993) sedimentary sequences (Table 1). The stratigraphic
the Cauca–Almaguer fault that bounds the Arquı́a Com- relationships among these units are difficult to establish,
plex to the west and, according to some petrogenetic models but their discordant deposition on top of the Cajamarca
(McCourt et al., 1984; Aspden and McCourt, 1986; Aspden Complex and general transgressive character have been
et al., 1987), marks the limit between Palaeozoic metamor- described in several localities (Bürgl and Radelli, 1962; Hall
phic rocks of continental affinity and Cretaceous accreted et al., 1972; González, 1980a). Within the Valle Alto and
terranes of oceanic character. The latter include the Cañas- Abejorral formations, Rodrı́guez and Rojas (1985) recog-
gordas, Diabase, and Dagua groups, the Amaime and Vol- nize sedimentary facies that vary with time from continen-
canic formations, and so forth. For these, Nivia (1997) uses tal to offshore marine to brackish, marine-brackish, and
the term ‘‘Western Oceanic Cretaceous Lithospheric Prov- littoral-marine. The fossils in these rocks indicate they
ince’’ (Fig. 1, Table 1), which corresponds to the southern are not older than Berriasian–middle Albian (Etayo, 1985).
extreme of the Caribbean–Colombian Cretaceous Igneous Imbricated slices of gabbro and ultramafic rocks are
Province (Kerr et al., 1996, 1997a,b) that has been accreted closely associated with the Quebradagrande Complex in
onto the Northern Andes. several localities and often show the same degree of defor-
The Quebradagrande Complex is composed of an mation. The most studied outcrops are the Liborina and
assemblage of metavolcanic and metasedimentary rocks. Sucre peridotites, the Pereira gabbro, the Pacora and Cor-
The protoliths of the metavolcanic rocks were basaltic to doba complexes, and a series of small bodies mapped as the
andesitic lavas and pyroclastics affected by the metamor- Romeral gabbros (Calle et al., 1980; González, 1980b,c;
phism of zeolite, prhenite–pumpellyite, and greenschist Mejı́a et al., 1983a,b). Toussaint and Restrepo (1974)
facies. The metasedimentary rocks display a wide variation group some of these intrusive rocks and volcanic rocks of
in grain size, from breccias and conglomerates to coarse the Quebradagrande Complex within the Cauca ophiolitic
426
Table 1
Stratigraphic units of western Colombia
Western Oceanic Arquia Complex Quebradagrande Complex Cajamarca Complex

A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436


Cretaceous Lithospheric
Province
(Upper Cretaceous) Cauca–Almaguer (Neoproterozoic - ?) Silvia–Pijao (Berriasian to middle Albian) San Jeronimo (Neoproterozoic)
Fault Fault Fault
N Marine sediments Areno-rudaceous clastic sequences
Penderisco Fm.: Abejorral Fm. Cajamarca Series
Urrao Member Valle Alto Fm. Cajamarca Group
Nutibara Member La Soledad Fm.
Dagua Group: San Francisco sed. sequence
Cisneros Fm. Rojiza sedimentary sequence
Espinal Fm.
Rio Piedras Fm.
Ampudia Fm.
Marilopito Fm.
Aguaclara Fm.
S
N
Plateau volcanics Mainly meta-volcanics and meta-sediments Lavas and pyroclastics
Barroso Fm. Arquia Group Quebradagrande Fm.
Diabase Group Bugalagrande Schists Aranzazu-Manizales
(meta)-Sedimentary Complex
Amaime Fm. La Mina Greenschists
S
Volcanic Fm.
Mafic-ultramafic rocks Metamorphic basic plutonics Mafic- ultramafic rocks
N Bolivar Ultramafic Complex Rosario Amphibolites Liborina peridotite
Ginebra Ophiolitic Massif Bolo Azul Metagabbroids Sucre peridotite
San Antonio Amphibolites Pacora Complex
and Metagabbroids
Romeral Gabbros
S
Pereira gabbro
A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436 427

Complex and suggest a possible cogenetic relationship as Th. The scatter of the Quebradagrande Complex sam-
between them (see also González, 1980a). ples on a variation diagram (Fig. 2) suggests that the alkalis
have been relatively mobile in the rocks, which might have
3. Geochemistry displaced some samples to the mugearite field in the total
alkali-silica diagram (Fig. 3). Similarly, the only sample
3.1. Sampling localities that contains greater than 62 wt% SiO2 is petrographically
identical to the andesites but looks far more altered in thin
The volcanic rocks of the Quebradagrande Complex section. It is generally considered that large-ion lithophile
were sampled on a regional basis (Fig. 1) between latitudes elements (LILE), such as K, Ba, Sr, and Rb, are relatively
6°35 0 N (Santa Fé de Antioquia) and 1°45 0 N (El Rosal– more mobile during low-grade metamorphism than high
Cauca). Sampling was performed along main roads that field strength elements (HFSE), such as Ti, P, Nb, Y, Zr,
cut the Quebradagrande Complex outcrop areas, such as and rare earth elements (REE) (Wood et al., 1979; Pearce,
Santa Fé de Antioquia–Planadas (6°35’ N), Medellı́n–Ebéj- 1983). Consequently, we place more emphasis on the geo-
ico (6°22 0 N), Itagüı́–Heliconia (6°13 0 N), Fredonia–Santa chemical behavior of the relatively immobile, trace HFSE.
Barbara (5°55 0 N), Arma–Aguadas (5°37 0 N), Aguadas– Five samples that show petrographical evidence of alter-
Pacora (5°35 0 N), Pacora–San Bartolome (5°33 0 N), Pac- ation and scatter on variation diagrams that include the
ora–Salamina (5°27 0 N), Salamina–La Merced (5°24 0 N),
Armenia–Cajamarca (4°30 0 N), Pijao–Cordoba (4°20 0 N),
and Bolı́var–El Rosal (1°45 0 N).

3.2. Analytical methods

Twenty-six samples from the Quebradagrande Complex


were analyzed for major and trace elements at the geo-
chemistry laboratories of Royal Halloway University of
London and University of Leicester. Samples were broken
into chips using manual and hydraulic jaw splinters to
remove weathered surfaces and thin veins of altered mate-
rial. The samples were further crushed using a flypress of a
steel die and shoe. Some 150 g of the coarse material was
ground into a fine powder (<240 mesh) in an agate tema
swing mill. Pressed pellets were made from the fine powders
by adding 15–20 drops of an aqueous solution of polyvinyl Fig. 2. Th versus K2O diagram for Quebradagrande Complex volcanic
alcohol (Mowiol) al 15 tonnes/in.2 pressure in a steel die. rocks. Solid triangles, Group 1; stars, Group 2; solid circles, Group 3;
These pellets were analyzed for trace elements on Phillips solid diamonds, Group 4.
PW1400 X-ray spectrometers using Rh and W radiations.
Major elements were analyzed on the same spectrome-
ters using fusion discs for which the samples were prepared
according to the technique of Marsh et al. (1980). The fine
powder was ignited at 950 °C, then mixed with a flux of
lithium metaborate/lithium tetraborate (Johnson Mattey
Spectroflux 100B) at a 1:5 ratio. After fusion at 1250 °C,
the melt was pressed into an aluminium mould. The discs
were transferred to heated insulating blocks for annealing
at 200 °C overnight and kept in polypropylene bags.

3.3. Alteration

Given the altered nature of the rocks (by ocean floor


hydrothermal fluids and/or dynamic metamorphism), the
subsolidus mobility of the elements must be established
before any petrological inferences can be made from their
composition. Silica and alkalis (Na2O, K2O, Rb, and Ba)
are notoriously mobile during ocean floor metamorphism
Fig. 3. Total alkali versus silica diagram (Le Bas et al., 1986) for
(Stern and Elthon, 1979). The extent of mobility of the Quebradagrande volcanic rocks. Symbols as in Fig. 2. Shaded field defined
alkalis in relation to other less mobile elements can be by 186 samples of volcanic rocks from the Western Oceanic Cretaceous
ascertained by plotting a relatively immobile element such Lithospheric Province (Nivia, 1987; Kerr et al., 1997a,b).
428 A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436

latter elements were discarded for geochemical Compared with these samples, the Quebradagrande Com-
considerations. plex rocks display a greater range in SiO2 contents.
In an AFM diagram (Fig. 4), the Quebradagrande Com-
3.4. Major and trace element geochemistry plex samples straddle the tholeiitic–calk-alkaline boundary;
most follow a calk-alkaline differentiation trend, but some
Representative major and trace elements analysis of the evolved along a tholeiitic trend, as indicated by iron enrich-
Quebradagrande Complex samples appear in Table 2. Ana- ment. The field defined by the samples from the Western Oce-
lyzed samples were separated into four groups on the basis anic Cretaceous Lithospheric Province also indicates a
of their trace element characteristics. On a total alkali-silica difference in the geochemical evolution of the two provinces.
diagram (Le Bas et al., 1986), most samples plot in the In a variation diagram including Ti (Fig. 5), the higher degree
basaltic andesite field, but some also fall in the basalt, of differentiation compared with the field defined by the Wes-
mugearite, and andesite fields (SiO2 = 48.25–61.7 wt%), tern Oceanic Cretaceous Lithospheric Province’s samples
and one sample is a dacite (Fig. 3) whose SiO2 content contrast with the low TiO2 (1.3–0.42 wt%) concentration,
(66.35 wt%) might be modified by alteration, as indicated which shows that the Quebradagrande Complex samples
previously. The diagram suggests the field defined by 186 are not differentiated components of tholeiitic mid-ocean
samples from the Amaime and Volcanic formations of ridge or plateau series, which show higher Ti contents.
the Western Oceanic Cretaceous Lithospheric Province Fig. 6 shows primordial mantle-normalized multielement
reported by Nivia (1987) and Kerr et al. (1997b, 2001). diagrams (Sun and McDonough, 1989) for representative

Table 2
Representative XRF analyses of Quebradagrande Complex volcanic rocks
Sample QBG95-1 QBG95-3 AN1410A AN1425 AN1426 AN1409 AN1414 AN1412 AN1416
SiO2 52.03 52.48 49.29 53.54 48.25 57.64 61.72 52.04 54.11
Al2O3 20.77 17.11 19.59 16.8 18.42 16 16.53 15.67 18.13
Fe2O3a 9.33 11.52 13.07 11.7 13.16 11.91 7.76 8.01 10.07
MgO 3.33 5.13 7.16 7.93 7.97 5.22 2.77 10.54 4.02
CaO 10.83 8.57 8.33 7.31 9.79 2.07 6.06 9.73 9.15
Na2O 2.96 3.04 0.74 1.33 1.02 4.08 3.26 2.01 2.49
K2O 0.3 1.29 0.38 0.4 0.35 1.46 0.7 0.39 0.81
TiO2 0.66 0.78 1.05 0.6 0.74 0.96 0.82 1.03 0.85
MnO 0.18 0.16 0.08 0.14 0.19 0.18 0.18 0.13 0.21
P2O5 0.29 0.25 0.1 0.08 0.08 0.14 0.16 0.14 0.12
Total 100.67 100.34 99.94 100.02 100.12 99.94 100.07 100.01 100.14
LOI% 4.38 3.2 4.54 2.14 6.2 3.11 2.26 3.93 2.42
Trace elements in ppm
Ni 1.9 2.2 18.5 28.9 21.1 3.1 4.9 295.1 7.2
Cr 8 23.4 35.4 39.2 28.8 4 5.8 851.8 13.7
V 210.4 336 399.5 315.5 398 266.5 214 239.5 271
Sc 27.4 33.5 39.9 39.6 41 37.9 26.8 36.8 38.8
Cu 88.7 114.3 147.5 29.4 91.5 53.3 71.8
Zn 90.8 77.4 94.9 112.5 116.2 60.2 112.9
Ga 19 17.2 17.6 14.2 16.2 17.9 14.3 13.5 18.2
Pb 8.5 4.7 3.1 4.4 3.3 2.6 6.1 0.8 6.6
Sr 268.5 483.5 261.6 336.5 189.3 125.7 242.5 445.3 325.7
Rb 7 28.2 4 8.3 6.8 11.6 12.2 5.2 15.1
Ba 212 553.8 122.5 209.2 143.1 1658.7 175.8 199.2 443.3
Zr 63.1 86.7 46 19.4 27.7 67.8 71.8 72.6 64.3
Nb 0.7 1 1 2.5 0.9 1.3 4.7 2 1.4
Th 4.2 3.3 0.4 0.4 1.1 1.8 1.9 0.8 1
Y 17.6 17.3 17.9 11.4 12.4 29 24 23.2 26.7
La 8.6 9.4 1.9 2.2 2.9 6 5.7 2.6 3.6
Ce 18.7 22.8 8.7 7.9 10.7 16.4 15.1 9.3 11.7
Nd 13 15.1 6.5 4.6 6.3 11.8 10.7 10.7 9.3
Interelement selected ratios
La/Nb 12.3 9.4 1.9 0.9 3.2 4.6 1.2 1.3 2.6
Ba/Zr 3.4 6.4 2.7 10.8 5.2 24.5 2.4 2.7 6.9
CeN/YNb 2.9 3.7 1.3 1.9 2.4 1.6 1.7 1.1 1.2
Major elements recalculated to a volatile-free total.
a
Total iron reported as Fe2O3. LOI – Losses on ignition.
b
Chondrite-normalized values.
A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436 429

Group 1 possesses the highest LILE enrichments relative


to HFSE, as indicated by their highest Th contents (2.6–
4.2 ppm) and La/Nb ratio values (1–12.3) compared with
the other defined groups (Group 2: Th = 0.4–2.7, La/
Nb = 0.9–3.8; Group 3: Th = 1.3–1.9, La/Nb = 0.7–4.6;
Group 4: Th = 0.5–1.4, La/Nb = 0.8–4.4). Fig. 7 suggests
that Group 1 also displays the most enriched LREE pat-
terns and a general intergroup trend of increasing CeN/
YN ratios with increasing fractionation (Group 1 = 2.6–
3.8; Group 2 = 1–2.7; Group 3 = 1.4–1.7; Group 4 = 0.6–
1.3). According to the FeO* and TiO2 contents, Group 1
exhibits the most calc-alkaline behavior, as also is indicated
by its position on the AFM diagram (Fig. 4) and the
decreasing TiO2 content with increasing fractionation
(Group 1 = 0.42–0.78; Group 2 = 0.6–1.3; Group
Fig. 4. AFM diagram for Quebradagrande Complex volcanic rocks.
3 = 0.7–0.9; Group 4 = 0.6–1.3 wt%).
Symbols as in Figs. 2 and 3.

4. Petrogenesis

4.1. Fractional crystallization ‘‘Fe–Ti oxide arguments’’

Most of the analyzed Quebradagrande Complex sam-


ples seem to have evolved along two crystallization
trends: a calc-alkaline and a more tholeiitic (Fig. 4).
The occurrence of calc-alkaline characteristics supports
a supra-subduction zone environment of origin for the
Quebradagrande Complex. Although tholeiitic crystalliza-
tion trends are present in most environments where basic
Fig. 5. Th versus TiO2 diagram for Quebradagrande Complex volcanic volcanic igneous rocks are generated, calc-alkaline trends
rocks. Symbols as in Figs. 2 and 3.
are exclusive to magmatic environments with a conver-
gent margin (Wilson, 1987). Both trends are interpreted
samples of the Quebradagrande Complex. All the diagrams as the result of fractional crystallization of olivine, pla-
display LILE enrichment relative to HFSE, particularly gioclase, and clinopyroxene. The main difference between
Nb. Also, Ba, K, and Sr enrichments are conspicuous by the two trends is the control exerted over the Fe–Ti oxi-
their pronounced peaks in the diagrams that may reach, des during crystallization (Gill, 1981). Experimental evi-
as in the case of Ba and Sr, up to 237 and 58 times the dence (Grove and Baker, 1984) shows that subalkaline,
mantle-normalized concentrations (1659 ppm Ba and anhydrous magmas crystallizing in the crust under geo-
1228 ppm Sr). Conversely, the depletion in Nb is outstand- logically reasonable oxygen fugacity follow tholeiitic dif-
ing by the throat it displays in all diagrams. The LILE ferentiation trends. In hydrated basaltic magmas,
enrichment relative to HFSE is usually coupled with LREE dissolved water reduces olivine, pyroxene, and plagio-
enrichment relative to HREE. The degree of enrichment in clase stability without affecting the thermal stability of
LREE can be monitored using the chondrite-normalized Fe–Ti oxides (Sisson and Grove, 1993), which results in
(Nakamura, 1974) CeN/YN ratio that, in the Quebrada- the early crystallization of Fe–Ti oxides from a calc-alka-
grande Complex, varies between 1 and 4 times the chon- line magmatic system and lower Fe and Ti contents in
dritic values, suggesting flat to enriched REE patterns the resultant magmas than in a tholeiitic crystallization
(Fig. 7). sequence where the fractionation of Fe–Ti oxide is
According to the shape of their multielement-normal- delayed, which increases the Fe and Ti contents of the
ized pattern, the samples can be divided into four different magma. Thus, water activity influences the early or late
groups (Fig. 6). This separation is based mainly on inter- crystallization of titaniferous magnetite, and subalkaline
HFSE ratios, which we believe represent inherited features magmas with high water contents follow calc-alkaline
from the source of the magmas. For example, Group 1 is differentiation trends, whereas those with low water con-
characterized by its low Y/P ratio values (Group 1 = 0.2– tents display tholeiitic trends.
0.5; Group 2 = 0.3–0.9; Group 3 = 0.7–1; Group 4 = 0.6–
1.1), and Group 2 is distinguished by its high Ti/Zr ratios 4.2. Trace element arguments
(Group 1 = 0.3–0.7; Group 2 = 1–2.1; Group 3 = 0.7–1;
Group 4 = 0.7–1.2). However, groups are also homoge- The LILE enrichment in volcanic rocks can be produced
neous in their LILE and major element characteristics. by several processes, such as oceanic mantle contaminated
430 A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436

Fig. 6. Normalized multielement plot of Quebradagrande Complex volcanic rocks. Normalizing values from Sun and McDonough (1989).

by deeper, uncirculated mantle plumes; low percentages of


fusion in the source; mantle metasomatism by subducting
plate-derived fluids; or contamination. The characteristic
feature of magmas originated from supra-subduction zone
mantle wedges is the LILE enrichment relative to HFSE
(Saunders et al., 1980; Tarney et al., 1981; Saunders and
Tarney, 1982; Pearce, 1983). This feature, produced by
chemical fractionation between LILE and HFSE in the
hydrous environment associated with subduction zones, is
outstanding in the pronounced Nb negative anomaly dis-
played by the Quebradagrande Complex in the spider-
grams. The strong chemical behavior affinity (i.e., similar
incompatibility with the original mantle and basic volcanic
rocks mineralogies) of elements like La (LILE), Nb, and Ta
Fig. 7. CeN/YN versus Th diagram for Quebradagrande Complex (HFSE) is well known, such that these elements usually are
volcanic rocks. Symbols as in Fig. 2. placed at adjacent positions in multielement-normalized
A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436 431

diagrams (Bougault et al., 1985). These elements are not 5. Discussion


fractionated by processes like partial fusion or fractional
crystallization. However, metamorphic processes acting On the basis of major element chemical analyses, Gon-
on the subduction zone produce dehydration of the sub- zález (1980a) proposes that the Quebradagrande Complex
ducting oceanic crust with consequent production of is composed of tholeiitic rocks generated in an oceanic rift.
hydrous supercritical fluids that transport water and solu- Bourgois et al. (1982, 1985, 1987) and Toussaint and Rest-
ble LILE (e.g., La) to the supra-subduction zone mantle repo (1993) propose the same origin but also suggest that
wedge (Saunders et al., 1980; Tarney et al., 1981). The the Quebradagrande Complex had been thrust from the
HFSE (e.g., Nb) can be retained by the subducting oceanic west onto the continental margin, along with basic rocks
crust (then an eclogitic assemblage) and returned to the that outcrop west of the Cauca-Almaguer fault (i.e., the
mantle (Saunders and Tarney, 1982). The latter can be Western Oceanic Cretaceous Lithospheric Province).
ascribed to the hydrous environment in the subduction According to this model, a single common mantle source
zone, which favors the stability of mineral phases such as region is responsible for both the Quebradagrande Com-
rutile or ilmenite that retain HFSE (Tarney et al., 1981). plex and the volcanic rocks of the Western Oceanic Creta-
Fig. 8 shows a plot of La/Y versus Nb/Y (La and Nb nor- ceous Lithospheric Province (Barroso, Amaime, Volcanic
malized to Y to eliminate the effects of fractional crystalli- formations, Table 1). Similarly, Kammer (1995) and Kam-
zation) for the Quebradagrande Complex samples. This mer and Mojica (1996) consider a close affinity between the
diagram is useful for separating samples that contain a sub- rocks of the accreted Western Oceanic Cretaceous Litho-
duction-related component (low La/Nb ratios) from those spheric Province and the Quebradagrande Complex. The
derived from an oceanic mantle source region, far from the results presented herein clearly show a subduction zone-de-
influence of a subduction zone (e.g., mid-ocean ridge bas- rived component in the geochemical composition of the
alts, mantle plume-generated oceanic plateau basalts). This Quebradagrande Complex. However, the geochemical
diagram also shows fields encompassing the basalts from characteristics of the Quebradagrande Complex differ fun-
the Amaime and Volcanic formations of the Western Oce- damentally from those of the basic volcanic rocks of the
anic Cretaceous Lithospheric Province (Nivia, 1987; Kerr Western Oceanic Cretaceous Lithospheric Province
et al., 1997b, 2001) and lavas of the recent Andean volca- exposed in western Colombia (Millward et al., 1984; Nivia,
noes of Colombia (Marriner and Millward, 1984). The 1987, 1989; Kerr et al., 1996, 1997a,b). In addition, the
environment of formation for the former rocks was unre- reported ages of the Western Oceanic Cretaceous Litho-
lated to a subduction zone (Millward et al., 1984; Nivia, spheric Province indicate these rocks were formed during
1987; Kerr et al., 1996, 1997a,b, 2001), whereas the latter the Late Cretaceous, whereas evidence for the Quebrada-
are modern examples of rocks produced in a destructive grande Complex indicates an Early Cretaceous age. Thus,
plate margin. In this diagram, the chemical differences the volcanic rocks of the Quebradagrande Complex and
between the basalts of the Western Oceanic Cretaceous the Western Oceanic Cretaceous Lithospheric Province
Lithospheric Province and the subduction-related Quebra- were generated from two different, unrelated mantle source
dagrande Complex are evident. It is thus highly unlikely regions; in turn, the geotectonic models and interpretations
that the Quebradagrande Complex lavas and pyroclastics of a common genesis for both are inherently incorrect.
represent part of the same tectonomagmatic event. Geochemical data suggest that the volcanic rocks of the
Quebradagrande Complex formed during the Lower Creta-
ceous in a supra-subduction zone environment, in an island
arc, marginal basin, or active continental margin. To eval-
uate these possibilities, the regional relationships between
the Quebradagrande Complex and adjacent rocks must
be considered. The Quebradagrande Complex outcrops
are bounded by the metamorphic Cajamarca and Arquı́a
complexes (Fig. 1, Table 1). The age of the Cajamarca
Complex, to the east, is constrained as Neoproterozoic
(Gómez and Nuñez, 2003), but the age of the Arquı́a Com-
plex to the west is controversial. Some researchers believe
the Arquı́a Complex was formed during the Lower Creta-
ceous (Toussaint and Restrepo, 1989, 1993; Restrepo et al.,
1991; González and Nuñez, 1991; González, 1993, 2001),
whereas others think it Paleozoic in age (McCourt and
Aspden, 1983; McCourt et al., 1984; Aspden et al., 1987).
However, both the presence of Triassic granitoid plutons,
such as the Santa Barbara Batholith and Amaga and Cam-
Fig. 8. Nb/Y versus La/Y diagram for Quebradagrande Complex bumbia stocks, intruding schists west of the Quebrada-
volcanic rocks. Symbols as in Fig. 2. grande Complex (Fig. 1; McCourt et al., 1984; Mejı́a
432 A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436

et al., 1983b; Restrepo et al., 1991) and the occurrence of To explain the Quebradagrande Complex genesis, we
Paleozoic rocks west of the Quebradagrande Complex propose a petrogenetic model in which the rocks formed
(Mosquera, 1978; Calle et al., 1980; Mejı́a et al., 1983a,b; during the opening of a marginal ensialic basin. Åberg
González, 2001) are strong arguments in favor of a Neo- et al. (1984) and Aguirre (1987) propose that these basins
proterozoic age for the Arquı́a Complex. Assuming a might form by ensialic expansion or subsidence mecha-
Paleozoic age for the Arquı́a Complex, McCourt et al. nisms promoted by the rollback action of the subducting
(1984) present an evolutionary model for the northern lithosphere on the active continental margin. These mech-
Andes. According to this model, during the Carbonifer- anisms may lead to crustal thinning with consequent adia-
ous, the continental margin of the northern Andes was batic decompression melting of the mantle, which
composed of an autochthonous block (Cajamarca Com- produced both basins and magmatism. The volcanic prod-
plex) and accreted island arc (Bugalagrande schists, ucts erupt over the continental crust or, in cases of extreme
Rosario amphibolites, and Bolo Azul metagabbroids; extension, result in the complete rupture of the continental
Arquı́a Complex, Table 1). We concur with this model margin and generation of oceanic crust (Åberg et al., 1984;
and believe that the data presented here support a model Aguirre, 1987; Atherton and Aguirre, 1992).
in which the Quebradagrande Complex is related to a Ophiolitic sequences proposed as generated by these
magmatic environment associated with a continental processes are common to the Uppermost Jurassic and
active margin. Lower Cretaceous of the South American Pacific margin,
Processes operating in subduction zones result in dehy- extending from its southernmost extreme in South Geor-
dration of the subducting oceanic crust and transport of gia Island (Chile) to central-western Mexico; the latter is
generated fluids into the overlying mantle wedge. The continuous with eastern Colombia in most Triassic–Ju-
influx of such volatile-rich fluids into the mantle wedge rassic Pangea reconstructions (Coney and Evenchick,
lowers the solidus of the mantle, resulting in melting 1994). Well-documented ensialic marginal basins are
and the formation of hydrated magmas. However, the found in the ophiolitic complexes of the Larsen Peninsu-
water (volatile) content of the magmas may vary accord- la, Tortuga and Sarmiento at South Georgia Island,
ing to the distance between the centers of volcanic erup- Tierra de Fuego, and Patagonia (Tarney et al., 1981;
tion and the trench. For marginal basins, it depends on Saunders and Tarney, 1982; Miller et al., 1994; Stern
the degree of evolution, that is, the development of the et al., 1976); the Rocas Verdes of central and southern
basin. As marginal basins (sensu stricto) open and Chile (Åberg et al., 1984); the Puente Piedra and Casma
the expansion center separates from the island arc, the formations in Peru (Atherton et al., 1985; Aguirre and
hydrous (subduction) component becomes less pro- Offer, 1985; Atherton and Aguirre, 1992); the Célica For-
nounced (Saunders and Tarney, 1984; Atherton and Agu- mation of southern Ecuador and northern Peru (Aguirre
irre, 1992). In this way, the calc-alkaline and tholeiitic and Atherton, 1987; Lebras et al., 1987); and the inter-
trends and the LILE/HFSE ratio in the Quebradagrande bedded calc-alkaline volcanics and sedimentary strata
Complex could be controlled by the preeruptive contents containing shallow marine fauna and imbricated with
of the hydrous component in the mantle. As we noted serpentinite bodies, ultramafic cumulates, and podiform
previously, the samples can be separated into four differ- chromite of the Siuna Terrane of Nicaragua and Hondu-
ent chemical types (Fig. 6), which may be interpreted as ras, which extends northward into the Guerrero Terrane
heterogeneities produced in the mantle during the evolu- of southern Mexico. In the latter, Elias-Herrera and Ort-
tion of the basin. Thus, Group 1 samples, which appear ega-Gutierrez (1998) posit the upper volcanoclastic
to contain more of the subduction component(s) (higher sequence was generated in a continental margin backarc
LILE/HFSE ratios), could have been generated during basin setting. The spatial, chronological characteristics
the initial stages of basin opening. During latter stages and geochemical composition of the Quebradagrande
of basin opening, progressive dilution of this component Complex suggest they belong to this belt.
in the source region, as reflected in the lower LILE/HFSE The petrogenetic model proposed for the origin of the
ratios of Groups 2 and 3, culminates in the Group 4 sam- Quebradagrande Complex is illustrated in Fig. 9. In this
ples, which display the lowest La/Nb and CeN/YN ratios model, subduction along the proto-Pacific Colombian mar-
(Figs. 7 and 8). Parallel trends between groups in bivari- gin produced distension and crustal thinning over the con-
ate diagrams for the same degree of differentiation favor tinental margin. The thinning resulted in the formation of a
this hypothesis. Furthermore, increasing values of the basin, as evidenced by the accumulation of transgressive
CeN/YN ratios with increasing Th in a Th versus CeN/ sedimentary sequences (e.g., Valle Alto, Abejorral; Table
YN diagram (Fig. 7) could be related to intersample var- 1). The genesis of these arenorudaceous clastic sequences
iation due to differentiation processes, whereas parallel would agree with the ensialic marginal basin model if they
intergroup trends suggest different LILE enrichment in represent and show the first stages of basin opening. On the
the mantle. However, the strong deformation of the oceanic plate, the progressive increase of temperature and
Quebradagrande Complex inhibits any reconstruction of pressure as subduction processes produced metamorphism
the basin that might help evaluate this hypothesis in terms with consequent dehydration. The corresponding subduc-
of geochemical characteristics. tion-liberated fluids moved upward, adding H2O, LILE,
A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436 433

Fig. 9. Diagrammatic sketch illustrating evolution of the Quebradagrande Complex–QGC marginal ensialic basin. (A) Subduction and consequent
metasomatism of subcontinental mantle under Paleozoic crust of an oceanic crustal block (Arquı́a Complex) accreted to the continental crustal block
(Cajamarca Complex + shield). (B) Rollback of the continental margin leads to lithospheric thinning and subsequent generation of basins and adiabatic
mantle melting. (C) Formation of marginal basin by generation of oceanic floor in the zone of backarc spreading. (D) Closure of basin, probably due to
forces produced on the continental plate during the aperture of the South Atlantic.
434 A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436

and LREE to the subcontinental mantle wedge. The addi- are well documented to have formed in an oceanic pla-
tion of these fluids lowered the solidus; combined with adi- teau setting.
abatic decompression induced by crustal thinning, it
resulted in mantle melting. The magmas produced were References
calc-alkaline basalts that differentiated at crustal levels to
form andesites and dacites (Quebradagrande Complex Åberg, G., Aguirre, L., Levi, V., Nystrom, J.O., 1984. Spreading-
Groups 1 and 2). Basin opening culminated in the genera- subsidence and generation of ensialic marginal basins: an example
tion of oceanic crust, represented today by ophiolitic com- from the early Cretaceous of central Chile. In: Kokelaar, B.P., Howells,
M.F., Roach, R.A. (Eds.), Volcanic Processes in Marginal Basins:
plexes. According to this petrogenetic model, more basic
Geological Society of London Special Publication 16, pp. 185–193.
volcanic rocks (Groups 2 and 3) and sedimentary rocks Aguirre, L., 1987. Andean modelling. Geology Today 3, 47–48.
of the Quebradagrande Complex accumulated on top of Aguirre, L., Atherton, M.P., 1987. Low grade metamorphism and
the basin, whereas the associated mafic and ultramafic plu- geotectonic setting of the Macuchi Formation, Western Cordillera of
tonic rocks (Table 1) represent deeper horizons of the Ecuador. Journal of Metamorphic Geology 5, 473–494.
Aguirre, L., Offer, R., 1985. Burial metamorphism in the Western
ophiolitic complexes developed by ocean floor formation
Peruvian Trough: its relation to Andean magmatism and tectonics.
during the opening of the basin. In: Pitcher, W.S., Atherton, M.P., Cobbing, E.J., Beckinsale, R.D.
The opening of the basin led to the gradual movement of (Eds.), Magmatism at a Plate Edge, the Peruvian Andes. Blackie, John
the locus of magmatism away from the trench, which may Willey and Sons, pp. 59–71.
have resulted in the dilution of the subduction zone-derived Aspden, J.A., McCourt, W.J., 1986. Mesozoic oceanic terrane in the
Central Andes of Colombia. Geology 14, 415–418.
component and a change in the magmatic products from
Aspden, J.A., McCourt, W.J., Brook, M., 1987. Geometrical control on
calc-alkaline to tholeiitic. According to Saunders and Tar- subduction-related magmatism: the Mesozoic and Cenozoic plutonic
ney (1984), marginal basins are short-life geotectonic fea- history of Western Colombia. Journal of the Geological Society,
tures. Furthermore, Dalziel (1981) suggests that in London 144, 893–905.
marginal basins of southern Chile, collapse or closure coin- Atherton, M.P., Aguirre, L., 1992. Thermal and geotectonic setting of
Cretaceous volcanic rocks near Ica, Peru, in relation to Andean crustal
cides with the opening of the South Atlantic. Thus, the
thinning. Journal of South American Earth Sciences 5, 47–69.
change in the velocity of plate displacement induced by Atherton, M.P., Pitcher, W.S., Warden, V., 1983. The Mesozoic marginal
the South Atlantic opening also may have promoted the basin of central Peru. Nature 305, 303–306.
closure of the Quebradagrande Complex marginal basin. Atherton, M.P., Warden, V., Sanderson, M., 1985. The Mesozoic
However, the accretion of the Caribbean–Colombian oce- marginal basin of Central Peru: a geochemical study of within-plate-
edge volcanism. In: Pitcher, W.S., Atherton, M.P., Cobbing, E.J.,
anic plateau in the late Cretaceous–early Tertiary likely
Beckinsale, R.D. (Eds.), Magmatism at a Plate Edge, the Peruvian
contributed significantly to the closure of the basin and Andes. John Willey and Sons, Blackie, pp. 47–58.
deformation of this unit. Barrero, D., 1979. Geology of the central Western Cordillera, West of
Buga and Roldanillo, Colombia. Publicaciones Geológicas Especiales
6. Conclusions de INGEOMINAS 4, 75.
Botero, A., 1963. Contribución al conocimiento de la geologı́a de la zona
central de Antı́oquia. Anales Facultad de Minas (Medellı́n) 57, 101.
Regional sampling of the Quebradagrande Complex Bougault, H., Joron, J.L., Treuil, M. Maury, R., 1985. Local versus
between 6°35 0 N and 1°45 0 N in Colombia shows that basal- regional mantle heterogeneities: evidence from hygromagmatophile
tic andesites and andesites have geochemical characteristics elements. In: Bougault, H. and Cande, S.C. (Eds.), Initial Reports of
typical of rocks formed in supra-subduction zone magmat- the Deep Sea Drilling Project, U.S. Government Printing Office.
Washington, 82, pp. 459–477.
ic environments. These rocks follow two contrasting differ-
Bourgois, J., Calle, B., Tourmon, J., Toissaint, J.F., 1982. The Andean
entiation trends: One is calc-alkaline, the other tholeiitic. ophiolitic megastructure on the Buga-Buenaventura transverse (Wes-
The geochemical characteristics of the Quebradagrande tern Cordillera – Valle, Colombia). Tectonophysics 82, 207–229.
Complex rocks and their spatial and chronological rela- Bourgois, J., Toussaint, J.F., Gonzáles, H., Azema, J., Calle, B., Desmet,
tionships with the Cajamarca and Arquı́a complexes, the A., Murcia, L.A., Alvarado, P., Parra, E., Tourmon, J., 1987.
Geological history of the Cretaceous ophiolitic complexes of northern
ultramafic and mafic Cretaceous rocks, and the arenoruda-
South America (Colombian Andes). Tectonophysics 143, 307–327.
ceous Lower Cretaceous related sequences can be integrat- Bourgois, J., Toussaint, J.F., Gonzales, H., Orrego, A., Azema, J., Calle,
ed into an evolutionary petrogenetic model of ensialic B., Desmet, A., Murcia, A., Alvarado, P., Parra, E. Tourmon, J., 1985.
marginal basin opening. Les ophiolites des Andes de Colombie. Evolution Structural et
The model has global tectonic implications, in that it signification geodinamic, In: Mascle, A. (Ed.), Geodinamic des
Caraibbes, Symposium: Technip. Paris, pp. 475–493.
forms an important link, through the northern Andes, of
Bürgl, H., Radelli, L., 1962. Nuevas localidades fosilı́feras en la Cordillera
the chain of marginal basins that extended from Tierra Central de Colombia (SA.). Geologı́a Colombiana 3, 133–138.
de Fuego to Mexico in the Early Cretaceous. During this Calle, B., González, H., De La Peña, R., Escorce, E., Durango, J.,
stratigraphic interval, subduction was active along the Ramı́rez, O., Alvarez, E., Calderón, M., Alvarez, J., Guarı́n, G.
South American margin. Rodrı́guez, C., Muñosz, J. Durán, J., 1980. Mapa Geológico de
Colombia – Escala 1:100.000, Plancha 166 – Jericó: INGEOMINAS.
The results we present clearly demonstrate that there
Bogotá.
is no genetic relationship between the Quebradagrande Coney, P.J., Evenchick, C.A., 1994. Consolidation of the American
Complex and the Upper Cretaceous volcanic rocks that Cordilleras. Fifth Circum-Pacific Terrane Conference (Santiago).
outcrop west of the Cauca-Almaguer fault – rocks that Pergamon. pp. 241–262.
A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436 435

Dalziel, I.W.D., 1981. Back-arc extension in the southern Andes: a review Lebras, M., Mégard, F., Dupuy, C., Dostal, J., 1987. Geochemistry and
and critical reappraisal. Philosophical Transactions of the Royal tectonic setting of pre-collision Cretaceous and Paleogene volcanic
Society of London 300, 319–335. rocks of Ecuador. Geological Society of America Bulletin 99, 569–578.
Dalziel, I.W.D., de Wit, M.J., Palmer, K.F., 1974. Fossil marginal basin in Lozano, H., Perez, H., Mosquera, D., 1984a. Prospección geoquı́mica
the southern Andes. Nature 215, 291–294. para oro, plata, antimonio y mercurio en los municipios de Salento,
Elias-Herrera, M., Ortega-Gutierrez, F., 1998. The Early Cretaceous Qundio y Cajamarca, Tolima. Boletı́n Geológico INGEOMINAS 27,
Arperos oceanic basin (western Mexico), Geochemical evidence for an 5–76.
aseismic ridge formed near a spreading center – comment. Tectono- Lozano, H., Perez, H., Vesga, C., 1984b. Prospección geoquı́mica y
physics 292 (3–4), 321–326. génesis del mercurio en el flanco occidental de la Cordillera Central de
Etayo, F., 1985. Documentacion paleontológica del Infracretácico de San Colombia, Municipios de Aranzazu, Salamina y Pacora, Departamen-
Félix y Valle Alto, Cordillera Central, Proyecto Cretácico. Publicac- to de Caldas. Boletı́n Geológico INGEOMINAS 27, 77–169.
iones Geológicas Especiales del INGEOMINAS 16, XXV1–XXV7. Marriner, G.F., Millward, D., 1984. The petrology and geochemistry of
Gill, J., 1981. Orogenic Andesites and Plate Tectonics:. Springer Verlag, Cretaceous to Recent volcanism in Colombia. Geological Society of
Berlin, 390 p. London 141, 473–486.
Gómez, A., Moreno, M., Pardo, A., 1995. Edad y origen de Complejo Marsh, N.G., Saunders, A.D., Tarney, J., Dick, H.J., 1980. Geochemistry
metasedimentario de Aranzazu-Manizales en los alrededores de of basalts from the Shikoku and Daito Basins, Deep Sea Drilling
Manizales (Departamento de Caldas, Colombia). Geologı́a Colombi- Project leg 58. Initial Reports of the Deep Sea Drilling Project. U.S.
ana 19, 83–93. Government Printing Office, Washington, 58, pp. 805–842.
Gómez, J., Nuñez, A., 2003. Las metasedimentitas de Santa Teresa y la Maya, M., Gonzalez, H., 1996. Unidades litodémicas en la Cordillera
edad del Complejo Cajamarca (Cordillera Central, Departamento del Central de Colombia. Boletı́n Geológico INGEOMINAS 35, 43–57.
Tolima – Colombia). IX Congreso Colombiano de Geologı́a, Resum- McCourt, W.J., Aspden, J.A., 1983. A plate tectonic model for the
enes, Medellı́n, 35 p. phanerozoic evolution of central and southern Colombia. In: 10th
González, H., Nuñez, A., 1991. Mapa Geológico Generalizado del Caribbean Geological Conference Transactions: INGEOMINAS, pp.
Departamento del Quindio, INGEOMINAS, 42 p. 38–47.
González, H., 1980a. Geologı́a de las planchas 167 (Sonsón) y 187 McCourt, W.J., Aspden, J.A., Brook, M., 1984. New geological and
(Salamina), Boletı́n Geológico INGEOMINAS, 23, 174 p. geochronological data from the Colombian Andes: continental growth
González, H., 1980b. Mapa Geológico de Colombia – Escala 1:100.000, by multiple accretion. Journal of the Geological Society, London 141,
Plancha 167 – Sonsón, INGEOMINAS. Bogotá. 835–841.
González, H., 1980c. Mapa Geológico de Colombia – Escala 1:100.000, Mejı́a, M., Alvarez, E., González, H., Grosse, E., 1983a. Mapa Geológico
Plancha 187 – Salamina, INGEOMINAS. Bogotá. de Colombia – Escala 1:100.000, Plancha 130 – Santa Fé de Antioquia.
González, H., 1993. Mapa Geológico del Departamento de Caldas, INGEOMINAS. Bogotá.
INGEOMINAS, 62 p. Mejı́a, M., Alvarez, E., González, H., Grosse, E., 1983b. Mapa Geológico
González, H., (2001). Geologı́a de las planchas 206, Manizales y 225, de Colombia - Escala 1:100.000, Plancha 146 - Medellı́n Occidental.
Nevado del Ruiz, Escala 1:100.000, Memoria explicativa. INGEOM- INGEOMINAS. Bogotá.
INAS, Bogotá, 92 p. Miller, C.A., Barton, M., Hanson, R.E., Fleming, T.H., 1994. An Early
Grove, T.L., Baker, M.B., 1984. Phase equilibrium controls on the Cretaceous volcanic arc/marginal basin transition zone, Peninsula
tholeiitic versus calc-alkaline differentiation trends. Journal of Geo- Hardy, southernmost Chile. Journal of Volcanology and Geothermal
physical Research 89, 3253–3274. Research 63, 33–58.
Hall, R.B., Alvarez, J., Rico, H., 1972. Geologı́a de parte de los Millward, D., Marriner, G., Saunders, A.D., 1984. Cretaceous tholeiitic
departamentos de Antioquia y Caldas (sub-zona II-A), Boletı́n volcanic rocks from the Western Cordillera of Colombia. Journal of
Geológico INGEOMINAS. XX, 85 p. the Geological Society, London 141, 847–860.
Kammer, A., 1995. Las fallas de Romeral y su relación con la tectónica de Molnar, P., Atwater, T., 1978. Interarc spreading and cordilleran tectonics
la Cordillera Central. Geologı́a Colombiana 18, 27–46. as alternates related to the age of subducted oceanic lithosphere. Earth
Kammer, A., Mojica, J., 1996. Una comparación de la tectónica de and Planetary Science Letters 41, 330–340.
basamento de las cordilleras Central y Oriental. Geologı́a Colombiana Mosquera, D., 1978. Geologı́a del Cuadrangulo K-8. Informe 1763
20, 93–106. (unpublished): INGEOMINAS. Bogotá. 63 p.
Kerr, A.C., Tarney, J., Marriner, G.F., Nivia, A., Saunders, A.D., Klaver, Nakamura, N., 1974. Determination of REE, Ba, Fe, Mg, Na and K in
G.T., 1996. The geochemistry and tectonic setting of late Cretaceous carbonaceous and ordinary chondrites. Geochimica et Cosmochimica
Caribbean and Colombian volcanism. Journal of South American Acta 38, 757–775.
Earth Sciences 9, 111–120. Nelson, H.W., 1957. Contribution to the geology of the Central and
Kerr, A.C., Marriner, G.F., Tarney, J., Nivia, A., Saunders, A.D., Western Cordillera of Colombia in the sector between Ibague and Cali.
Thirlwall, M.F., Sinton, C., 1997a. Cretaceous Basaltic Terranes in Leidse Geologische Medelelingen 22, 1–76.
Western Colombia: elemental, chronological and Sr–Nd isotopic Nelson, H.W., 1962. Contribución al conocimiento de la Cordillera
constraints on Petrogenesis. Journal of Petrology 38, 677–702. Central de Colombia. Seccion entre Ibagué y Armenia. Boletı́n
Kerr, A.C., Tarney, J., Marriner, G.F., Nivia, A., Saunders, A.D., 1997b. Geológico INGEOMINAS, X, pp. 161–203.
The Caribbean–Colombian Cretaceous igneous province: the internal Nivia, A., 1987. Geochemistry and origin of the Amaime and Volcanic
anatomy of an oceanic plateau, In: Mahoney, J.J., Coffin, M. (Eds.), Sequences, Southwestern Colombia: Unpublished Master of Philoso-
Large Igneous Provinces: Continental, Oceanic, and Planetary Flood phy thesis. University of Leicester, Leicester, UK, 163 p.
Volcanism, American Geophysical Union, Geophysical Monograph Nivia, A., 1989. El Terreno Amaime-Volcánica una provincia acrecionada
100, pp. 123–144. de basaltos de meseta oceánica. In: V Congreso Colombiano de
Kerr, A.C., Tarney, J., Kempton, P.D., Spadea, P., Nivia, A., Marriner, Geologı́a, Memorias, I, pp. 1–30.
G.F., Duncan, R., 2001. Pervasive mantle plume head heterogeneity: Nivia, A., Gálvis, N. Maya, M., 1997. Geologı́a de la Plancha 242 –
evidence from the late Cretaceous Caribbean–Colombian oceanic Zarzal. INGEOMINAS, Bogotá. 73 p.
plateau. Journal of Geophysical Research 107/B7. doi:10.1029/ North American Commission on Stratigraphic Nomenclature, 1983.
2001JB000790. North American Stratigraphic Code. American Association of Petro-
Le Bas, M.J., Le Maitre, R.W., Streckeisen, A., Zanettin, B., 1986. A leum Geologist Bulletin, 67/5, pp. 841–875.
chemical classification of volcanic rocks based on the total alkali-silica Orrego, A., 1993. Geologı́a de la Plancha 364-Timbı́o. INGEOMINAS.
diagram. Journal of Petrology 27, 745–750. 36 p.
436 A. Nivia et al. / Journal of South American Earth Sciences 21 (2006) 423–436

Orrego, A., Rossman, D., Paris, G., 1976. Geologı́a del Cuadrangulo N-6 implication for ocean floor metamorphism, seismic layering, and
Popayán. Informe 1711 (unpublished): INGEOMINAS. Bogotá. 179 magnetism. Journal of Geophysical Research 81, 4370–4380.
p. Stern, C.R., Elthon, D., 1979. Vertical variations in the effects of
Paris G., Marı́n, P.A., 1979. Generalidades acerca de la geologı́a del hydrothermal metamorphism in the Chilean ophiolites: their
Departamento del Cauca. INGEOMINAS. Bogotá. 38 p. implications for ocean floor metamorphism. Tectonophysics 55,
Pearce, J.A., 1983. The role of sub-continental lithosphere in magma 179–213.
genesis at active continental margins. In: Hawkesworth, C.J., Norry, Sun, S.S., McDonough, W.F., 1989. Chemical and isotopic systematics
M.J. (Eds.), Continental Basalts and Mantle Xenoliths. Shiva, of oceanic basalts: implications for mantle composition and
Nantwich, UK, pp. 230–249. processes, In: Saunders, A.D., Norry, M.J. (Eds.), Magmatism in
Restrepo, J.J., Toussaint, J.F., González, H., Cordani, U., Kawashita, K., the Ocean Basins, Geological Society of London Special Publication,
Linares, E., Parica, C., 1991. Precisiones geocronológicas sobre el 42, pp. 313–345.
Occidente Colombiano. Memorias Simposio sobre Magmatismo Tarney, J., Saunders, A.D., Mattey, D.P., Wood, D.A., Marsh, N.G.,
Andino y su Marco Tectónico I, 1–22. 1981. Geochemical aspects of back-arc spreading in the Scotia Sea and
Rodrı́guez, C., Rojas, R., 1985. Estratigrafı́a y tectónica de la Serie Western Pacific. Philosophical Transactions of the Royal Society of
Infracretácica en los alrededores de San Felix, Cordillera Central de London A300, 263–285.
Colombia. Publicaciones Geológicas Especiales del INGEOMINAS Toussaint, J.F., Restrepo, J.J., 1974, Algunas consideraciones sobre la
16, XXI1–XXI21. evolución structural de los Andes Colombianos. Publicaciones Espec-
Saunders, A.D., Tarney, J., 1982. Igneous activity in the southern Andes iales de Geologı́a, 4: Facultad Nacional de Minas, Medellı́n, 17 p.
and northern Antarctic Peninsula: a review. Journal of the Geological Toussaint, J.F., Restrepo, J.J., 1978. Edad K–Ar de dos rocas básicas del
Society, London 139, 691–700. flanco noroccidental de la Cordillera Central. Publicaciones Especiales
Saunders, A.D., Tarney, J., 1984. Geochemical characteristics of basaltic de Geologı́a, 15: Facultad de Ciencias, Medellı́n.
volcanism within back-arc basins, In: Kokelaar, B.P., Howells, M.F., Toussaint, J.F., Restrepo, J.J., 1989. Acreciones sucesivas en Colombia:
Roach, R.A. (Eds.), Volcanic Processes in Marginal Basins, Geological Un nuevo modelo de evolución geológica. Memorias V Congreso
Society of London Special Publication, 16, pp. 59–76. Colombiano de Geologia I, 127–147.
Saunders, A.D., Tarney, J., Weaver, S.D., 1980. Transverse geochemical Toussaint, J.F., Restrepo, J.J., 1993. Tectónica de terrenos durante el
variations across the Antarctic Peninsula: Implications for the genesis Cretácico en Colombia. Memorias VI Congreso Colombiano de
of calc-alkaline magmas. Earth and Planetary Science Letters 46, 344– Geologı́a I, 97–114.
360. Wilson, M., 1987. Igneous Petrogenesis: A Global Tectonics Approach.
Sisson, T.W., Grove, T.L., 1993. Experimental investigations of the role of Harper Collins Academic, London, 466 p.
H2O in calc-alkaline differentiation and subduction zone magmatism. Wood, D.A., Joron, J.L., Treuil, M., Norry, M.J., Tarney, J., 1979.
Contributions to Mineralogy and Petrology 113, 146–166. Elemental and Sr isotope variations in basic lavas from Iceland and the
Stern, C.R., De Witt, M.J., Lawrence, J., 1976. Igneous and metamorphic surrounding ocean floor: the nature of mantle source inhomogenities.
processes associated with the formation of Chilean ophiolites and their Contributions to Mineralogy and Petrology 70, 319–339.

View publication stats

You might also like