Sie sind auf Seite 1von 26

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

PAGES 677702

1997

Cretaceous Basaltic Terranes in Western Colombia: Elemental, Chronological and SrNd Isotopic Constraints on Petrogenesis
A. C. KERR1, G. F. MARRINER2, J. TARNEY1, A. NIVIA3, A. D. SAUNDERS1, M. F. THIRLWALL2 AND C. W. SINTON4
1 2 3 4

DEPARTMENT OF GEOLOGY, UNIVERSITY OF LEICESTER, UNIVERSITY ROAD, LEICESTER LE1 7RH, UK DEPARTMENT OF GEOLOGY, ROYAL HOLLOWAY UNIVERSITY OF LONDON, EGHAM TW20 0EX, UK INGEOMINASREGIONAL PACIFICO, AA 9724, CALI, COLOMBIA COLLEGE OF OCEANOGRAPHY, OREGON STATE UNIVERSITY, CORVALLIS, OR 97331, USA

RECEIVED AUGUST 19, 1996 REVISED TYPESCRIPT ACCEPTED JANUARY 10, 1997

Accreted terranes comprising Mid to Late Cretaceous picrites, basalts and dolerites occur in three northsouth trending belts in western Colombia, in the Central Cordillera, Western Cordillera and along the Pacic coast. The geochemistry of these rocks is consistent with an oceanic plateau (plume-related) origin, and they most probably formed in the Pacic as part of the Caribbean oceanic plateau. These igneous rocks display small but signicant inter-cordillera variations, being younger and more depleted in incompatible trace element ratios (and with more positive Nd values) to the west. The igneous rocks of the Pacic coast (Serrana de Baudo) are dated at 7378 Ma ( 40Ar/39Ar), and those of the Western Cordillera at ~90 Ma, whereas the volcanics of the Central Cordillera are believed to be older than 100 Ma. Most of the igneous rocks are basaltic, and it is suggested that they have fractionated from picritic primary magmas, generated by partial melting within a hot mantle plume. Variable and positive Nd values reveal that the plume must have been heterogeneous, originating from a mantle source with a long-term history of depletion. Partial melt modelling suggests that the composition of the basalts requires at least some input from a mantle source region containing garnet and that the extent of partial melting required to reproduce the composition of the erupted basalts is of the order of ~20%. Mixing of melts from dierent depths, either in the mantle melting column or during fractionation in lithospheric magma chambers, can explain the relative homogeneity of basaltic lavas erupted to form this (and other) oceanic plateaux. The CaribbeanColombian oceanic plateau may have formed at
Corresponding author. Telephone: +44 116 2523639. Personal fax: +44 116 2523639. Department fax: +44 116 2523918. e-mail: ack2@le.ac.uk Present address: Graduate School of Oceanography, Rhode Island University, Narragansett, RI 02882, USA

an oceanic spreading centre, and valuable comparisons can be made between Iceland and the CaribbeanColombian plateau.

KEY WORDS:

basalt; Colombia; geochemistry; mantle plume; oceanic

plateau

INTRODUCTION
The relative contribution of lithospheric vs asthenospheric (plume) mantle sources to continental ood basalt volcanism is a contentious subject in the petrological literature (e.g. Gallagher & Hawkesworth, 1992; Saunders et al., 1992; Arndt et al., 1993; Thirlwall et al., 1994; Gibson et al., 1995). One thing seems relatively clear: only a small percentage of plume-derived continental ood basalts pass cleanly through the continental lithosphere, with the rest being contaminated. The nature of this contamination and the discussions bearing on its relative importance have tended to divert attention away from the primary cause of the volcanism, the plume itself. Increasingly, however, it has been realized that mantle plumes form not only continental ood basalts but also

Oxford University Press 1997

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

oceanic ood basalts or, more specically, oceanic plateaux that have thickened oceanic crust (>8 km) (e.g. Carlson et al., 1980; Ben-Avraham et al., 1981; Mahoney, 1987). Such plateaux are only locally exposed at the surface (e.g. Ontong Java in the Pacic and Kerguelen in the Indian Ocean), but they oer us the chance to study what is, in eect, an oceanic analogue of a continental ood basalt province, without the complication of sub-continental lithospheric contamination. A major problem is that most parts of extant oceanic plateaux are still submerged beneath deep water. To a certain extent, they can be sampled by drilling (e.g. Vallier et al., 1980; Saunders, 1985; Mahoney, 1987; Mahoney & Spencer, 1991; Storey et al., 1992) but this really just scratches the surface (Con & Eldholm, 1993), as the greatest depth to which any of these oceanic large igneous provinces has been drilled is 700 m into the Nauru Basin (Saunders, 1985). Nevertheless, another way in which oceanic plateaux have been made accessible for detailed study stems from the fact that these plateaux (especially if they are young) are more buoyant than normal oceanic crust and will therefore resist subduction (Ben-Avraham et al., 1981; Cloos, 1993; Saunders et al., 1996). Thus the upper reaches of such a plateau may be obducted onto the margin of the overriding plate (see Kimura & Ludden, 1995). Such a situation is believed to have occurred in the Solomon Islands (Tejada et al., 1996) and the Wrangellia terrane (Lassiter et al., 1995). Arguably, the best exposed oceanic plateau available for study is the Late Cretaceous Caribbean oceanic plateau, which makes up most of the Caribbean Plate, and is exposed around its tectonically disturbed margins (Donnelly et al., 1990; Kerr et al., 1996b, c, 1997) (Fig. 1). In this paper we focus on the less well-known parts of the Caribbean plateau which were accreted against northwestern South America, and now form a series of mac terranes in western Colombia. We present new geochemical and SrNd isotopic data to demonstrate that these Cretaceous igneous rocks are part of the Caribbean oceanic plateau, rather than supra-subduction zone ophiolites. We assess the nature of the mantle plume involved in the petrogenesis of these Colombian volcanics, and present new 40Ar/39Ar radiometric ages for the basalts, so as to estimate the spatial and temporal extent of this plume-related volcanism.

REGIONAL TECTONIC SETTING


The distribution of volcanic rocks belonging to the CaribbeanColombian Cretaceous Igneous Province is shown in Fig. 1. These Cretaceous volcanics around the margin of the Caribbean have been discussed by various workers [see Donnelly et al. (1990) and Kerr et al. (1996c, 1997) for reviews], and indeed the basalts forming the

Caribbean sea oor itself have been drilled by DSDP Leg 15 (Donnelly et al., 1973; Bence et al., 1975). The Caribbean ocean oor is composed of anomalously thick crust (up to 20 km; Edgar et al., 1971), which Donnelly (1973) and Donnelly et al. (1973) considered a ood basalt province. It is now generally accepted that the Caribbean plate was formed in the eastern Pacic as a Large Igneous Province (LIP) in the Late Cretaceous (e.g. Duncan & Hargraves, 1984; Burke, 1988; Pindell & Barrett, 1990; Kerr et al., 1997). Using a xed hotspot reference frame, Duncan & Hargraves (1984) and Hill (1993) suggested that the magmas of the CaribbeanColombian Cretaceous Igneous Province were produced by partial melting within the initial plume head of the Galapagos hotspot. East ward movement of the Farallon plate in the Late CretaceousEarly Tertiary forced the northern half of the plateau into the ocean basin which had been opening between North and South America since the Jurassic. The eastward moving plateau appears to have been too buoyant (owing to its thermal structure and crustal thickness) to be subducted beneath the westward moving American oceanic plate (Burke et al., 1978; Hill, 1993), thus jamming the subduction zone and causing a ip in the direction of subduction from east to west, such that the Atlantic plate was being consumed at the subduction zone, as opposed to the FarallonCaribbean plate (Pindell & Barrett, 1990; Lebron & Pert, 1994). Further south, the CaribbeanColombian plateau impinged against the northwestern continental margin of South America, leading to imbrication and obduction of the plateau and progressive westward back-stepping of the subduction zone to form the accreted oceanic plateau terranes of northwestern South America (Millward et al., 1984; Kerr et al., 1996c). The Romeral Fault zone (Fig. 2) represents a major terrane boundary in northwestern South America. To the east of the Romeral Fault, Bouguer gravity anomalies are strongly negative (220 mgal; Case et al., 1973), conrming geological observations that the basement is composed of continental crust. In contrast, to the west of the Romeral Fault system, anomalies are strongly positive (+135 mgal at the Pacic coast and +75 mgal in the Western Cordillera of Colombia; Case et al., 1971). Therefore, the basement to the west of the Romeral Fault is composed of high-density material, consistent with an oceanic origin. Because present-day volcanoes in western Colombia erupt subduction-related lavas, several earlier interpretations of the tectonic setting of the Cretaceous basalts suggested a subduction-related environment of formation (Barrero, 1979; McCourt et al., 1984; Bourgois et al., 1987; Spadea et al., 1987; Grosser, 1989). Other workers have argued that the basic lavas of western Colombia represent former oceanic crust (Pichler et al.,

678

KERR et al.

CRETACEOUS COLOMBIAN BASALTS

Fig. 1. Map showing the location of Cretaceous basalts of the CaribbeanColombian oceanic plateau. Also shown are Deep Sea Drilling Project (DSDP) Leg 15 drill sites.

1974; Mooney, 1980; Bourgois et al., 1982). However, as Nivia (1987) has pointed out, typical ophiolitic sequences, most of which have sheeted dyke complexes, indicative of spreading ridges, are not found in the Cretaceous accreted terranes of Colombia. Here we evaluate more fully the suggestion by Millward et al. (1984), Nivia (1987), Storey et al. (1991) and Kerr et al. (1996c) that the Cretaceous Colombian mac terranes formed as part of an oceanic plateau, resulting from partial melting of a mantle plume, which generated thicker than normal oceanic crust. Previous studies (e.g. Millward et al., 1984; Spadea et al., 1987; Storey et al., 1991) have tended to focus on the major and trace element composition of basalts from relatively small areas. This study, however, is the rst to report comprehensive major and trace element data, SrNd isotopic analyses, and precise 40Ar/39Ar ages, for the Cretaceous basalts throughout a wide area of Colombia. In doing this we hope to shed some new light on the petrogenesis and original tectonic setting of the Cretaceous Colombian basalts.

CRETACEOUS COLOMBIAN BASALTIC TERRANESFIELD RELATIONS AND SAMPLING


The Cretaceous mac sequences form three main belts (Fig. 2) which trend approximately NNESSW, namely the Central Cordillera, the Western Cordillera and the Serrana de Baudo along the Pacic coast.

Central Cordillera
The mac igneous rocks of the western ank of the Central Cordillera occur in several discontinuous lenses, from Medelln in the north to Pasto in the south. The exposures are bounded to the east by the Romeral Fault (Fig. 2). Cretaceous volcanics of the Central Cordillera also extend along the trace of the Romeral Fault into northern Ecuador (Lebras et al., 1987). The largest continuous outcrop of mac volcanics in the Central Cordillera, known as the Amaime Formation

679

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

Fig. 2. Map of western Colombia showing the three main belts of Cretaceous igneous rocks, along with sample localities mentioned in the text.

(Fig. 3b) (McCourt et al., 1984), is 140 km long, 520 km wide and bounded by the Romeral Fault to the east and by the GuabasPradera Fault to the west. The formation consists of both massive and pillowed tholeiitic basalts and occasional cumulate picrites which occur in several fault-bounded blocks of 510 km width. The occurrence of picrites within the Amaime Formation has also been noted by McCourt et al. (1984) and Spadea et al. (1989). To the west of the GuabasPradera Fault an elongate (northsouth) body of mac and ultramac rocks (30 km long 7 km wide) is exposed, known as the Ginebra

Ultramac Complex (Nivia, 1987) (Fig. 3b). This body consists of a sequence of dunites, wehrlites and layered and isotropic gabbros, which are overlain by amphibolitized basalts. Aspden & McCourt (1986) proposed that this sequence represents a lower-crustal level of the Amaime Formation. Other exposures of the igneous rocks of the Central Cordillera shown in Fig. 2 include the pillowed picritic and picritic-basalts and tu-breccias found at El Encenillo, 20 km SSW of Popayan, and the Los Azules complex, which consists mostly of a series of ultramac cumulates with some massive and pillowed picritic to basaltic lavas (Spadea et al., 1989). Some

680

KERR et al.

CRETACEOUS COLOMBIAN BASALTS

Fig. 3. Close-up maps of (a) the Serrana de Baudo and (b) the area around Cali [after Nivia (1987)] detailing sample locations. The areas covered by these maps are outlined in Fig. 2.

40 km north of Pasto a sequence of pillowed and massive lavas is exposed along a new road cut traversed by the Pan American Highway, near the village of Taminango. Along the eastern margin of the Romeral Fault a 510 km wide belt of high-pressure lawsoniteglaucophane schists and eclogites can be found stretching from southern Ecuador to northcentral Colombia. These highgrade metamorphic rocks are associated with highly tectonized and serpentinized ultramac rocks and gabbros (McCourt & Feininger, 1984), and are known as the Arqua Complex (Fig. 2). Exposed to the east of the Arqua Complex is a discontinuous (510 km wide) belt of basalts, andesites and tus, tectonically intermixed with Palaeozoic low-grade schiststhe Quebradagrande Complex (Fig. 2). Preliminary geochemical studies (A. Nivia et al., unpublished data, 1996) of lavas and tus from the Quebradagrande Complex suggests that they have been formed in a subduction-related tectonic setting. During the present study, basalts and several cumulate picrites were sampled from the main outcrop of the Amaime Formation east of Cali [SW and NE of the

town of Sevilla and due west of Buga (AMA112), as well as west of Florida (FLO14); Fig. 3]. Basaltic samples were collected along the Pan American Highway near Taminango (YAN18), and at El Encenillo (Fig. 2) blocks in a picrite breccia (ROM4ii) and a picritic ow (ROM2) were sampled. The volcanic section of the Los Azules complex (Fig. 2) yielded basalts and cumulate and noncumulate picrites. Only the non-cumulate picrites and basalts (LER4, AZU2 and -3 and ROM510) will be discussed in this paper. A basaltic sample (BAR2) was also collected from a minor fault-bounded sliver 30 km SW of Medelln, which is believed to be part of the basaltic sequence of the Central Cordillera.

Western Cordillera
The most extensive outcrops of Cretaceous basalt in Colombia occur in the Western Cordillera (Figs 2 and 3), which is separated from the Central Cordillera by the CaucaPatia Graben, and from the coastal Serrana de

681

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

Table 1: Stratigraphic nomenclature of the rocks of the Western Cordillera


Northern Western Cordillera Barroso Formation (volcanic) SouthCentral Western Cordillera

Penderisco Formation (sedimentary)

Volcanic Formation (volcanic) Canasgordas Group Espinal Formation (sedimentary) Cisneros Formation (sedimentary)

Baudo sequence by the San JuanAtrato Trough (Fig. 2). The igneous rocks crop out in NNESSW trending fault-bounded slices which can be up to 15 km wide, although they are more often <10 km in width. Steeply eastward-dipping fault-bounded lenses of metasediments sometimes separate the volcanic slices (Fig. 3). Sedimentary and structural data indicate that the sequence in each fault block youngs towards the east, with vertical to sub-vertical bedding planes (Millward et al., 1984). The rocks of the Western Cordillera are known by several dierent formation names depending on their location in Colombia, and these are summarized in Table 1. In central and southern Colombia sedimentary rocks comprise ~30% of the Western Cordillera and they have been divided into two main formations: the Cisneros and Espinal Formations (Barrero, 1979; Aspden, 1984). The Cisneros Formation may be up to 2000 m thick (Barrero, 1979; Millward et al., 1984); it consists of a sequence of strongly deformed slates and phyllites with thin lenses of meta-limestones, cherts and greywackes, and is in faulted contact with igneous rocks of the Western Cordillera. In contrast, the ~700 m thick Espinal Formation is composed of a series of relatively unmetamorphosed cherts, shales and pebbly sandstones, and appears to be conformable with the volcanics of the Western Cordillera (McCourt et al., 1984; Millward et al., 1984). It has recently been suggested by Nivia et al. (1996) that the dierence in deformation between the two formations is due to the relative competence of each formation, with the generally ner-grained Cisneros sediments being more deformed than the coarser Espinal Formation sediments. Palaeontological evidence (Barrero, 1979), suggests that the Cisneros Formation is older than the Espinal (and Volcanic) Formation. The Volcanic Formation (formerly the Diabase Group) consists of a >5 km thick (Barrero, 1979) sequence of both pillowed and massive basaltic lavas, dolerites, local gabbros and rare tus. The pillow lavas can be up to 50 m thick, and hyaloclastite breccia is commonly found between the individual pillows, which sometimes preserve chilled skins and occasionally concentrically arranged amygdales. Coarser-grained basalts and dolerites form either massive ows or intrusive sheets (200300 m thick). The basalts and dolerites are highly sheared in places

and have undergone low-grade (zeolite and prehnite pumpellyite facies) metamorphism (Barrero, 1979). Associated sedimentary lenses (chert and grey shales) are usually <30 m thick. Very few mac dykes cut the volcanic formation, although there are several small gabbroic plugs. McBirney (1963) has calculated that at hydrostatic pressures >315 bars (315 km seawater depth) explosive fragmentation and vesicle formation are not possible. Thus the abundance of pillow lavas with only rare vesicles (now inlled) and the scarcity of intercalated limestones and tus, combined with the occurrence of cherts and grey shales within the basaltic succession, imply that the bulk of the lava succession was erupted in a considerable depth of water (>3 km). The presence of a few tus within the lava succession of the Western Cordillera suggests that as the lava pile accumulated, eruptions may have occurred in shallower water; nevertheless, no evidence of subaerial eruptions has so far been found In the eastern part of the Western Cordillera, at 415N (Fig. 3b) a suite of ultramacmac rocks crops out (Barrero, 1979; Nivia, 1996), known as the BolvarRo Frio ultramac complex (Kerr et al., 1997). This complex consists of both layered and isotropic gabbros and norites, structurally underlain by serpentinized dunite containing bands of both clinopyroxenite and olivine gabbronorite. The composition of the rocks from the Bolvar complex strongly suggests that they are genetically related to the basalts of the Western Cordillera (Nivia, 1996; Kerr et al., in preparation) and that they could represent the lower section of the CaribbeanColombian oceanic plateau (Kerr et al., 1997). In the northern section of the Western Cordillera, sediments make up more than half of the outcrop. These sediments, the Penderisco Formation, consist of a sequence of cherts, black micritic limestones and deposits with possible turbiditic characteristics (Alvarez & Gonzalez, 1978). The associated volcanic rocks, the Barroso Formation, include basalts (with some pillows), dolerites, hyaloclastites, thin tus, and volcanic breccias, with occasional lenses of pelites and cherts. Along the western periphery of the Western Cordillera, a sequence of Early Tertiary subduction-related volcanics crop out, the Dabeiba Volcanic Arc (Tistl & Salazar,

682

KERR et al.

CRETACEOUS COLOMBIAN BASALTS

Table 2: 40Ar39Ar plateau and isochron age calculations from Western Colombia
Sample Location Material Plateau age by 1/ SDB11 SDB11 SDB16 AN1464 PAN6 Serrana de Baudo Serrana de Baudo Serrana de Baudo Dabeiba Volcanic Arc Western Cordillera whole rock plagioclase plagioclase plagioclase whole rock
2 39

Ar % of

Isochron age (Ma) 72908 71814 76211 41486 94764

SUMS/ (N 2)

40

Ar/36Ar

(m.y.)

total 98 95 35 72 89

intercept1 304429 324243 3106118 3232137 2924156

73608 72504 77910 43104 91727

5 10 3 5 3

260 055 058 1703 025

Ages calculated using the following decay constants: e=05811010 yr1, b=49631010 yr1. Isotopic interferences for the OSU TRIGA reactor are: ( 36Ar/37Ar)Ca=264104, ( 39Ar/37Ar)Ca=673104 and ( 40Ar/39Ar)K=100103. Reported errors are 2 . Sample PAN6 is from Sinton (1996).

1994). The basalts, andesites and pyroclastic rocks of this arc are best developed to the west of Medelln (Fig. 2), and the rocks appear to be signicantly younger than the basalts of the Western Cordillera. 40Ar/39Ar dating of a basalt from the Dabeiba Volcanic Arc (AN1464), using step heating techniques, yielded a plateau age of 43104 Ma (i.e. Eocene) (Table 2). To the south, in Ecuador, the Macuchi Formation has been widely correlated with the Volcanic Formation (McCourt et al., 1984; Lebras et al., 1987; WallrabeAdams, 1990; Van Thournout et al., 1992). This formation consists of a more westerly subduction-related sequence of lavas and volcaniclastic sediments (Henderson, 1979; Lebras et al., 1987). A scattered discontinuous belt of basalts intercalated with silicied shale, along with peridotites and layered gabbros, crops out along the eastern margin of the belt, and has been termed the midocean ridge basalt of the Macuchi Formation (Lebras et al., 1987). Western Cordillera igneous rocks were sampled from new road cuts along the Cali to Buenaventura road (CBU217; Fig. 3b). Additionally, volcanic rocks were sampled in the Vijes area, 30 km NNE of Cali (VIJ 14, PAN211 and 19) and at Calima, 50 km north of Cali (COL-1; Fig. 3b). A suite of lavas have also been collected from the Barroso Formation type-section 4060 km SW of Medelln (Fig. 2), on the Remolino to El Barroso road, and from the Barroso valley along the road to Jardn (BAR311).

not reported by previous workers (Goossens et al., 1977; Macia, 1985), intercalations of ne-grained sandstone, chert and limestones were found between some of the basalts. The nature of these intercalated sediments implies a shallower eruption depth for these basalts than for those of the Western Cordillera, and this is perhaps a reection of their younger age (see below). The basalts are unconformably overlain by a poorly exposed sequence of subduction-related basalts which are intercalated with Eocene limestones (Gansser, 1973). McGeary & Ben-Avraham (1986) proposed that the volcanic succession on Gorgona Island (Fig. 2) represents a continuation of the basalts of the Serrana de Baudo. Gorgona Island is the location of the worlds only known Phanerozoic komatiites (Echeverra, 1980; Kerr et al., 1996a). However, despite extensive sampling along the coast from 710N to 530N (Fig. 3a), no evidence of lavas with spinifex (komatiitic) textures was found. Basalts and some gabbros were collected from the Serrana de Baudo coast (SDB125; Fig. 3a). Dense jungle and deep tropical weathering meant that sampling was restricted to coastal exposures.

ANALYTICAL METHODS
After powdering in an agate Tema mill, major and trace elements were analysed by X-ray uorescence (XRF) at Royal Holloway University of London and Leicester University using conventional techniques [see Tarney & Marsh (1991) and Kerr et al. (1996b) for further details]. The 2 SD errors are as follows: SiO2, 02; TiO2, 003; Al2O3, 01; Fe2O3, 01; CaO, 005; MgO, 01; Na2O, 01; K2O, 003; MnO, 003; P2O5, 003; Ba, 20; Cr, 1%; Ga, 10; Nb, 02; Ni, 1%; Rb, 03; Sr, 1%; V, 15%; Y, 04; Zr, 05. The XRF data are reported in Table 3. The rare earth elements (REE) and Th, U, Co, Sc, Ta and Hf (Table 4) have been analysed by instrumental neutron activation analysis (INAA) at the University of

Serrana de Baudo
The westernmost belt of Cretaceous volcanic rocks outcrops principally along the Pacic coast of NW Colombia, known as the Serrana de Baudo (Figs 2 and 3a). The coastal exposures in the region consist of pillowed and massive basalts, with some basaltic breccias (with clasts up to 30 cm in diameter), dolerites and gabbros. Although

683

Table 3: X-ray uorescence major and trace element data for Cretaceous Colombian volcanic and intrusive rocks
CaO Na2O K2O TiO2 MnO P2O5 Total LOI % Ni Cr V Ga Sr Rb Ba Zr Nb Y

Type:

Sample SiO2

Al2O3

Fe2O3 MgO

JOURNAL OF PETROLOGY

Serrana de Baudo

VOLUME 38

684
863 822 892 832 1029 1039 868 925 829 776 871 893 1247 1234 1244 1346 1011 1001 1117 1192 1284 1172 1313 1177 172 204 160 104 256 220 257 209 231 278 211 269 007 027 007 004 037 076 050 077 017 008 029 033 085 094 091 095 124 115 118 110 125 128 116 113 0202 0162 0158 0091 0233 0225 0200 0205 0153 0182 0190 0186 0076 0082 0078 0085 0087 0091 0081 0076 0088 0097 0092 0087 10037 10068 10049 10097 9990 9984 9995 9999 9999 9990 9995 9996 146 169 169 178 297 301 271 288 309 302 271 280 133 143 2 18 113 118 128 147 103 94 129 140 441 573 0 70 261 301 336 522 288 240 420 365 252 254 149 257 331 336 328 305 333 331 311 299

1459 1470 1403 1485 1457 1404 1405 1351 1443 1386 1367 1570 1691 1312 1425 1419 1410 1384 1462 1448 1392 1615 1605

1214 1222 1339 1131 1156 1267 1263 965 1092 1262 924 1079 980 1186 1153 1334 1299 1462 1114 1107 1321 1305 1235

821 780 759 834 719 734 746 994 835 753 1010 710 1013 710 690 677 681 618 797 805 508 759 773

1204 1210 1096 1203 1138 1210 798 1453 1183 1155 1375 1181 1212 1257 900 1094 1153 1123 1010 1154 1127 841 857

318 328 402 310 409 301 703 205 275 334 148 243 142 235 475 289 260 247 233 280 298 319 479

022 021 014 019 015 010 005 009 059 016 009 023 009 015 013 015 022 016 111 018 104 075 091

117 122 140 100 126 137 115 057 098 118 063 110 076 122 120 147 129 166 092 094 291 257 249

0191 0212 0143 0143 0189 0223 0169 0171 0184 0210 0177 0207 0141 0217 0163 0230 0166 0192 0170 0223 0215 0173 0182

0096 0097 0117 0075 0102 0106 0089 0039 0075 0088 0051 0090 0061 0088 0096 0117 0108 0092 0070 0074 0339 0258 0257

9977 10027 9995 9971 9985 9946 9992 9883 9916 9861 9903 9940 10010 9933 9973 10072 10065 10069 9984 10062 10034 10021 10025

223 181 163 110 132 147 399 139 092 096 210 230 207 167 342 157 289 059 179 210 090 291 351

115 112 87 127 110 95 97 178 97 96 226 105 282 80 104 97 79 41 111 116 77 88 81

318 317 169 408 335 242 236 489 250 255 917 252 418 100 228 208 157 18 353 342 145 136 139

276 292 305 311 286 309 309 220 321 315 202 228 170 295 261 331 273 527 278 283 387 315 305

154 157 172 154 163 168 127 132 173 175 138 165 151 172 130 191 174 216 150 150 201 202 194

220 247 216 184 124 228 39 88 106 95 91 191 90 129 95 156 186 112 306 213 331 345 363

22 19 19 34 28 12 24 22 57 20 24 33 26 26 30 25 33 29 69 26 251 70 68

191 109 71 31 44 33 15 28 34 45 27 63 13 26 64 37 283 26 196 36 131 449 1042

69 74 81 57 67 82 50 28 48 61 33 70 34 63 59 70 73 55 67 58 217 170 176

33 35 51 36 40 47 35 12 35 35 13 39 25 33 36 43 42 27 26 35 199 133 148

223 231 284 201 242 257 229 133 194 249 144 233 142 246 238 294 249 256 186 203 365 297 318

NUMBER 6 JUNE 1997

bas SDB1 4793 bas SDB2 4844 bas SDB4 4816 bas SDB5 4868 bas SDB7 4937 bas SDB8 4850 bas SDB9 4932 bas SDB10 4827 bas SDB11 4906 bas SDB12 4808 bas SDB13 4985 dol SDB14 4993 gab SDB15 4867 gab SDB16 5066 bas SDB17 5171 bas SDB18 5063 bas SDB19 5085 br cl SDB20 5025 br cl SDB21 5141 bas SDB22 5126 br cl SDB23 4938 bas SDB24 4806 br cl SDB25 4693 Western Cordillera bas BAR5 5161 bas BAR6 5202 bas BAR7 5159 bas BAR8 5193 bas CBU2 4895 bas CBU3 4887 bas CBU4 4907 bas CBU5 4839 bas CBU6 4857 bas CBU7 4900 bas CBU8 4790 bas CBU9 4877 141 138 152 132 173 160 164 163 165 155 157 156 99 93 417 390 137 159 207 218 176 187 358 154 69 23 470 226 25 55 58 68 16 07 36 30 49 40 918 654 138 229 289 345 57 67 258 130

1431 1444 1441 1456 1446 1535 1509 1488 1477 1513 1512 1514

1043 1016 1032 1050 1161 1080 1141 1132 1154 1188 1125 1093

51 46 76 59 58 61 57 50 62 66 58 60

45 46 71 24 34 33 32 26 33 35 34 32

181 175 240 142 213 241 216 197 232 241 232 223

1211 1479 1478 1732 1191 1050 1593 1599 1117 1180 1181 1243 1192 1391 1118 1066 1520 1511 1430

866 601 592 152 797 853 503 511 870 817 779 850 744 658 490 929 578 620 650

1221 698 777 666 1127 1068 754 784 1169 1166 1089 1138 1090 986 571 1222 957 915 996

229 507 462 344 219 251 290 269 219 216 250 224 257 312 477 167 229 204 220

004 014 034 012 022 017 004 006 008 007 020 024 013 022 002 031 007 007 007

124 197 186 145 114 113 194 188 099 106 113 123 121 162 079 111 202 202 198

0180 0205 0249 0265 0182 0141 0203 0213 0178 0198 0182 0234 0192 0231 0134 0164 0185 0230 0179

0101 0162 0157 0159 0089 0087 0192 0163 0084 0092 0101 0095 0091 0128 0120 0093 0171 0177 0158

9992 9950 9960 9955 10000 9985 10016 10022 9995 9979 10009 9969 9972 10013 9845 9947 9959 9997 9989

300 217 169 031 235 327 278 270 098 064 109 224 112 118 176 222 090 118 159

117 67 56 30 140 97 35 45 130 109 89 96 91 65 52 145 58 57 72

355 69 62 3 193 262 13 17 235 199 170 183 190 54 73 360 63 61 85

311 475 464 8 316 290 454 490 305 326 342 401 370 430 210 246 531 531 510

159 190 152 202 156 157 176 171 157 167 163 166 155 188 149 161 202 188 173

117 63 33 116 99 79 74 94 100 91 101 110 100 99 40 131 89 89 100

11 11 27 24 24 19 04 01 12 07 29 26 15 29 00 15 04 07 05

78 92 153 35 21 701 35 34 20 19 34 86 35 58 13 69 27 27 26

66 110 108 240 56 56 128 102 49 55 58 62 56 80 67 64 119 120 107

39 72 76 255 38 30 86 70 36 41 42 39 38 55 41 26 80 82 72

244 388 383 762 229 211 464 379 204 232 235 253 247 320 247 242 417 417 393

KERR et al.

CRETACEOUS COLOMBIAN BASALTS

685
1037 907 810 1324 937 1664 713 961 1139 864 1041 1579 1065 1150 1254 947 175 382 376 126 310 190 334 389 009 114 046 079 044 022 003 027 221 269 137 102 148 091 166 127 0257 0120 0183 0161 0176 0230 0191 0176 0246 0355 0142 0083 0196 0092 0172 0103 9909 9964 10003 9988 9976 9982 9965 9953 201 311 190 161 182 219 327 189 437 109 123 223 210 436 96 186 729 334 394 747 411 1556 150 844 286 259 369 378 365 325 423 361

bas CBU10 4819 1489 bas CBU11 5106 1312 bas CBU12 5141 1250 dio CBU13 5794 1068 dol CBU14 5037 1467 dyke CBU17 5222 1388 bas PAN2 5373 1265 bas-a PAN3 5390 1240 dol PAN5 5007 1479 dol PAN6 5025 1433 dol PAN7 5130 1419 dol PAN8 4829 1506 dol PAN9 5089 1438 dol PAN11 5113 1334 and PAN19 5807 1275 dol COL1 4909 1487 bas VIJ1 5145 1285 bas VIJ2 5185 1312 bas VIJ4 5120 1334 Central Cordillera bas AMA1 5414 1432 bas AMA2 5045 1469 bas AMA3 5022 1379 bas AMA4 4991 1436 bas AMA5 5011 1421 bas AMA6 5010 1437 bas AMA8 4998 1398 bas AMA9 5024 1436 bas AMA11 5004 1361 dol AMA12 5035 1488 bas FLO1 5043 1421 bas FLO2 4692 1523 bas FLO3 4981 1378 bas FLO4 4987 1382 dol YAN1 5040 1416 bas YAN2 4938 1316 bas YAN3 5041 1467 bas YAN4 5086 1422 dyke YAN5 5125 1331 bas YAN6 4976 1266 bas YAN7 4869 1500 bas YAN8 5128 1355 bas BAR2 5140 1452 Los Azules and El Encenillo bas ROM2 4678 1234 bas ROM4ii 4906 1366 bas LER4 4973 1325 p-bas AZU3 4899 866 bas ROM5 4853 1462 pic ROM7 4614 974 bas ROM9 4727 1438 bas ROM10 4959 1243 766 915 1030 906 799 902 771 878 728 715 843 747 898 927 652 523 940 857 650 499 920 693 748 167 191 163 117 197 135 167 154 179 176 219 263 295 40 83 328 970 1058 1223 1275 1201 1205 1152 1285 1191 1276 1211 2004 1404 1379 894 889 1184 1322 996 1304 1202 994 890 265 297 145 162 196 215 203 181 165 124 271 012 177 169 284 247 219 139 234 044 184 307 352 006 077 008 030 011 032 031 004 015 008 020 000 010 006 053 018 017 003 011 002 015 023 031 091 087 088 092 109 099 201 099 221 038 099 082 094 094 147 202 081 087 144 178 097 123 131 0114 0171 0175 0171 0205 0177 0182 0180 0183 0251 0173 0155 0167 0170 0180 0234 0156 0179 0202 0232 0215 0205 0219 0098 0068 0078 0081 0083 0085 0181 0087 0204 0032 0086 0070 0077 0080 0112 0171 0061 0066 0120 0148 0079 0099 0102 9987 9984 9977 9988 9980 9993 9974 9993 9971 9983 9983 9980 9994 9992 9983 9979 9992 9994 9989 9976 9974 9988 10049 071 226 173 140 007 233 163 214 077 033 144 364 084 120 393 141 116 359 057 353 136 150 271 96 124 229 143 148 144 136 110 119 52 135 137 159 158 71 53 157 135 82 57 147 82 105 204 489 602 415 403 438 310 213 212 96 433 422 454 456 27 30 562 404 81 40 426 103 206 333 298 282 291 293 295 352 363 359 295 304 257 285 286 416 558 294 322 420 510 336 407 312 144 128 131 133 149 145 173 140 199 125 143 199 145 148 167 193 130 136 177 157 153 155 139 90 132 149 139 97 110 238 812 235 30 125 12 86 103 120 99 96 85 90 22 98 421 174 05 103 14 50 00 46 62 15 24 07 40 03 16 07 96 23 31 03 13 03 36 31 55 29 161 49 99 64 32 03 32 23 48 27 62 17 40 82 40 51 15 23 3 23 15 26 29 21 7 25 50 45 42 52 135 149 939 133 52 15 80 60 41 47 48 56 54 125 47 139 16 54 44 51 51 69 113 41 45 74 98 50 61 74 148 198 69 41 92 46 93 61 41 27 45 44 51 51 152 34 172 07 48 41 46 47 41 71 26 27 48 62 32 43 35 248 303 51 30 88 31 60 46 208 177 164 181 195 175 246 224 268 99 190 154 175 172 280 432 175 190 304 385 207 277 252 239 265 205 159 236 137 236 205

1022 1012 1057 1072 1203 1067 1184 1060 1247 1271 1049 897 1028 1023 1469 1805 1022 1054 1466 1669 1158 1335 1272

1366 1109 1263 989 1121 1244 1294 1272

bas, basalt; dol, dolerite; gab, gabbro; br cl, breccia clast; dio, diorite; and, andesite; bas-a, basaltic andesite; p bas, picritic basalt; pic, picrite. All iron reported as Fe2O3. All totals reported on a volatile-free basis. ROM2 and ROM4 are from El Encenillo; LER4, AZU3 and ROM5, -7, -9 and -10 are from Los Azules.

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

Leicester [see Fitton et al. (1997) for analytical details]. Additionally, several samples were analysed for REE by inductively coupled plasma-atomic emission spectroscopy at Royal Holloway [see Walsh et al. (1981) for analytical details]. Sr and Nd isotope ratios (Table 5) were measured at Royal Holloway, on a VG354 mass spectrometer tted with ve Faraday collectors. The isotopic analysis procedures have been described by Thirlwall (1991a, b). The powders used for Sr isotope analysis were leached in 6M HCl for 1 h before commencing the chemical procedures. The more altered samples were leached a second time in 6M HCl. During the period of the analyses, international reference material SRM987 averaged 071024317 (2 SD, n=64) and laboratory standard Low Aldrich averaged 05114219 (2 SD, n=14). [See Thirlwall (1991a) for comparison of Low Aldrich with international standards.] Age determinations for whole-rock and plagioclase mineral separates from the least altered samples were performed at Oregon State University using standard 40 Ar39Ar incremental heating techniques (Duncan & Hargraves, 1990; Duncan & Hogan, 1994) Whole-rock samples were either crushed in bulk in a ceramic jaw crusher (PAN6) and sieved to a uniform 0510 mm grain size or made into mini-cores. Plagioclase crystals were magnetically separated, checked for purity under a binocular microscope, briey washed in 5% HF and ultrasonically cleaned in distilled water. Samples were sealed in evacuated quartz glass vials and irradiated for 6 h at the OSU TRIGA nuclear reactor facility. Neutron ux during irradiation was monitored by FCT-3 biotite (277 Ma, Hurfurd & Hammerschmidt, 1985). Ar isotopes of the whole-rock samples (~500 mg) were determined using an AEI MS-10S mass spectrometer. The plagioclase separates (~100 mg) were analysed using an MAP 21550 mass spectrometer. Individual ages for each 40Ar39Ar temperature step were calculated after corrections for background, mass fractionation, isotopic interferences and atmospheric argon contamination ( 40Ar/36Ar=2955). Plateau ages were calculated from consecutive steps that are concordant within 2 error using the procedure described by Dalrymple et al. (1988), in which step ages were weighted by the inverse of their variance (Table 2). Isotope correlation diagrams ( 36Ar/40Ar vs 39Ar/40Ar) were made for each analysis, in which the slope is proportional to the age (isochron) and the inverse of the y-intercept gives the initial 40Ar/36Ar composition. The isochron ages are calculated using the same steps as in the plateau ages. Plateau ages almost invariably have smaller errors because the weighting procedure emphasizes the most precisely determined age steps, whereas uncertainties in the isochron ages reect the analytical errors of the more

poorly determined step ages and the particular distribution of Ar isotopic compositions in 36Ar/40Ar vs 39Ar/ 40 Ar plots. We will therefore restrict our discussion to the plateau ages. Samples CBU11 and CBU12 from the Western Cordillera displayed erratic step ages such that no age plateaux or isochrons could be discerned. This can be attributed to a combination of alteration, the low K2O contents of the rocks and interferences from hydrocarbons.

GEOCHRONOLOGY
Dierent environments of formation as well as dierent relative ages have been proposed for the three belts of Cretaceous igneous rocks: (1) The igneous rocks of the three belts form part of one near-synchronous province (Goossens & Rose, 1973; Goossens et al., 1977; Marriner & Millward, 1984). (2) The volcanic and intrusive rocks of the Western and Central Cordillera form one province, whereas those of the Serrana de Baudo constitute a younger province (Barrero, 1979; Feininger & Bristow, 1980). (3) The Cretaceous volcanic belts young westwards from the Central Cordillera, through the Western Cor dillera, to the Serrana de Baudo (McCourt et al., 1984; Aspden et al., 1987; Lebras et al., 1987). These previous interpretations are unfortunately heavily dependent on unreliable (owing to Ar and K loss, or K gain) K/Ar ages, and on palaeontological evidence from associated sediments that are only in tectonic contact with the volcanic rocks. We have dated two basalts from the Serrana de Baudo, one from the Western Cordillera, and one from the Dabeiba Volcanic Arc by 40Ar/39Ar step heating. We have also restricted palaeontological evidence to those fossils that are derived from sediments which are clearly stratigraphically intercalated (rather than tectonically intercalated) with the volcanic rocks. The Serrana de Baudo rocks yield 40Ar/39Ar stepheating plateaux which range from 72504 to 77910 Ma (Table 2), consistent with the occurrence of Upper Cretaceous bivalves in intercalated sediments (Gansser, 1973). 40Ar/39Ar dates for the basalts and gabbros of Gorgona Island, the proposed oshore continuation of the Serrana de Baudo (Kerr et al., 1996a), have yielded signicantly older ages (86046 to 88319 Ma) than the basalts of the Serrana de Baudo. The implications of this will be discussed below. The basalts of the Western Cordillera have proved more problematical to date, owing to both their altered nature and low K2O contents. Nevertheless, one sample (PAN6) has been successfully dated (Table 2) and has yielded a plateau age of 91727 Ma. This age is also consistent with palaeontological evidence from the

686

Table 4: Additional trace element data for selected Cretaceous Colombian intrusive and volcanic rocks
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Yb Lu Hf Ta Th U

Sample Method Sc

Co

Serrana de Baudo SDB8 INAA 452


34 15 27 15 34 38 32 126 23 28 27 59 49 96 31 27 39 25 61 58 32 119 05 37 35 85 78 24 09 101 79 27 10 44 17 15 bdl 03 02 08 06 280 184 45 16 54 08 86 66 21 07 24 05 164 1210 377 142 551 095 166 124 40 15 09 61 65 21 08 06 111 15 89 27 11 46 10 73 09 58 17 07 29 06 19 31 85 73 23 09 30 07 274 223 72 24 96 18 145 131 38 14 58 10 40 99 26 19 30 22 43 40 21 21 13 29 28 150 122 39 14 09 41 75 69 22 09 05 22 76 620 224 096 326 058 22 035 026 062 067 132 034 029 048 026 063 065 035 030 020 046 039 73 58 20 08 04 20 031 298 2189 53 196 582 096 26 038 459 165 170 165 310 308 860 174 143 314 297 173 343 048 165 175 80 64 24 11 06 25 036 168 98 93 29 11 08 28 043 205 030 020 095 018 018 045 046 161 024 017 048 046 041 092 bdl 029 031 82 72 24 10 06 24 033 181 024 41 35 13 06 52 04 14 020 091 012 62 61 19 08 06 21 032 140 023 040 012 026 030 031 119 025 031 023 059 061 120 029 027 062 059 060 119 bdl 043 043 55 541 062 032 131 021 008 029 107 84 27 12 33 07 24 040 225 029 033 bdl bdl bdl bdl bdl bdl bdl bdl 012 bdl 011 025 bdl bdl bdl bdl 021 bdl bdl bdl bdl bdl 017

470

SDB10

INAA

572

567

SDB11

INAA

496

466

SDB13

INAA

517

522

SDB16

INAA

550

474

SDB18

INAA

462

499

SDB20

INAA

545

540

SDB25

INAA

344

503

KERR et al.

Western Cordillera

CBU5

INAA

485

505

CBU8

INAA

482

517

CBU9

INAA

443

461

CBU11

INAA

482

480

CBU12

INAA

423

422

CBU13

INAA

524

483

CRETACEOUS COLOMBIAN BASALTS

687
198 56 48 47 138 107 125 102 31 33 143 19 105 30 451 2619 588 21 11 12 13 638 37 38 41 085 08 07 36 07 21

CBU14

INAA

464

497

PAN5

ICP-AES

PAN11

ICP-AES

COL1

INAA

463

492

VIJ1

INAA

476

480

VIJ4

INAA

483

395

Central Cordillera

AMA1

INAA

369

390

AMA8

INAA

396

455

AMA12 INAA

526

482

YAN1

INAA

507

502

YAN8

INAA

491

457

Southern Central
19 18 26 25 027 026 039 044 501 246 253 171 044 051 212 040 045 bdl

Cordillera

ROM4ii XRF

305

395

ROM5

ICP-AES

Standards (INAA)

BOB-1

335

563

measured

BOB-1

recomended

INAA, instrumental neutron activation analysis; XRF, X-ray uorescence; ICP-AES, inductively coupled plasma-atomic emission spectroscopy.

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

Table 5: Radiogenic isotope data for Cretaceous Colombian intrusive and volcanic rocks
87

Sr/86Sr m

(87Sr/86Sr)i

143

Nd/144Nd m

(Nd)i

Serrana de Baudo
SDB8 SDB11 SDB13 SDB16 SDB18 SDB20 0703039 10 0703854 11 0703073 11 0703128 13 0703429 9 0702998 11 0703224 11 0703246 11 0703855 11 0704572 9 0704714 12 0704909 10 0703686 11 0703297 12 0703795 10 0704404 11 0703420 10 0705033 10 0703856 11 0703563 13 0703023 0703687 0702991 0703065 0703380 0702918 0513081 5 0513048 7 0513073 7 0513070 5 0513067 8 0513073 7 0513053 11 0513047 14 0513020 20 0513019 5 0513014 5 0513063 5 0513049 5 0512953 4 0512909 5 0513056 13 0513027 6 0513078 4 866 806 816 836 848 818

Western Cordillera
VlJ1 VlJ4 CBU4 CBU9 CBU11 CBU12 CBU13 CBU14 COL1 0703207 0703227 0703751 0704500 0704650 0704606 0703610 0703208 0703753 813 801 754 746 756 831 814

Central Cordillera
AMA1 AMA8 AMA12 YAN1 YAN8 0704376 0703292 0704919 0703461 0703527 619 601 777 740 877

m, measured isotope ratio. i = 75 Ma for Serrana de Baudo; i = 90 Ma for Western Cordillera; i = 120 Ma for Central Cordillera.

intercalated sediments. For example, Bourgois et al. (1987) noted the presence of Cenomanian to Turonian (9788 Ma) microfaunas in limestones and cherts interbedded with basalts, 50 km NW of Medelln, whereas Barrero (1979) reported the occurrence of TuronianConiacian (9187 Ma) microfossils and ammonites from the Espinal Formation. As was noted above, the Macuchi Formation in Ecuador can be divided into two petrological provinces: an eastern, more basaltic province, and a western tu-rich andesitic province. The only palaeontological ages come from intercalated sediments in the western arc-derived province, where foraminifera of Eocene age have been reported (Henderson, 1979). This suggests that the more andesitic portion of the Macuchi Formation, like the Dabeiba Volcanic Arc sequence (sample AN1464 dated at 43104 Ma; see above), is signicantly younger than, and should not be correlated with, the basalts of the Western Cordillera. The tectonic signicance of these Eocene subduction-related volcanics along the western periphery of the Western Colombian and Ecuadorian Cordillera, and the Quebradagrande Complex, the

metasediments of which have yielded poorly preserved fossils ranging from Hauterivian to Albian [13597 Ma; summarized by Nivia (1987)] will be discussed in a later section. Unfortunately, no Central Cordillera basalts or dolerites were fresh enough or contained enough K2O (>01 wt %) to yield reliable 40Ar/39Ar ages. Additionally, no fossils have been found within the intercalated sediments (Aspden & McCourt, 1986). One of the few constraints on the age of the Amaime Formation comes from the fact that it is intruded by the Buga tonalitic batholith. This tonalite has yielded an Rb/Sr mineral isochron (on biotite and hornblende) of 994 Ma (McCourt et al., 1984), suggesting that the formation of the Amaime basaltic crust and its accretion onto the margin of NW Colombia must have occurred in the Early Cretaceous, well before 100 Ma. Thus the igneous rocks of the Central Cordillera would appear to be appreciably older than those of the Western Cordillera, which are in turn younger than the volcanic and intrusive rocks of the Serrana de Baudo. Nevertheless, the compositions of basalts from the Central Cordillera are similar to those

688

KERR et al.

CRETACEOUS COLOMBIAN BASALTS

of basalts from the other two Cordilleras, thus implying a similar petrogenetic history.

PETROGRAPHY AND MINERAL CHEMISTRY Central Cordillera


The basalts and dolerites consist of plagioclase, clinopyroxene and FeTi oxides, with secondary chlorite, iron oxides, zeolites, pumpellyite and quartz. Euhedral to subhedral olivine phenocrysts, which are altered to serpentine, are more common than in the Western Cordillera (see below), but aphyric basalts still predominate. The clinopyroxenes are relatively fresh and range in composition from diopside to augite (Nivia, 1987), whereas the fresher plagioclase crystals vary between An67 and An50 (Nivia, 1987). Textures range from plagioclase laths and euhedral clinopyroxene microphenocrysts in a groundmass of plagioclase and clinopyroxene spherulites, to ophitic to subophitic clinopyroxene enclosing or partially enclosing plagioclase laths, or occasionally a very ne-grained granular texture. An excellent summary of the petrography and mineral chemistry of the picrites and basalts from El Encenillo and Los Azules has been given by Spadea et al. (1989).

zeolites occur in small amounts within the groundmass and also in amygdales and sometimes in veinlets. In several samples altered olivine has been recognized in the groundmass, and occasionally altered (to chlorite and iddingsite) subhedral olivine microphenocrysts can be found. Ophitic textures are common, and these lavas can contain pyroxene and plagioclase phenocrysts, within a granular groundmass of plagioclase, pyroxene and secondary minerals. Occasionally, plagioclase and pyroxene form glomeroporphyritic clusters.

GEOCHEMISTRY Alteration and elemental mobility


The altered nature (up to prehnitepumpellyite grade) of the rocks from all three cordilleras means that before any petrological inferences can be drawn from the chemistry of the rocks, the possible chemical eects of subsolidus mobility of elements must be considered. Zirconium is widely regarded as being essentially immobile during low-grade alteration of basaltic rocks by hydrothermal uids (e.g. Humphris & Thompson, 1978; Gibson et al., 1982; Kerr, 1995) and so has been plotted against all the other minor and trace elements. A selection of these diagrams is shown in Fig. 4. Nb, Y and TiO2, which are also believed to be relatively immobile, produce good correlations against Zr. The minor dierences in ratio seen in the Y and Nb vs Zr plots are signicant, and are due either to variable degrees of partial mantle melting or to a heterogeneous mantle source region. In contrast, Ba, Sr, Rb and K2O display virtually no correlation with Zr, which strongly implies that, as in other altered basalts, these large ion lithophile elements have been extensively mobilized. Accordingly, variation in these elements will not be discussed further. All the other minor and trace elements are, like TiO2, Y and Nb, well correlated with Zr contents, implying relative immobility.

Western Cordillera
Most of the basalts and dolerites are ne- to mediumgrained holocrystalline rocks, with altered glass only being found in the outer skin of pillows. Textures vary from ophitic to subophitic with plagioclase laths poikilitically enclosed within anhedral clinopyroxenes, to intergranular and intersertal. The basalts and dolerites are generally aphyric but occasionally microphenocrysts of clinopyroxene can be found, set in a groundmass of variably altered plagioclase laths (An7250; Nivia, 1987) and anhedral to occasionally euhedral diopside and augite (En5035Wo4435Fs630; Nivia, 1987), with <5% FeTi oxides. Olivine phenocrysts (altered to chlorite and iddingsite) are occasionally found as pseudomorphs, but olivine is more common as an altered groundmass mineral. Other minor minerals are mostly secondary in origin, and include chloritized glass, chlorite, zeolites, pumpellyite, quartz and calcite.

Analytical results Serrana de Baudo


The volcanic and intrusive rocks of the Serrana de Baudo fall into two main groups. The rst group is represented by three samples (SDB2325) which come from the southernmost exposure of basalts on the Pacic coast of Colombia (Fig. 3a). The rocks of this small group, comprising one basaltic lava and two volcanic breccia blocks, are mildly alkaline, having more enriched incompatible trace element contents than the rest of the Serrana de Baudo rocks at similar MgO (58 wt %) levels (Figs 4 and 6). Figure 5 shows that the primitive

Serrana de Baudo
The basalts and dolerites are mostly holocrystalline, ne to medium grained, and contain augite and variably altered plagioclase laths (labradorite to andesine) with <5% anhedral FeTi oxides as the principal minerals. Secondary minerals such as chlorite, albite, quartz and

689

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

Fig. 4. Plots of Y, TiO2, Nb, Sr, K2O, Ba and Rb vs Zr. (Note that the large ion lithophile elements display virtually no correlation with Zr content.)

690

KERR et al.

CRETACEOUS COLOMBIAN BASALTS

major element composition [480517 wt % SiO2; 62 101 wt % MgO (Fig. 6); 92146 wt % Fe2O3(total)]. In terms of trace element contents, two basalts (SDB10 and -13) and one gabbro (SDB15) have slightly lower Zr, Y, Nb and TiO2 values than the rest of the basalts and gabbros, and lie slightly below the main fractionation trend (Fig. 6). This depletion is characteristic of all the other incompatible trace elements, as the primitive mantle-normalized pattern of SDB13 shows (Fig. 5a). These three samples (SDB10, -13 and -15) have lower Nb/Zr ratios (Fig. 7b) and higher 147Sm/144Nd (Fig. 8) than the rest of the Serrana de Baudo rocks. The rest of the basalts and gabbros of the Serrana de Baudo possess relatively at primitive mantle-normalized patterns (Fig. 5) ranging from 2 to 65 times primitive mantle, with the more evolved basalts having generally higher levels of incompatible trace elements. All the rocks (except SDB2225) are slightly depleted in light REE (LREE) [(La/Nd)pn 07510; Fig. 7a] and have (Sm/Yb)pn ratios ranging from 10 to 12 (pn denotes primitive mantlenormalized; Sun & McDonough, 1989). The isotopic data for the Serrana de Baudo samples are given in Table 5 and plotted in Fig. 8. ( Nd)i (where i=75 Ma age corrected value) span a narrow range from +80 to +87. Four of the samples also display a restricted range in ( 87Sr/86Sr)i from 070292 to 070307; however, SDB18 and SDB11 have elevated ( 87Sr/86Sr)i of 070338 and 070369, respectively.

Western Cordillera
The lavas and intrusives of the Western Cordillera are tholeiitic in character. Most of the rocks are basaltic in composition, and range from 48 to 53 wt % SiO2, 101 to 147 wt % Fe2O3(total) (Table 3) and 60 to 104 wt % MgO (Fig. 6). In addition, two basaltic andesites (PAN2 and -3) and an andesite PAN19, occur near the town of Vijes (Fig. 3b) in an area where rhyolitic dykes of uncertain origin cut the lava succession (Kerr et al., 1996c). These more evolved lavas have SiO2>53 wt % and MgO<51 wt %. A diorite (CBU13; CaliBuenaventura road) intruding the basaltic pile has only 15 wt % MgO (Fig. 6) and higher levels of incompatible trace elements than the basalts of the Western Cordillera (Figs 4, 5 and 6), and exhibits marked troughs in P and Ti on a primitive mantle-normalized diagram (Fig. 5). Plots of incompatible trace elements vs MgO (Fig. 6) show a typical negative correlation. This feature is also observed in Fig. 5, in which basalts with >60 wt % MgO (VIJ1, CBU11 and CBU12) have higher abundances of incompatible trace elements. All the basalts shown in Fig. 5 have essentially parallel and at primitive mantlenormalized patterns, and thus have relatively small ranges in incompatible trace element ratios (Figs 7 and 8c). As Fig. 7a shows, the basalts of the Western Cordillera are

Fig. 5. Primitive mantle-normalized (Sun & McDonough, 1989) multielement plots for the igneous rocks from (a) the Serrana de Baudo, (b) the Western Cordillera and (c) the Central Cordillera (ROM4ii, Los Azules; ROM5, El Encenillo).

mantle-normalized pattern of SDB25 is relatively steep, and that the La/Y and Nb/Zr ratios of this group are consistently higher (Fig. 7) in comparison with the rest of the Serrana de Baudo rocks. It is interesting to note that the more enriched (e-)basalts from Gorgona Island (Fig. 2), also part of the CaribbeanColombian oceanic plateau (Kerr et al., 1996a), possess similar incompatible trace element contents and ratios to SDB2325 (Fig. 7). The second group comprises the remaining volcanics and intrusives from the Serrana de Baudo. These tholeiitic basalts and gabbros exhibit a moderate range in

691

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

Fig. 6. Plots of Zr, Nb, TiO2 and Y vs MgO, showing the compositions of the Cretaceous igneous rocks of the three cordilleras in Colombia. (The oblong eld represents the El Encenillo samples.) Also shown are modelled (using the TRACE3 program; Nielsen, 1988) fractional crystallization trends for three starting compositions: a, CUR7 (picrite); b, ROM7 (picrite); c, SDB13 (basalt). Ticks on the fractional crystallization trends represent 10% crystallization intervals.

moderately depleted in the LREE [(La/Nd)pn 075095], with essentially at heavy REE (HREE) patterns [(Sm/ Yb)pn 1011]. The rocks of the Western Cordillera have ( Nd)i (i=90 Ma) in the range +75 to +81, whereas the diorite is slightly higher at +83. Three of the samples have ( 87Sr/86Sr)i in the range 0703207033, the diorite being distinctly higher at 07036 (Fig. 8a). However, like the basalts of the Serrana de Baudo, three of the basalts have elevated ( 87Sr/86Sr)i (>07045) with only a small concomitant decrease in ( Nd)i values.

Central Cordillera
The igneous rocks of the Central Cordillera are the most compositionally diverse of the three cordilleras, ranging from moderately enriched (relative to proposed primitive mantle compositions) in incompatible trace elements to relatively depleted (Fig. 5). The rocks can be divided into two broad groups: the moderately incompatible trace element-enriched basalts and picrites from Los Azules and El Encenillo, and the mostly less enriched basalts

and dolerites of the Amaime Formation and the Yana area. Lavas from Los Azules are tholeiitic, whereas those from El Encenillo are more alkaline. The non-cumulate lavas range from 71 to 166 wt % MgO (Fig. 6), the higher values falling within the range of calculated MgO content (128173 wt %) (Spadea et al., 1989) for a liquid in equilibrium with Fo89 olivine (maximum in Los Azules picrites). Therefore ROM7, the sample containing 166 wt % MgO, may represent a potential primary magma for the Los Azules complex. In terms of trace elements, the two analysed lavas from El Encenillo are signicantly more enriched in incompatible trace elements (at equivalent MgO content) than the lavas from Los Azules (Fig. 6). This is also reected in the steeper primitive mantle-normalized trace element patterns for the El Encenillo lavas (Fig. 5) and incompatible trace element ratio plots (Fig. 7), with the El Encenillo lavas having, for example, higher La/Y and Nb/Zr than the Los Azules lavas. The tholeiitic basalts and dolerites of the Amaime Formation and Yana area range in MgO content from

692

KERR et al.

CRETACEOUS COLOMBIAN BASALTS

exception to this is a group of basalts from the more southerly parts of the Amaime Formation (AMA36 and FLO14), which possess higher Nb/Zr ratios than the rest of the Amaime and Yana basalts (Fig. 7). This enrichment seems only to aect Nb contents, as ratios not involving Nb (e.g. Ti/Zr) are the same as for the rest of the basalts. Two of the basalts from the Amaime Formation (AMA8 and -11) are relatively enriched in incompatible trace element contents and ratios (Figs 5, 6 and 7), whereas one basalt (AMA12) has been found which possesses more depleted levels of incompatible trace elements (Figs 5, 6 and 7). In terms of ( Nd)i (i=120 Ma), the lavas span a comparatively wide range from +60 in AMA8, one of the most enriched Amaime basalts, to +81 in YAN-8 (Fig. 8). The most trace element depleted Amaime basalt (AMA12) has an ( Nd)i of +77. As with the basalts of the Western Cordillera, ( 87Sr/86Sr)i appears be decoupled from ( Nd)i, with several of the basalts having ( 87Sr/ 86 Sr)i>07043, but most lie in the range 0703207035.

DISCUSSION Original tectonic setting and relative ages of the Colombian basalts Tectonic setting
The chemical data presented can help to resolve the long-standing controversy regarding the tectonic setting of the Colombian mac volcanic terranes, namely, whether they represent a subduction zone environment, ocean oor or oceanic plateau. This problem can be addressed in two ways: (1) Mature subduction zone volcanic sequences are characterized by abundant explosion-derived tus and lavas of andesitic composition, as seen in recent Colombian subduction-related volcanoes (Marriner & Millward, 1984). However, the Cretaceous volcanic terranes are predominantly basaltic, and although pyroclastic deposits do occur, they are relatively rare and are mostly basaltic in composition.

50 to 103 wt % (Fig. 6). Most of the lavas possess relatively at primitive mantle-normalized trace element patterns (Fig. 5), and their incompatible trace element ratios are similar to those from the Western Cordillera and from Serrana de Baudo (Fig. 7). An interesting

Fig. 7. Incompatible element ratio plots displaying the Cretaceous Colombian igneous rocks along with elds for lavas from Gorgona, Curacao, Iceland and the Ontong Java plateau: (a) (Sm/Yb)pn vs (La/ Nd)pn; (b) Ti/Zr vs Nb/Zr; (c) Nb/Y vs La/Y [also shown in (c) are Cretaceous arc lavas from Bonaire Island in the southern Caribbean and Recent Colombian arc-derived lavas]. Symbols are as in Fig. 4, apart from those which have been separately denoted. Data sources: GorgonaAitken & Echeverra (1984), Arndt et al. (1997), Kerr et al. (1996a); CuracaoKerr et al. (1996b); IcelandHemond et al. (1993); Ontong JavaMahoney et al. (1993); BonaireG. Klaver (unpublished data, 1979); Recent Colombian lavasMarriner & Millward (1984); the eld for Cretaceous Colombian lavas in (c) is taken from Nivia (1987).

693

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

Fig. 8 . (a) ( Nd)i vs ( 87Sr/86Sr)i, (b) ( 143Nd/144Nd)i vs (Sm/Yb)pn, (c) ( 143Nd/144Nd)i vs 147Sm/144Nd. i denotes initial age corrected values. Symbols are as in Fig. 4, except those separately marked. Data sources: HaitiSen et al. (1988); GalapagosWhite et al. (1993); all other data sources are as for Fig. 7.

694

KERR et al.

CRETACEOUS COLOMBIAN BASALTS

(2) As Fig. 5 shows, these Cretaceous volcanic rocks do not possess the characteristic Nb depletion (with corresponding LREE enrichment), which results in a negative Nb anomaly on multi-element normalized plots of subduction-related lavas. Figure 7c conrms that the Cretaceous Colombian basalts and picrites have consistently higher Nb/La ratios (relative to Y) than both Recent Colombian volcanic arc rocks (Marriner & Millward, 1984) and Cretaceous subduction-related lavas from Bonaire in the southern Caribbean (G. Klaver, unpublished data, 1979). Therefore, both eld and geochemical evidence militate strongly against a subduction-related origin for the Cretaceous Colombian volcanic terranes. Moreover, the predominance of basalts with relatively at primitive normalized trace element patterns, and the not uncommon occurrence of lavas with more picritic compositions, are more consistent with an origin in plumederived oceanic plateau, rather than in normal ocean oor. This is further supported by Figs 7 and 8, which reveal that lavas drilled from the Ontong Java plateau (Mahoney et al., 1993) possess similar trace element and isotopic ratios to most of the basalts and picrites from Colombia. However, there is a greater chemical variability within the Colombian lavas, and possible explanations for this will be discussed in a later section. The ~90 Ma basalts and picrites from Curacao are part of the CaribbeanColombian oceanic plateau (Kerr et al., 1996b), and the close similarity in isotopic and trace element ratios between these Curacao lavas and the Colombian basalts and picrites (Figs 7 and 8) supports an origin within the Caribbean oceanic plateau.

Timing of volcanism
The volcanism associated with the Colombian portion of the Caribbean oceanic plateau appears to be of three distinct ages: two well-dated events ( 40Ar/39Ar; fossils from intercalated sediments), one of Late Cenomanian Turonian (8892 Ma) and another of Late Campanian Early Maastrichtian (7278 Ma), plus another eruptive episode older than 100 Ma. Other fragments of Early Cretaceous oceanic plateau material have been reported from elsewhere in the Caribbean region, for example, in the Duarte Complex of Hispaniola (Lapierre et al., 1997) and Cuba (Iturralde-Vinent, 1994). Similarly, more evidence for the younger plateau eruptive event within the Caribbean region (possibly contemporaneous with the formation of the Serrana de Baudo rocks) is also beginning to accumulate. The westernmost Caribbean site drilled during DSDP Leg 15, Site 152 (Fig. 1) encountered basalt with the chemical characteristics of an oceanic plateau which contained fragments of Campanian (8374 Ma) limestone (Donnelly et al., 1973). Recent Ocean Drilling Program (ODP) drilling in the Caribbean during

Leg 165, at Site 1001 (40 km WSW of Site 152), encountered Middle Campanian limestone mixed with clay and basaltic ashlapilli overlying, and grading down into, plateau basalt (Pearce & Pearson, 1996). Thus the presence of Campanian limestone intimately associated with the uppermost volcanics at both these sites places a maximum age of 83 Ma on the underlying basalt. Additionally, a dolerite sill intruding the upper part of the Curacao lava succession has recently been dated, using 40Ar/39Ar step-heating, at 75819 Ma (C. Sinton, unpublished data, 1996). It is interesting to note that, as in Colombia, where the westernmost basalts (Serrana de Baudo) are the youngest (7278 Ma), so it is the westernmost drilled holes in the Caribbean Sea that have also produced the youngest (Campanian) ages. These observations support a younger episode of CaribbeanColombian plateau volcanism mostly in the west of the province, consistent with eastward movement of the plate(s) above a stationary plume (possibly Galapagos). The CaribbeanColombian oceanic plateau is not unique in this regard, because many of the worlds plumerelated LIPs display at least two distinct periods of major eruptive eruptions, separated by between 20 and 90 m.y. (Bercovici & Mahoney, 1994). The Ontong Java plateau, like the CaribbeanColombian oceanic plateau, appears to have peaks of volcanic activity at about 120 and 90 Ma (Mahoney et al., 1993). However, perhaps a more suitable analogue for the CaribbeanColombian Cretaceous oceanic plateauwhich appears to have been volcanically active over ~40 Mais the long-lived volcanism associated with the Icelandic plume. Interestingly, there are other similarities in both chemistry and tectonic setting between the CaribbeanColombian and Icelandic basalts, which will be explored in a later section. Finally, whereas it is possible that the earlier (>100 Ma) oceanic plateau forming event in the Caribbean Colombian province is linked with the same plume as that which produced the 90 and 75 Ma events, we caution that there is no way of discounting the possibility that the earlier event was linked to an entirely dierent plume.

Petrogenetic aspects The high 87Sr/86Sr puzzle


Arguably, the most enigmatic feature of the chemistry of these lavas is their tendency to high ( 87Sr/86Sr)i values relative to ( Nd)i (Fig. 8a), which do not appear to correlate with any other chemical parameter. Repeated leaching of powders from several of these samples failed to signicantly reduce the high 87Sr/86Sr more than just a single leaching. Within the CaribbeanColombian Cretaceous igneous province these high 87Sr/86Sr values are not

695

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

unique, and they have been found in well-leached komatiites and picrites from Gorgona Island (Aitken & Echeverra, 1984), basalts from Nicoya in Costa Rica (F. Hau, personal communication, 1996) and basalts from Curacao (Kerr et al., 1996b). The last leaching ex periments support the view that these high 87Sr/86Sr values are magmatic in origin and are not caused by sub-solidus hydrothermal alteration (Kerr et al., 1996b). Accordingly, Kerr et al. (1996b) proposed that the high 87 Sr/86Sr values could result from assimilation of altered oceanic crust, a process which has also been proposed to account for the moderately elevated 87Sr/86Sr values in Icelandic lavas (e.g. Hemond et al., 1993).

Fractional crystallization
Although the vast majority of rocks from the three volcanic terranes are basaltic in composition, these may not be primary mantle melts. Basalts from the Ontong Java plateau not only have similar trace element ratios to the Cretaceous Colombian basalts, but display a similar range in MgO content (Mahoney et al., 1993). Farnetani et al. (1996) recently proposed that high seismic velocities near Moho levels within the Ontong Java plateau represent crystal cumulates (chiey olivine) from the largescale fractionation of primary picritic magmas to produce the erupted basalts. The Cretaceous igneous association in both the Western and Central Cordilleras is in part composed of mac and ultramac cumulates (Spadea et al., 1987; Nivia, 1996). The chemistry of these rocks reveals that they are genetically related to the basalts of their respective cordillera (A. C. Kerr et al., unpublished data, 1996). These mac and ultramac cumulates may be akin to the high seismic velocity layers found at the base of the crust in the Ontong Java plateau, and so may represent the lower-crustal levels of the 8792 Ma phase of CaribbeanColombian oceanic plateau. It is likely that the Cretaceous Colombian basalts were mostly derived from more picritic melts which ponded and fractionated deep within the plateau to produce the ultramac cumulates and lower-MgO basaltic liquids. The relatively widespread occurrence of 8792 Ma high-MgO lavas (picrites and komatiites) throughout the Caribbean Colombian province (Kerr et al., 1997) also implies that high-MgO melts were produced by a large proportion of the proposed plume head. Thus it is not unreasonable to consider that the Colombian basalts are derived from more picritic parental magmas. In light of these eld observations, we have attempted to reproduce the composition of the Colombian basalts using the TRACE3 computer program of Nielsen (1988), which, given a starting magma composition, calculates mineral and residual magma compositions during fractional crystallization. The residual magma compositions obtained from this modelling are presented in Fig. 6. We

have used three dierent starting magma compositions: a picrite (CUR7) containing 20% MgO from Curacao (Fig. 1); a slightly more enriched picrite (ROM7) from Los Azules (166% MgO); and one of the more depleted, higher-MgO basalts from the Serrana de Baudo (SDB13; 101 wt % MgO). Not surprisingly, the rst phase to crystallize in all three models is olivine (Cr-spinel); then, when 89 wt % MgO is reached, plagioclase and clinopyroxene join the fractionating assemblage, to be followed by FeTi oxide at <5 wt % MgO. The results (Fig. 6) show that, between them, the three dierent parental magmas could fractionate to produce the incompatible trace element compositions of most of the Cretaceous Colombian basalts. However, some of the basalts are too enriched or too depleted in incompatible trace elements to be fractionates of any of the three magma starting compositions. As will be discussed in the following section, this wide range in incompatible trace element contents at a given MgO (Fig. 6) reects variable degrees of partial mantle melting and/or a heterogeneous mantle source region. The composition of the diorite (CBU13) from the Western Cordillera appears to be modelled fairly well by the TRACE3 program. More than 80% fractionation from a picritic parent is required to reproduce the composition of this diorite, which in addition to olivine, plagioclase, clinopyroxene and FeTi oxide, seems to have fractionated minor apatite (as evidenced by the negative phosphorus anomaly in Fig. 3b). Additionally, it should be noted that the marked positive NbTa spike in Fig. 3b militates strongly against an arc-derived origin for this intrusion.

Mantle sources and melting


The variations in Nd isotope and incompatible trace element ratios (Figs 7 and 8) unequivocally demonstrate that the plume source region of the Cretaceous Colombian basalts was heterogeneous. Although some lavas are more enriched in incompatible trace elements compared with others, the positive Nd values mean that all the lavas have been derived from a mantle source region with a long-term depletion in the LREE. Thus both the enriched and chondritic LREE patterns (Fig. 5) of some of the Colombian basalts (and for that matter, most Ontong Java plateau basalts) must represent a relatively recent enrichment, probably during the melting process. The range of (La/Nd)pn (1) found in the majority of the Colombian basalts (Fig. 7a) obviously reects the depleted nature of their source regions. Nevertheless, this depletion in incompatible trace elements is subtly dierent in character from that observed in present-day East Pacic Rise mid-ocean ridge basalt (MORB), and Fig. 9a elegantly reinforces this point. This diagram was originally used by Fitton et al. (1997) to distinguish incompatible element-depleted, plume-derived Icelandic

696

KERR et al.

CRETACEOUS COLOMBIAN BASALTS

Fig. 9. (a) Log plot of Nb/Y vs Zr/Y [after Fitton et al. (1997)], showing the compositions of the Cretaceous Colombian igneous rocks. Lavas from the neovolcanic zones of Iceland fall between the two parallel lines, whereas East Pacic Rise MORBs [from Mahoney et al. (1994)] plot below this eld. (b) Nb/Y vs Zr with modelled partial fractional melting curves for a garnet lherzolite (Gnt lz), spinel lherzolite (Sp lz) and a 50:50 spinelgarnet mixture (SpGnt lz). Numbered ticks on the melting curves indicate percentage of partial melting (see text for more details). Near-horizontal arrows represent several possible fractional crystallization trajectories (based on the modelling shown in Fig. 6). Numbered ticks on the lowermost trajectory indicate the composition of the residual liquid after 50 and 70% crystallization from a picritic parent.

basalts from North Atlantic MORB. In Fig. 9a, our new data from Colombia are plotted along with the tramlines of Fitton et al. (1997) (between which all the plumederived Icelandic lavas plot), and elds for East Pacic Rise MORB and the Ontong Java plateau (Mahoney et al., 1993, 1994). Virtually all the Colombian basalts fall within the Icelandic tramlines, strongly implying derivation from a plume source region. Even the most depleted Colombian rocks (SDB13 and AMA12) do not appear to originate from a depleted upper-mantle (MORB-source) region; rather, they seem to be derived from a depleted (relative to Bulk Earth) source region within the plume. In modelling the mantle melting processes several factors need to be considered: (1) The plume source region was isotopically heterogeneous, comprising at least two components which both

appear to have a long-term history of depletion, with one component being slightly more enriched (lower Nd) than the other. (2) The Colombian basalts are not primary melts from the mantle, but may have been trapped and fractionated from picritic melts in magma chambers en route to the surface. However, as well as fractionation, the entrapment of picritic liquids in magma chambers will also serve to homogenize individual heterogeneous melt fractions. This suggestion is reinforced by the fact that the ColombianCaribbean Cretaceous high-MgO lavas are, as a group, signicantly more heterogeneous than their associated basalts (Kerr et al., 1996a, c). Thus, although many of the Cretaceous basalts in Colombia (and in other oceanic plateaux) are relatively homogeneous, this could just reect mixing of compositionally distinct magmas derived from a heterogeneous plume source region by variable degrees of melting. An LREE-depleted mantle source composition was used in the modelling of pooled fractional melting. Three dierent mineral assemblages were useda garnet lherzolite, a spinel lherzolite and a 50:50 mixture of these two mineralogiesto simulate melting in the garnetspinel transition zone [source compositions, mineral proportions and partition coecients are from McKenzie & ONions (1991)]. The results of the mantle melt modelling are presented in Fig. 9b, along with fractional crystallization trends that show a slight increase in Nb/Y ratios as the magmas fractionate. Thus, when fractional crystallization has been accounted for, it can be inferred that many of the Colombian basalts result from fairly extensive degrees (~20%) of partial mantle melting, or that they were derived from a more depleted source than that used in the melt modelling. As Fig. 9b shows, at 20% melting the dierences in incompatible trace element contents and ratios between the three mantle mineral assemblages are signicantly less than at smaller degrees of melting. It thus becomes more dicult to assess the approximate depth of melting of the erupted lavas, a diculty which is compounded by the possibility of mixing melts derived from both the garnet and spinel lherzolite stability elds. However, melting seems to have occurred over a rather wide depth range (polybaric melting) from relatively deep, mostly within the garnet lherzolite eld, to shallower melting of spinel lherzolite. Figure 9b shows that basalts with Nb/Y>02 probably had a melt input from a mantle source region containing some garnet. These more incompatible element-enriched basalts also generally have lower ( 143Nd/144Nd)i values (Fig. 8b and c), implying that they are derived from a more enriched source region, or contain a higher proportion of enriched component. The negative correlation between ( 143Nd/144Nd)i and (Sm/Yb)pn displayed by the samples from the Western and Central Cordilleras reveals that

697

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

the more enriched melts are derived from deeper source region and vice versa. This source heterogeneity and negative correlation between the degree of enrichment and depth of melting support the recent models of Kerr et al. (1995) and Arndt et al. (1997). These models have proposed that most mantle plumes are heterogeneous and consist predominantly of a relatively refractory depleted matrix component, with a small proportion (p10%) of more enriched and more fusible streaks or blobs. Thus deeper, and therefore smaller, extents of mantle melting will preferentially sample these more enriched streaks, whereas at shallower depths, more extensive melting means that a greater proportion of the magma produced will be derived from the more depleted matrix. A feature worth stressing is that many of the Colombian basalts, and indeed basalts from many other oceanic plateaux, particularly Ontong Java, possess positive Nd values but paradoxically, have near-chondritic ratios of incompatible trace elements. This problem has become known as the plateau paradox (Babbs et al., 1996). One possible explanation for this paradox may be that mixing of small degree (<2%) relatively LREE-enriched melts (with Nd>0) from the garnet lherzolite stability eld, with larger degree (~20%) LREE-depleted melts, predominantly derived from the spinel lherzolite eld, has occurred. Such mixing of dierent degree melts could occur either in the mantle plume source region, or perhaps more likely within crustal or lithospheric magma chambers en route to the surface, and has the potential to produce magmas with near-chondritic incompatible trace element ratios but with positive Nd values. A key issue is whether the sequence of late Cretaceous komatiites, picrites and basalts on Gorgona are not only an integral part of the CaribbeanColombian Cretaceous oceanic plateau (Storey et al., 1991; Kerr et al., 1996a) but can be correlated with, and are also a southward continuation of, the basaltic sequence of the Serrana de Baudo (McGeary & Ben-Avraham, 1986). The new chemical data presented above support this latter proposal, in that the three more enriched basalts (from the southernmost exposure of the Serrana de Baudo) are very similar to the most enriched e-basalts (Kerr et al., 1996a) from Gorgona, whereas the rest of the Serrana de Baudo samples possess incompatible trace element ratios which are within the range displayed by the rest of the Gorgona basalts. However, the older ages (8688 Ma; Kerr et al., 1996a) for the Gorgona lavas and intrusives, compared with the Serrana de Baudo (7278 Ma) appear to rule out the idea that the basalts of the Serrana de Baudo are a continuation of the sequence found on Gorgona. As well as becoming younger towards the west, the lavas and intrusives of the three cordilleras also generally become more depleted; for example, the average Nd for the Central Cordillera is +72, for the Western Cordillera

+78 and for the Serrana de Baudo +83. In addition to this, the younger lavas drilled by DSDP Leg 15 at Site 152 are similarly the most elementally and isotopically depleted lavas sampled during Leg 15 (Kerr et al., 1997; G. F. Marriner & A. D. Saunders, unpublished data, 1987). Thus if one single plume is responsible for the volcanism in the CaribbeanColombian oceanic plateau, then it appears to have become more depleted with time.

Is the Caribbean Plateau an Icelandic analogue?


Recently, linear NESW and eastwest long-wavelength magnetic anomalies have been discovered over the Venezuelan and Colombian Basins (Fig. 1) in the Caribbean (Hall, 1995). It has been suggested by Hall (1995) that these anomalies may have resulted from an Early Cretaceous phase of seaoor spreading at the Farallon PacicPhoenix triple junction, at or near which the CaribbeanColombian Oceanic Plateau formed in the Late Cretaceous. It is interesting to note that the Ga lapagos plume, which may represent the present-day expression of the plume responsible for the Caribbean Colombian oceanic plateau, is currently situated close to the Galapagos Spreading Centre. Indeed, the occurrence of a plume track on both the Cocos and Nazca Plates suggests that it was only within the last 5 m.y. that the ridge moved northwards away from the Galapagos hotspot (Hay, 1977), so for a large proportion of its history the Galapagos plume may have impinged on the base of the lithosphere at, or close to, an oceanic spreading centre. Thus, as well as both the Galapagos and Icelandic plumes having a long history of activity, it also seems that the two plumes may have been situated below a mid-ocean ridge for a considerable part of their histories. It is therefore a constructive exercise to amplify the initial studies of Nivia (1987) and compare the chemistry of the CaribbeanColombian oceanic plateau with that of lavas produced by the Icelandic plume. In terms of trace element and radiogenic ratios, the Icelandic lavas [from Hemond et al. (1993)] span a very similar range to the Cretaceous Colombian lavas and intrusives reported in this paper (Figs 7, 8 and 9). The data in Fig. 8b and c show that, at a given incompatible trace element ratio, the Icelandic lavas possess slightly higher ( Nd)i. Nevertheless, the basic similarity between the Icelandic and Colombian lavas adds further weight to the supposition that the CaribbeanColombian oceanic plateau formed at, or near an oceanic spreading centre. Like presentday Iceland, the more enriched lavas in the Caribbean Cretaceous oceanic plateau could have been produced by smaller degrees of melting below thicker lithosphere, further away from the spreading centre (see Hards et al.,

698

KERR et al.

CRETACEOUS COLOMBIAN BASALTS

1995). Similarly, the more depleted basalts found in Colombia could represent the equivalent of those depleted Icelandic picrites and basalts which are produced by more extensive shallower melting close to the ridge.

The signicance of Early Tertiary and Early Cretaceous arc volcanism in Colombia
Although not considered in detail in this paper, it is nevertheless important to assess the tectonic signicance of the subduction-related Early Cretaceous Quebradagrande Complex and the Eocene Dabeiba Volcanic Arc, along with the Eocene arc lavas and pyroclastics of the Macuchi Formation in Ecuador. It is probable that the Quebradagrande complex formed in either a marginal basin or a continental volcanic arc setting (A. Nivia, in preparation) before ~120 Ma. The radiometric age constraints imposed by the Buga batholith mean that the Early Cretaceous Amaime Formation must have been obducted onto the Colombian margin before 100 Ma. The Quebradagrande Complex may thus represent Early Cretaceous pre-collision magmatism on the Colombian continental margin. The obduction of the 7590 Ma lavas and intrusives of the Cordillera and the Serrana de Baudo appears to have occurred in the Late Cretaceous to Early Tertiary (Kerr et al., 1997). It is possible that the volcanics of the Dabeiba arc were formed during the obduction of the 7590 Ma portion of the Caribbean Colombian oceanic plateau onto the Colombian continental margin. If this is the case, then the date of the obduction may well be late Eocene. However, more 40 Ar/39Ar step heating work on other possible obduction related pegmatite veins intruding the BolvarRo Frio ultramac complex is currently under way, and these dates may help to better constrain the possible age of the obduction event.

part of the plume. Importantly, neither depleted plume component can be attributed to entrainment of depleted upper asthenospheric (MORB-source) mantle. (3) Fractional melt modelling using a depleted mantle source composition suggests that most of the Western Colombian lavas require a small melt input from a mantle source region containing garnet as a residual phase, along with more extensive melting in the spinel lherzolite stability eld. After allowing for fractional crystallization it appears that these Cretaceous Colombian lavas and intrusives are the result of ~20% partial mantle melting. However, mixing of magmas generated at dierent depths possibly in Moho-level magma chambers during fractionation will also result in mixing, thus masking mantlederived heterogeneities. (4) Three distinct ages of volcanic activity can be identied in the Colombian portion of the Caribbean Colombian oceanic plateau (and also in the Caribbean part of the province). The westernmost accreted belt, the Serrana de Baudo, is the youngest at 7378 Ma, the Western Cordillera rocks are ~90 Ma old, whereas the volcanic sequence of the easternmost Central Cordillera is Early Cretaceous in age. The geochemical signatures of the igneous rocks of the three cordilleras also become more depleted from east to west.

ACKNOWLEDGEMENTS
We are grateful to the Natural Environmental Research Council (UK) for supporting this work through Grants GR9/583A and GR3/8984 to J.T. and A.D.S., and to the Leverhulme Trust for a fellowship to A.C.K. We also thank INGEOMINAS for invaluable logistic support in Colombia, Nick Marsh for assistance with XRF analysis at Leicester, and Marion Weber for her help with Spanish spelling. Ray MacDonald, Henriette Lapierre and Christian Coulon are thanked for their constructive reviews of the manuscript. XRF and radiogenic isotope laboratories at Royal Holloway are London University intercollegiate facilities.

CONCLUSIONS
(1) The accreted Cretaceous basaltic and picritic terranes found in western Colombia appear to have formed part of an oceanic plateau linked to a mantle plume. These terranes represent the southern portion of the large late-Cretaceous Caribbean oceanic plateau, which was obducted and imbricated against the western margin of the northern South American continent. (2) It is unlikely that the Cretaceous plateau-derived basalts of western Colombia are primary mantle melts. Rather, they may have fractionated from more picritic parental magmas. These picritic partial mantle melts must have been derived from a heterogeneous mantle plume source region containing at least two components. Both components display time integrated depletion (positive Nd), and both appear to have been an integral

REFERENCES
Aitken, B. G. & Echeverra, L. M., 1984. Petrology and geochemistry of komatiites and tholeiites from Gorgona Island, Colombia. Contributions to Mineralogy and Petrology 86, 94105. Alvarez, E. & Gonzalez, H., 1978. Geologa y Geoqumica del Cu adrangulo I-7 Urrao, No. 1761. Medelln, Colombia: IN GEOMINAS, p. 370. Arndt, N. T., Czamanske, G. K., Wooden, J. L. & Fedorenko, V. A., 1993. Mantle and crustal contributions to continental ood basalt volcanism. Tectonophysics 223, 3952. Arndt, N. T., Kerr, A. C. & Tarney, J., 1997. Dierentiation in plume heads: the formation of Gorgona komatiites and basalts. Earth and Planetary Science Letters 146, 289301.

699

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

Aspden, J. A., 1984. The geology of the Western Cordillera, Department of Valle, Colombia (sheets 261, 278, 280 and 299). Cali, Colombia: INGEOMINASMission Britanica (British Geological Survey). Aspden, J. A. & McCourt, W. J., 1986. Mesozoic oceanic terrane in the central Andes of Colombia. Geology 14, 415418. Aspden, J. A., McCourt, W. J. & Brook, M., 1987. Geometrical control of subduction-related magmatism: the Mesozoic and Cenozoic plutonic history of western Colombia. Journal of the Geological Society, London 144, 893905. Babbs, T., Saunders, A. D., Kerr, A. C. & Kempton, P. D., 1996. Composition of the lower mantle: the plateau paradox. Journal of Conference Abstracts 1, 32. Barrero, D., 1979. Geology of the Central Western Cordillera, west of Buga and Roldanillo, Colombia. Publicaciones Geologicas Especials del INGEOMINAS 4, 175. Ben-Avraham, Z., Nur, A., Jones, D. & Cox, A., 1981. Continental accretion: from oceanic plateaux to allochthonous terranes. Science 213, 4754. Bence, A. E., Papike, J. J. & Ayuso, R. A., 1975. Petrology of submarine basalts from the Central Caribbean: DSDP Leg 15. Journal of Geophysical Research 80, 47754804. Bercovici, D. & Mahoney, J., 1994. Double ood basalts and plume head separation at the 660-km discontinuity. Science 266, 13671369. Bourgois, J., Calle, B., Tournon, J., Azema, J. & Toussaint, J. F., 1982. The Andean ophiolitic megastructure in the BugaBuenaventura traverse (Western Cordillera, Valle, Colombia). Tectonophysics 82, 207229. Bourgois, J., Toussaint, J. F., Gonzales, H., Azema, J., Calle, B., Desmet, A., Murcia, L. S., Acevedo, A. P., Parra, E. & Tournon, J., 1987. Geological history of the Cretaceous ophiolitic complexes of NW South America (Colombian Andes). Tectonophysics 143, 307327. Burke, K., 1988. Tectonic evolution of the Caribbean. Annual Review of Earth and Planetary Sciences 16, 201230. Burke, K., Fox, P. J. & Sengor, M. C., 1978. Buoyant ocean oor and the origin of the Caribbean. Journal of Geophysical Research 83, 39493954. Carlson, R. L., Christensen, N. I. & Moore, R. P., 1980. Anomalous crustal structures in ocean basins: continental fragments and oceanic plateaux. Earth and Planetary Science Letters 51, 171180. Case, J. E., Duran, S. L. G., Lopez, A. & Moore, W. R., 1971. Tectonic investigations on western Colombia and eastern Panama. Geological Society of America Bulletin 82, 26852712. Case, J. E., Barnes, J., Pars, G., Gonzalez, I. H. & Vina, A., 1973. Trans-Andean geophysical prole, southern Colombia. Geological Society of America Bulletin 84, 28952904. Cloos, M., 1993. Lithospheric buoyancy and collisional orogenesis: subduction of oceanic plateaux, continental margins, island arcs, spreading ridges, and seamounts. Geological Society of America Bulletin 105, 715737. Con, M. F. & Eldholm, O., 1993. Scratching the surface: estimating dimensions of large igneous provinces. Geology 21, 515518. Dalrymple, G. B., Clague, D. A., Vallier, T. L. & Menard, H. W., 1988. 40Ar39Ar age, petrology, and tectonic signicance of some seamounts in the Gulf of Alaska. Geophysical Monograph, American Geophysical Union 43, 297315. Donnelly, T. W., 1973. Late Cretaceous basalts from the Caribbean, a possible ood basalt province of vast size. EOS 54, 1004. Donnelly, T. W., Melson, W., Kay, R. & Rogers, J. J. W., 1973. Basalts and dolerites of late Cretaceous age from the central Caribbean. In: Initial Reports of the Deep Sea Drilling Project, 15. Washington, DC: US Government Printing Oce, pp. 9891004. Donnelly, T. W., et al., 1990. History and tectonic setting of Caribbean magmatism. In: Dengo, G. & Case, J. E. (eds) The Geology of North

America, Volume H, The Caribbean Region. Boulder, CO: Geological Society of America, pp. 339374. Duncan, R. A. & Hargraves, R. B., 1984. Plate tectonic evolution of the Caribbean region in the mantle reference frame. In: Bonini, W. E., Hargraves, R. B. & Shagam, R. (eds) The CaribbeanSouth America Plate Boundary and Regional Tectonics. Geological Society of America Memoir 162, 8193. Duncan, R. A. & Hargraves, R. B., 1990. 40Ar/39Ar geochronology of basement rocks from the Mascarene Plateau, Chagos Bank, and Maldives Ridge. In: Duncan, R. A., Backman, J. & Peterson, L. C. (eds) Proceedings of the Ocean Drilling Program: Scientic Results, 115. College Station, TX: Ocean Drilling Program, pp. 4352. Duncan, R. A. & Hogan, L. G., 1994. Radiometric dating of young MORB using the 40Ar39Ar incremental heating method. Geophysical Research Letters 21, 19271930. Echeverra, L. M., 1980. Tertiary or Mesozoic komatiites from Gorgona Island, Colombia: eld relations and geochemistry. Contributions to Mineralogy and Petrology 73, 253266. Edgar, N. T., Ewing, J. I. & Hennion, J., 1971. Seismic refraction and reection in the Caribbean Sea. American Association of Petroleum Geology Bulletin 55, 833870. Farnetani, C. G., Richards, M. A. & Ghiorso, M. S., 1996. Petrological models of magma evolution and deep crustal structure beneath hotspots and ood basalt provinces. Earth and Planetary Science Letters 143, 8194. Feininger, T. & Bristow, C. R., 1980. Cretaceous and Paleogene geologic history of coastal Ecuador. Geologische Rundschau 69, 849874. Fitton, J. G., Saunders, A. D., Larsen, H. C., Hardarson, B. S. & Norry, M. J., 1997. Volcanic rocks from the East Greenland margin at 63N: composition, petrogenesis and mantle sources. In: Saunders, A. D., Larsen, H. C., Wise, S., & Allen, J. R. (eds) Proceedings of the Ocean Drilling Program, Scientic Results, 152. College Station, TX: Ocean Drilling Program. In press. Gallagher, K. & Hawkesworth, C. J., 1992. Dehydration melting and the generation of continental ood basalts. Nature 358, 5759. Gansser, A., 1973. Facts and theories on the Andes. Journal of the Geological Society, London 129, 93131. Gibson, I. L., Kirkpatrick, R. J., Emmerman, R., Schmincke, H.-U., Pritchard, G., Okay, P. J., Thorpe, R. S. & Marriner, G. F., 1982. The trace element composition of the lavas and dykes from a 3 km vertical section through the lava pile of Eastern Iceland. Journal of Geophysical Research 87, 65326546. Gibson, S. A., Thompson, R. N., Dickin, A. P. & Leonardos, O. H., 1995. High-Ti and Low-Ti mac potassic magmas: key to plumelithosphere interactions and continental ood basalt genesis. Earth and Planetary Science Letters 136, 134149. Goossens, P. J. & Rose, W. I., 1973. Chemical composition and age determination of tholeiitic rocks in the Basic Igneous Complex, Ecuador. Geological Society of America Bulletin 84, 10431052. Goossens, P. J., Rose, W. I. & Flores, D., 1977. Geochemistry of tholeiites of the Basic Igneous Complex of NW South America. Geological Society of America Bulletin 88, 17111720. Grosser, J. R., 1989. Geotectonic evolution of the Western Cordillera of Colombia: new aspects from geochemical data on volcanic rocks. Journal of South American Earth Sciences 2, 359369. Hall, S. A., 1995. Oceanic basement of the Caribbean basins. Geological Society of America, Abstracts with Programs 27, 153. Hards, V. L., Kempton, P. D. & Thompson, R. N., 1995. The heterogeneous Iceland plume: new insights from the alkalic basalts of the Snaefell volcanic centre. Journal of the Geological Society, London 152, 10031011. Hay, R., 1977. Tectonic evolution of the CocosNazca spreading center. Geological Society of America Bulletin 88, 14041420.

700

KERR et al.

CRETACEOUS COLOMBIAN BASALTS

Hemond, C., Arndt, N. T., Lichtenstein, U., Hofmann, A. W., Os karsson, N. & Steinthorsson, S., 1993. The heterogeneous Iceland plume; NdSrO isotopes and trace element constraints. Journal of Geophysical Research 98, 1583315850. Henderson, W. G., 1979. Cretaceous to Eocene volcanic arc activity in the Andes of northern Ecuador. Journal of the Geological Society, London 136, 367378. Hill, R. I., 1993. Mantle plumes and continental tectonics. Lithos 30, 193206. Humphris, S. E. & Thompson, G., 1978. Trace element mobility during hydrothermal alteration of oceanic basalts. Geochimica et Cosmochimica Acta 42, 127136. Hurfurd, A. J. & Hammerschmidt, K., 1985. 40Ar/39Ar and KAr dating of the Bishop and Fish Canyon Tus: calibration ages for ssion-track dating standards. Chemical Geology 58, 2332. Iturralde-Vinent, M. A., 1994. Cuban geology: a new plate tectonic synthesis. Journal of Petroleum Geology 17, 3970. Kerr, A. C., 1995. The geochemistry of the MullMorvern Tertiary lava succession, NW Scotland: an assessment of mantle sources during plume-related volcanism. Chemical Geology 122, 4358. Kerr, A. C., Saunders, A. D., Tarney, J., Berry, N. H. & Hards, V. L., 1995. Depleted mantle plume geochemical signatures: no paradox for plume theories. Geology 23, 843846. Kerr, A. C., Marriner, G. F., Arndt, N. T., Tarney, J., Nivia, A., Saunders, A. D. & Duncan, R. A., 1996a. The petrogenesis of komatiites, picrites and basalts from the Isle of Gorgona, Colombia; new eld, petrographic and geochemical constraints. Lithos 37, 245260. Kerr, A. C., Tarney, J., Marriner, G. F., Klaver, G. Th., Saunders, A. D. & Thirlwall, M. F., 1996b. The geochemistry and petrogenesis of the late-Cretaceous picrites and basalts of Curacao, Netherlands Antilles: a remnant of an oceanic plateau. Contributions to Mineralogy and Petrology 124, 2943. Kerr, A. C., Tarney, J., Marriner, G. F., Nivia, A., Saunders, A. D. & Klaver, G. Th., 1996c. The geochemistry and tectonic setting of late Cretaceous Caribbean and Colombian volcanism. Journal of South American Earth Sciences 9, 111120. Kerr, A. C., Tarney, J., Marriner, G. F., Nivia, A. & Saunders, A. D., 1997. The CaribbeanColombian Cretaceous igneous province: the internal anatomy of an oceanic plateau. Geophysical Monograph, American Geophysical Union. In press. Kimura, G. & Ludden, J., 1995. Peeling oceanic crust in subduction zones. Geology 23, 217220. Lapierre, H., Dupuis, V., de Lepinay, B. M., Ruiz, J., Tardy, M., Maury, R. C., Hernandez, J. & Loubet, M., 1997. Is the lower Duarte igneous complex (Northern Hispaniola) a remnant of the Caribbean plume-generated oceanic plateau? Journal of Geology 105, 111120. Lassiter, J. C., DePaolo, D. J. & Mahoney, J. J., 1995. Geochemistry of the Wrangellia ood basalt province: implications for the role of continental and oceanic lithosphere in ood basalt genesis. Journal of Petrology 36, 9831009. Lebras, M., Megard, F., Dupuy, C. & Dostal, J., 1987. Geochemistry and tectonic setting of pre-collision Cretaceous and Paleogene volcanic rocks of Ecuador. Geological Society of America Bulletin 99, 569578. Lebron, M. C. & Pert, M. R., 1994. Petrochemistry and tectonic signicance of Cretaceous island arc rocks, Cordillera Central, Dominican Republic. Tectonophysics 229, 69100. Macia, C., 1985. Caractersticas petrogracas y geoqumicas de rocas basalticas de la pennsula de Cabo Corrientes (Serrana de Baudo), Colombia. Geologia Colombiana 14, 2537. Mahoney, J. J., 1987. An isotopic survey of Pacic oceanic plateaux: implications for their nature and origin. Geophysical Monograph, American Geophysical Union 43, 207220.

Mahoney, J. J. & Spencer, K. J., 1991. Isotopic evidence for the origin of the Manihiki and Ontong Java oceanic plateaux. Earth and Planetary Science Letters 104, 196210. Mahoney, J. J., Storey, M., Duncan, R. A., Spencer, K. J. & Pringle, M., 1993. Geochemistry and age of the Ontong Java Plateau. Geophysical Monograph, American Geophysical Union 77, 233261. Mahoney, J. J., Sinton, J. M., Macdougall, J. D., Spencer, K. J. & Lugmair, G. W. 1994. Isotope and trace element characteristics of a super-fast spreading ridge: East Pacic Rise, 1323S. Earth and Planetary Science Letters 121, 173193. Marriner, G. F. & Millward, D., 1984. The petrology and geochemistry of Cretaceous to Recent volcanism in Colombia: the magmatic history of an accretionary plate margin. Journal of the Geological Society, London 141, 473486. McBirney, A. R., 1963. Factors governing the nature of submarine volcanism. Bulletin of Volcanology 37, 455469. McCourt, W. J. & Feininger, T., 1984. High pressure metamorphic rocks in the Central Cordillera of Colombia. British Geological Survey Report Series 16/1, 2835. McCourt, W. J., Aspden, J. A. & Brook, M., 1984. New geological and geochronological data from the Colombian Andes: continental growth by multiple accretion. Journal of the Geological Society, London 141, 831845. McGeary, S. & Ben-Avraham, Z., 1986. The accretion of Gorgona Island, Colombia: multichannel seismic evidence. In: Howell, D. G. (ed.) Tectonostratigraphic Terranes of the Circum-Pacic Region. Houston, TX: Circum-Pacic Council for Energy and Mineral Resources, pp. 543554. McKenzie, D. P. & ONions, R. K., 1991. Partial melt distributions from inversion of rare earth element concentrations. Journal of Petrology 32, 10211091. Millward, D., Marriner, G. F. & Saunders, A. D., 1984. Cretaceous tholeiitic volcanic rocks from the Western Cordillera of Colombia. Journal of the Geological Society, London 141, 847860. Mooney, W. D., 1980. An East PacicCaribbean ridge during the Jurassic and Cretaceous and the evolution of western Colombia. In: R. H. Pigler (ed.) The Origin of the Gulf of Mexico and Early Tertiary Opening of the Central North Atlantic Ocean. Houston Geological SocietyContinuing Education Series. Baton Rouge, LA: School of Geoscience, Louisiana State University, pp. 5574. Nielsen, R. L., 1988. A model for the simulation of combined major and trace element liquid lines of descent. Geochimica et Cosmochimica Acta 52, 2738. Nivia, A., 1987. The geochemistry and origin of the Amaime and Volcanic sequences, SW Colombia. M.Phil. Thesis, University of Leicester, UK, 164 pp. Nivia, A., 1996. The Bolvar macultramac complex, SW Colombia: the base of an oceanic plateau. Journal of South American Earth Sciences 9, 5968. Nivia, A., Galvis, N. & Maya, M., 1996. Mapa Geologico de Co lombiaEscala 1:100 000, Plancha 242Zarzal. Bogota: IN GEOMINAS. Pearce, J. A. & Pearson, G., 1996. Leg 165. United Kingdom Ocean Drilling Project Newsletter 22, 1011. Pichler, H., Stibane, F. R. & Weyl, R., 1974. Magmatismus and Krustenbau in sudlichen Mittelamerika, Kolumbien and Ecuador. Neues Jahrbuch fur Geologie und Palaontologie, Monatshefte 2, 102136. Pindell, J. L. & Barrett, S. F., 1990. Geologic evolution of the Caribbean region: a plate tectonic perspective. In: Dengo, G. & Case, J. E. (eds) The Geology of North America HThe Caribbean Region. Boulder, CO: Geological Society of America, pp. 405432. Saunders, A. D., 1985. Geochemistry of basalts from the Nauru Basin, Deep Sea Drilling Project Legs 61 and 89: implications for the origin

701

JOURNAL OF PETROLOGY

VOLUME 38

NUMBER 6

JUNE 1997

of oceanic ood basalts. In: Initial Reports of the Deep Sea Drilling Project, 89. Washington, DC: US Government Printing Oce, pp. 499517. Saunders, A. D., Storey, M., Kent, R. W. & Norry, M. J., 1992. Consequences of plumelithosphere interactions. In: Storey, B. C., Alabaster, T. & Pankhurst, R. J. (eds) Magmatism and the Causes of Continental Breakup. Geological Society of London Special Publication 68, 4160. Saunders, A. D., Tarney, J., Kerr, A. C. & Kent, R. W., 1996. The formation and fate of large igneous provinces. Lithos 37, 8195. Sen, G., Hickey-Vargas, R., Waggoner, D. G. & Maurrasse, F., 1988. Geochemistry of basalts from the Dumisseau Formation, southern Haiti: implications for the origin of the Caribbean Sea crust. Earth and Planetary Science Letters 87, 423437. Sinton, C.W., 1996. A tale of two large igneous provinces: geochronological and geochemical studies of the North Atlantic volcanic province and the Caribbean oceanic plateau. Ph.D. Thesis, Oregon State University, Corvallis. Spadea, P., Delaloye, M., Espinosa, A., Orrego, A. & Wagner, J. J., 1987. Ophiolite complex from La Tetillia, SW Colombia, South America. Journal of Geology 95, 377395. Spadea, P., Espinosa, A. & Orrego, A., 1989. High-Mg extrusive rocks from the Romeral zone ophiolites in the southwestern Colombian Andes. Chemical Geology 77, 303321. Storey, M., Mahoney, J. J., Kroenke, L. W. & Saunders, A. D., 1991. Are oceanic plateaux sites of komatiite formation? Geology 19, 376379. Storey, M., Kent, R., Saunders, A. D., Salters, V. J., Hergt, J., Whitechurch, H., Sevigny, J. H., Thirlwall, M. F., Leat, P., Ghose, N. C. & Giord, M., 1992. Lower Cretaceous volcanic rocks on continental margins and their relationship to the Kerguelen Plateau. In: Proceedings of the Ocean Drilling Program, Scientic Results, 120. College Station, TX: Ocean Drilling Program, pp. 3353. Sun, S.-s. & McDonough, W. F., 1989. Chemical and isotope systematics of oceanic basalts: implications for mantle composition and processes. Geological Society of London, Special Publication 42, 313345.

Tarney, J. & Marsh, N. G., 1991. Major and trace element geochemistry of Holes CY-1 and CY-4: implications for petrogenetic models. In: Gibson, I. L., Malpas, J., Robinson, P. T. & Xenophontos, C. (eds) Initial Reports of the Cyprus Crustal Study. Geological Survey of Canada, Paper 90-20, 133176. Tejada, M. L. G., Mahoney, J. J., Duncan, R. A. & Hawkins, M. P., 1996. Age and geochemistry of basement and alkalic rocks of Maliata and Santa Isabel, Solomon Islands, southern margin of Ontong Java Plateau. Journal of Petrology 37, 361394. Thirlwall, M. F., 1991a. High precision multicollector isotope analysis of low levels of Nd as oxide. Chemical Geology (Isotope Geoscience Section) 94, 1324. Thirlwall, M. F., 1991b. Long-term reproducibility of multicollector Sr and Nd isotope ratios analysis. Chemical Geology (Isotope Geoscience Section) 94, 85104. Thirlwall, M. F., Upton, B. G. J. & Jenkins, C., 1994. Interaction between continental lithosphere and the Iceland plumeSrNdPb isotope geochemistry of Tertiary basalts, NE Greenland. Journal of Petrology 35, 839879. Tistl, M. & Salazar, G., 1994. The tectono-magmatic evolution of northwestern South America. Zentralblatt fur Geologie und Palaontologie 1993(12), 439453. Vallier, T. L., Windom, K. E., Seifert, K. E. & Thiede, J., 1980. Volcanic rocks cored on Hess Rise, western Pacic Ocean. Nature 286, 4850. Van Thournout, F., Hertogen, J. & Quevedo, L., 1992. Allochthonous terranes in northwestern Ecuador. Tectonophysics 205, 205221. Wallrabe-Adams, H.-J., 1990. Petrology and geotectonic development of the Western Ecuadorian Andes: the Basic Igneous Complex. Tectonophysics 185, 163182. Walsh, J. N., Buckley, F. & Barker, J., 1981. The simultaneous determination of rare earth elements in rocks using inductively coupled plasma mass spectrometry. Chemical Geology 33, 141153. White, W. M., McBirney, A. R. & Duncan, R. A., 1993. Petrology and geochemistry of the Galapagos Islands: portrait of a pathological mantle plume. Journal of Geophysical Research 98, 1953319563.

702

Das könnte Ihnen auch gefallen