Вы находитесь на странице: 1из 26

Sedimentology (2003) 50, 10791104

doi: 10.1046/j.1365-3091.2003.00595.x

Provenance analysis and tectonic setting of the Ordovician clastic deposits in the southern Puna Basin, NW Argentina
U DO ZIMMERMANN* and HEINRICH BAHLBURG *Department of Geology, RAU University, PO Box 524, 2006 Auckland Park, Johannesburg, South Africa (E-mail: uz@na.rau.ac.za) Geologisch-Plontologisches Institut, Westflische Wilhelms-Universitt Munster, aa a a Corrensstrasse 24, 48149 Munster, Germany (E-mail: bahlbur@uni-muenster.de)
ABSTRACT

Provenance studies on Early to Middle Ordovician clastic formations of the southern Puna basin in north-western Argentina indicate that the sedimentary detritus is generally composed of reworked crustal material. Tremadoc quartzrich turbidites (Tolar Chico Formation, mean composition Qt89 F7 L4) are followed by volcaniclastic rocks and greywackes (Tolillar Formation, mean Qt33 F42 L25). These are in turn overlain by volcaniclastic deposits (mean Qt24 F30 L46) of the Diablo Formation (late Arenigearly Llanvirn) that are intercalated by yic Orogeny during the Middle lava ows. All units were deformed in the Oclo and Late Ordovician. Sandstones of the Tolar Chico Formation are characterized by Th/Sc ratios > 1, La/Sc ratios  10, whereas associated ne-grained wackes show slightly lower values for both ratios. LREE (light rare earth elements) enrichment of the arenites is  50 chondrite, Eu/Eu* values are between 072 and 092, and at HREE (heavy rare earth elements) patterns indicate a derivation from mostly felsic rocks of typical upper crustal composition. The eNd(t sed) values scatter around )11 to )9. The calculated Nd-TDM residence ages vary between 18 and 20 Ga indicating contribution by a Palaeoproterozoic crustal component. The Th/Sc and La/Sc ratios of the Tolillar Formation are lower than those of the Tolar Chico Formation. Normalized REE (rare earth elements) patterns display a similar shape to PAAS (post-Archaean average Australian shale) but with higher abundances of HREEs. Eu/Eu* values range between 044 and 117, where the higher values reect the abundance of plagioclase and feldspar-bearing volcanic lithoclasts. Average eNd(t sed) values are less negative at )51, and Nd-TDM are lower at 16 Ga. This is consistent with characteristics of regional rocks of upper continental crust composition, which most probably represent the sources of the studied detritus. The rocks of the Diablo Formation have the lowest Th/Sc and La/Sc ratios, lower LREE abundances than the average continental crust and are slightly enriched in HREEs. Eu/Eu* values are between 063 and 117. The Nd isotopes (eNd(t sed) )3 to )1; TDM 12 Ga) indicate that one source component was less fractionated than both the underlying Early Ordovician and the overlying Middle Ordovician units. Synsedimentary vulcanites in the Diablo Formation show the same isotopic composition. Our data indicate that the sedimentary detritus is generally composed of reworked crustal material, but that the Diablo Formation appears to contain  80% of a less fractionated component, derived from a contemporaneous continental volcanic arc. There are no data indicating an exotic detrital source or the accretion of an exotic block at this part of the Gondwana margin during the Ordovician. Keywords Geochemistry, isotope geochemistry, Ordovician, provenance analysis, Puna, retroarc basin deposits.
2003 International Association of Sedimentologists

1079

1080

U. Zimmermann & H. Bahlburg

Fig. 1. (A) Geological map of the Puna with important outcrop regions: 1, Salar de Pocitos, including Complejo Igneo Pocitos and Complejo Basico Ojo de Colorados; 2, Central Sierra de Calalaste (Quebrada El Diablo); 3, Southern Sierra de Calalaste (SC); 4, Huaitiquina and Aguada de la Perdz; 5, Salar de Rincon; 6, Salar de Antofalla; 7, Cuchiyacu pluton, modied after Rapela et al. (1992). (B and C) Study areas in the southern Puna of north-western Argentina. Outcrops: 1, Vega Quiron; 2, Quebrada Honda; 3, Quebrada Carro Grande; 4, Falda Cienaga; 5, Los Nacimientos; 6, Antofagasta de la Sierra; 7, Quebrada del Diablo.

INTRODUCTION Provenance studies of sedimentary rocks aim to decipher the composition and geological evolution of the sediment source areas and to constrain the tectonic setting of the depositional basin. Here, results are presented of a provenance study of the Lower to Middle Ordovician clastic Tolar

Chico, Tolillar and Diablo Formations of the coeval southern Puna Basin of north-western Argentina. A meaningful provenance analysis necessarily concentrates on the evaluation of indicators that are inherited from the original source areas. However, the data need to be scrutinized for the effects of secondary factors that have the potential

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Provenance of sedimentary rocks, Puna Basin to obscure this information, i.e. sorting, weathering, diagenesis, metamorphism and recent alteration processes (Morton & Hallsworth, 1999). Accordingly, a combination of several independent analytical approaches is required to have condence in the provenance indicators. In view of the Early Palaeozoic age of the studied rocks and the effects of two orogenic overprints, i.e. during the end-Ordovician Ocloyic and the Tertiary Andean orogenies, a combination of eldwork, petrographical, geochemical and Nd isotope data is used to relate the composition and evolution of the source regions to the plate tectonic setting of the poorly understood southern Puna Basin, and to evaluate the potential role of exotic terranes in the development of this part of the Western Gondwana margin during the Early Palaeozoic.

1081

GEOLOGICAL SETTING The Argentinean Puna forms the southern continuation of the Bolivian highland plateau, situated at an altitude of  4000 m. Sedimentary and magmatic rocks of Ordovician age are widely exposed in a system of horsts and grabens formed since the middle Tertiary as a consequence of the Andean orogeny (Zeil, 1979; Allmendinger et al., 1997). This study concentrates on two representative outcrop areas of Lower Palaeozoic rocks in the southern Puna, namely at Salar de Pocitos and the middle to southern part of the Sierra de Calalaste (Fig. 1). The association of Ordovician sedimentary units with mac to ultramac rocks in the southern Puna has been interpreted as the remnants of an Ordovician ocean (e.g. Ramos et al., 1986) that closed during the collision of one of several proposed westerly terranes with the western Gondwana margin (Fig. 2). These terranes include the Puna Terrane (e.g. Coira et al., 1982), the composite PunaFamatina Terrane (Conti et al., 1996) and the ArequipaAntofalla Terrane (e.g. Ramos et al., 1986; Forsythe et al., 1993; Bahlburg & Herve, 1997), partly interpreted as derived from Laurentia (Dalla Salda et al., 1992; Dalziel, 1997). Other authors assume that the continental margin of Gondwana in this region alternated between compressive and extensional tectonic regimes (e.g. Mon & Hongn, 1991). These different tectonic scenarios imply very different palaeotectonic positions for the Puna Basin (oceanic, back-arc, retroarc or intracratonic basin). In this provenance study of

Fig. 2. Proposed geotectonic units between 15S and 40S including the assumed Puna and PunaFamatina Terranes. The black spots indicate the alleged ophiolitic rocks in the southern Puna (e.g. Ramos et al., 1986). The Famatina Terrane represents the southern part of a hypothetical PunaFamatina Terrane (modi ed after Bahlburg & Herve, 1997); STG, Santiago de Chile; BUE, Buenos Aires.

Ordovician clastic formations, new petrographic, geochemical and isotope geochemical analyses are presented for the southern Puna Basin. These data are the key to the interpretation of the source area evolution and establishing the basins palaeotectonic setting.

STRATIGRAPHY The Lower Ordovician clastic sedimentary successions in the southern Puna include synsedimentary lava ows. The formations are furthermore tectonically associated with mac to ultramac rocks of poorly dened pre-Arenig age; the youngest age limit is given by felsic to intermediate intrusives of the Complejo Igneo Pocitos (Fig. 1A, no. 1) dated at 476 2 Ma (early Arenig; Kleine et al., 1999). Ordovician deposits to the east of Salar de Pocitos and south of the Salar de Hombre Muerto (Fig. 1B and C: 1, Vega n; Quiro 2, Quebrada Honda; 3, Quebrada Carro Grande; 4, Falda Cienaga; 5, Los Nacimientos; 6, Antofagasta de la Sierra) preserve no record of contemporaneous magmatism. The plutonic rocks of the Complejo Igneo Pocitos intruded pre-Arenig mac to ultramac

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

1082

U. Zimmermann & H. Bahlburg medium-grained quartz arenites, which give way to volcaniclastic sandstones, quartz-rich siltstones and shales. Fine- to coarse-grained volcaniclastic greywackes and massive feldspathic greywackes are the most abundant lithologies. Flute marks, ripple, cross- and parallel lamination are consistent with deposition from turbidity currents. Debris-ow deposits consisting of immature detritus are interbedded with the turbidites and shale horizons. The immature material points to only minor reworking between the source and the site of deposition. In the central part of the Sierra de Calalaste (Quebrada Diablo, Fig. 1A, no. 2; Fig. 1C, no. 7), a volcanogenic unit of Arenig to Lower Llanvirn age crops out, which is informally named the Diablo Formation after its typical outcrop in the Quebrada El Diablo (260335,0S, 673639,3W; Fig. 3; Zimmermann, 2000). This formation is dominated by very coarse-grained debris-ow deposits and ne- to coarse-grained greywackes and siltstones representing turbidites. Rhyolitic to andesitic lavas, including some pillowed

bodies and show tectonic contacts with the sedimentary rocks. Tectonic juxtaposition of the mac to ultramac rocks most likely occurred during the Ocloyic Orogeny in the Late Ordovician. Geochemical data characterize the mac bodies as calc-alkaline rocks related to a magmatic arc environment (Zimmermann et al., 1999). An improved stratigraphy (Fig. 3) for the Lower Ordovician units in the southern Puna is based on new nds of graptolites (Zimmermann, 2000). The oldest Ordovician sedimentary rocks are represented by the Tolar Chico Formation (after Zappettini et al., 1994; redened here). This unfossiliferous formation contains thick quartz arenites as well as thin layers of silty wackes displaying grading and ripple cross-lamination. The presence of thick turbidite packages and the absence of shallow-water deposits indicate a marine environment below wave base. The Tolar Chico Formation is conformably overlain by the Tolillar Formation (after Zappettini et al., 1994; redened here). At the base, it consists of rare

Fig. 3. Stratigraphic table of the Early Palaeozoic of NW Argentina. The grey shading marks the stratigraphy dis cussed in this paper (compilation of stratigraphic references in Acenolaza & Baldis, 1987; Bahlburg & Herve, 1997; Zimmermann et al., 2002).
2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Provenance of sedimentary rocks, Puna Basin andesites, form occasional intercalations. Chilled margins mark the contacts of the lava ows with the underlying sedimentary rocks. Rare peperites and volcaniclastic ows are the result of subaqueous synsedimentary volcanism. This association of volcanogenic sedimentary rocks and interbedded lava ows could be interpreted as having been deposited in a volcanic apron (e.g. White & Busby-Spera, 1987). Equivalent associations are also exposed in the northern Puna (Bahlburg, 1991). These units of volcanogenic facies contrast markedly with the overlying siliciclastic Coquena and Falda Cienaga Formations exposed to the north-east and east (outcrops 16 in Figs 1B and C and 3; Acenolaza et al., 1975; Zimmermann et al., 2002). Flute marks (n 104) in the Lower Ordovician formations indicate a uniform axial sediment transport from the S and SSW to the N and NNE, which is also dominant in the continuation of the Ordovician basin in the northern Puna. Here, volcaniclastic detritus originated in a western volcanic arc source, whereas more mature basement detritus was derived from the eastwardlying Late ProterozoicEarly Cambrian Puncoviscana orogen (Bahlburg, 1991, 1998). The entire Lower Ordovician sedimentary succession from the Tolar Chico Formation to the Diablo Formation underwent only a very lowgrade metamorphism, as indicated by analyses of the illite and chlorite crystallinities (Zimmermann, 1999). The sedimentary successions are isoclinally folded with folds commonly verging westwards. Towards the south, the deformation clearly increases, and the vergence becomes more variable. The cleavage is predominantly parallel to the bedding in the coarse units. In the pelitic rocks, two cleavages may be developed locally. PROVENANCE STUDY

1083

Analytical methods
Framework mineral composition was quantied using the point-counting method of Gazzi and Dickinson as described by Ingersoll et al. (1984) with a Swift model F counter. The thin sections were partly stained to identify the different feldspars. A total of 350530 counts were executed in traverses to obtain a sufcient database. Heavy mineral composition was quantied using the counting method of Boenick (1983). The separated heavy minerals (60150 lm) were divided in two fractions (6090 lm and 90 120 lm) and counted in traverses up to 250 counts. The opaque minerals were not considered in the calculation of the mean percentages. X-ray diffraction (XRD) was carried out on 43 samples (< 2 lm) to determine the illite crystallinity. The samples were dried on a sample holder (2 2 cm) and measured in four directions glycolized and not glycolized. A Philips X-ray diffractometer was operated at 30 mA and 40 kV. The sample were measured from 2 to 45 2h, in steps of 002 per 2 s. The measurements of illite crystallinity were analysed using diffrak (Siemens, version 30). The analytical process is described in Warr & Rice (1994), and the standards of Warr & Rice (1994) were used. Scanning electron microscope (SEM) studies were carried out on light and heavy minerals. The separated minerals were mounted with a current conducting glue on sample holders (diameter 12 cm) and coated with gold. The samples were analysed using a SEM 505 (Philips). Cathodoluminescence (CL) studies were performed using an ASK-SEM-CL (ASK-Wesel) at the Department of Mineralogy (University of Heidelberg, Germany). The CL analyses were compared with backscatter electron images of the same samples using a BSE microscope LEO 440 (Firma LEO-Zeiss/Leica-Oberkochem, Cambridge). X-ray uorescence (XRF) analyses for major and trace elements were carried out using a SRS 303 (Siemens) wavelength-dispersive XRF spectrometer operating at 50 kV and 50 mA. Only fresh and homogeneous samples were selected, crushed and milled. Fudion disks were prepared with Spektroux 100 (Lithiumborate, Li2B4O7) as ux. Detection limits for major elements are related to the atomic number, and are between 1 and 10 lg for medium heavy elements. Precision is 05 (1 r); accuracy was controlled by repetitive measurements of standards, and each sample

PREVIOUS STUDIES Geological studies of the Ordovician in the remote southern Puna are rare. Only local studies on the metamorphic basement, early Palaeozoic magmatic rocks and Ordovician sedimentary rocks of the southern Puna have been published, including preliminary palaeontological work (see compilations in Ramos et al., 1986; Rapela et al., 1992; Bahlburg & Herve, 1997; Zimmermann, 2000; Zimmermann et al., 2002). However, Ordovician sedimentary rocks are extensively exposed in this region and are the key to interpreting the active margin evolution of the southern central Andes during the Early Palaeozoic.

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

1084

U. Zimmermann & H. Bahlburg sedimentary lithoclasts (e.g. sample A84, Table 1). Cathodoluminescence analysis of quartz grains indicates a dominance of metamorphic quartz (mean: Qmet 75%, Qplut 20%, Qvolc 5%; Zimmermann, 1999). Towards the top of the formation, volcanic quartz becomes more abundant. SEM pictures of individual quartz grains show v-shaped indentations/impact marks, which suggest a subaqueous turbulent transport regime (Krinsley & Marshall, 1987). Some feldspars show alteration to sericite and others, as observed with SEM, are fresh. Lithoclasts are subordinate and, in most cases, they are of phyllitic or other (meta-)sedimentary origin (Table 1). The heavy mineral composition of the Tolar Chico Formation is dominated by zircon, tourmaline and rutile in the 60150 lm fraction. Metamorphic minerals such as sillimanite, epidote, zoisite and pumpellyite are subordinated. Amphiboles and other mac minerals are extremely rare (< 2%; Table 2).

was measured twice. The errors of major elements vary between 05% and 2%. INAA (instrumental neutron activation analysis) was performed by Actlabs (Ontario, Canada). Samples were crushed and milled in an agate mill. The powder was dissolved in lithium metaborate ux fusion, and the resulting molten bead was rapidly digested in a weak nitric acid solution. Detection limits are between 05 p.p.m. and 5 p.p.m. for elements such as Ag and Cr. INAA precision and accuracy based on replicate analysis of international rock standards are 25% (1 r) for most elements and 10% for U, Sr, Nd and Ni. TIMS (thermal ionization mass spectrometry) for Sm-Nd analyses were performed at the NIGL (BGS, Keyworth, UK) by isotope dilution with a mixed 149 Sm150Nd spike, using HDEHP-coated polystyrene columns for the REE separation. Measurements were carried out on a fully automated VG 354 mass spectrometer, using static collection to determine the Sm and Nd concentrations, and mixed static/peak-jumping to determine the 143 Nd/144Nd ratios with an internal precision of 10 p.p.m. (1 SD). Long-term reproducibility of 143 Nd/144Nd ratios both on the La Jolla and in-house standards is better than 15 p.p.m. (1 r). Sm/Nd ratios on rock standards are reproducible to 0102% (1 r). La Jolla standard measured 0511900/)0000008 of seven analyses. Values used for calculations: Sm/NdCHUR 01966, 147 147 Sm/144Ndcrust 0115, Sm/144Ndmantle 0222, eDMt 0 86, Q 2513, f 0415, threestage model Sm/Ndlimit 0300, equiv. TDM 2063,1993896236. Nd model ages calculated after DePaolo et al. (1991).

Petrofacies Tolar Chico Formation The Tolar Chico Formation (Figs 3 and 4) is composed mainly of quartz arenites intercalated with some ne-grained wackes. Besides quartz, the quartz arenites contain slightly more alkali feldspar (microcline) than plagioclase (Fig. 4D). Rock fragments are usually rare but, in some samples, sedimentary lithoclasts, with internal silt to ne-sand grain size, are abundant (Fig. 4C, Table 1). Accessory phases include biotite, muscovite and calcite. XRD analyses show the presence of chlorite and illite. The pseudomatrix (Dickinson, 1970) has the same qualitative mineralogical composition as the framework fraction. The wackes are similar in their matrix composition, but show a higher amount of

Tolillar Formation The sandstones and wackes of the Tolillar Formation are relatively rich in feldspar including both plagioclase and potassium feldspar (Fig. 4A and D). The abundance of framework clast types is variable in the greywackes and volcaniclastic sandstones (Fig. 4; Table 1). Nearly 70% of the quartz grains are of volcanic origin according to CL analysis (mean: Qmet 10%, Qplut 20%, Qvolc 70%; Zimmermann, 1999). Quartz grains have ragged form and display resorption embayments probably indicating a rapid cooling process in volcanic rocks (Matter & Ramseyer, 1985). Also, some resorption embayments may be the result of solution processes (e.g. Schneider, 1993). In the upper part of the succession, massive feldspar-rich greywackes occur (plg + kf > 50%; A223, A273, A296, A309, X27; Table 1). The amount of plagioclase in relation to potassium feldspar increases slightly from the Tolar Chico Formation (Fig. 4D). Microcline is absent. Volcanic lithoclasts are abundant. They are most commonly of felsic origin (Lv(felsic)/ Lv(intermed) 3:1). Metamorphic lithoclasts are infrequent (Table 1). The arrow in Fig. 4A and B shows the stratigraphic trend in the Tolillar Formation to an increasingly volcanic composition, overlapping with the Diablo Formation. In the heavy mineral population, tourmaline and rutile are less abundant, and zircon is only slightly more frequent than in the Tolar Chico Formation. Apatite and amphibole abundance increases markedly from the Tolar Chico Forma-

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Provenance of sedimentary rocks, Puna Basin

1085

Fig. 4. Framework mineral composition of Lower Ordovician sedimentary rocks in the southern Puna. (AD) Framework mode diagrams according to Dickinson & Suczek (1979) and Dickinson et al. (1983): Q, quartz; F, feldspar; L, lithoclasts; Qm, monocrystalline quartz; Qp, polycrystalline quartz; Qt, total quartz Qm + Qp; P, plagioclase; K, potassium feldspar; Lt, total lithoclasts including Qp; Lv, volcanic lithoclasts; Ls, sedimentary lithoclasts. (A) Qt-F-L diagram: the arrow indicates the trend from bottom to top in the Tolillar Formation. (B) Qm-F-Lt diagram: the arrow indicates the trend from bottom to top in the Tolillar Formation. (C) Qp-Lv-Ls diagram: note the mixed composition of the Tolillar Formation and the relatively high amount of Ls in the Diablo Formation. (D) Qm-P-K: note that albitization may have reduced the P/F ratio. For comparison, the averages of Lower to Middle Ordovician sedimentary rocks and turbiditic sandstones of the northern Puna: VS, volcanosedimentary successions; LTS, lower turbidite system; UTS, upper turbidite system (Fig. 3); Qm-P-K data for the northern Puna rocks and Coquena Formation are not available. Falda Cienaga Formation from Zimmerman et al. (2002); northern Puna rocks from Bahlburg (1990) were counted after Suttner & Basu (1985) and corrected for comparison with the GazziDickinson method described by Ingersoll et al. (1984). UTS, LTS, VS and Coquena Formation averages are not available for (D) and Coquena Formation for (C).

tion. Metamorphic minerals, apart from epidote, are very rare (Table 2).

Diablo Formation This formation is also relatively feldspathic with slightly more plagioclase than potassium feldspar

(Fig. 4, Table 1). Plagioclase abundances are roughly similar to those in the Tolillar Formation. Plagioclase occurs as large grains (> 1 cm), and many preserve a magmatic zonation. In the feldspar-rich greywackes of this formation, quartz abundances vary between 18% and 46% (Fig. 4;

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

1086

U. Zimmermann & H. Bahlburg

Table 1. Selected framework mineral data of the Ordovician Puna Basin deposits [data this study, Bahlburg (1998) for the southern Puna Basin and Zimmermann et al. (2002) for the Falda Cienaga Formation]. Basin Ordovician southern Puna Basin Tolar Chico Formation (n 23) Tolillar Formation (n 26) Diablo Formation (n 14) Falda Cienaga Formation Ordovician northern Puna Basin Volcanosedimentary successions Lower turbidite system Upper turbidite system Qt 888 33 245 78 49 49 67 F 67 42 30 12 17 23 8 L 45 251 457 9 34 28 25 Matrix (%) 854 214 1688 85 Lv/L 006 03 03 0 098 085 062 P/F 036 04 06 023 081 064 072

Complete point-counting data for the Tolar Chico, Tolillar and Diablo Formations are available in Supplementary material, Table S1.

Table 1). Quartz grains often show resorption embayments pointing to derivation from volcanic sources. CL analysis shows that more than 80% of the quartz grains are originally from volcanic rocks. Quartz derived from metamorphic rocks is nearly absent (mean: Qmet 2%, Qplut 13%, Qvolc 85%; Zimmermann, 1999). The amount of volcanic lithoclasts increases, including a notable population of intermediate composition. Even though there is some overlap with some volcanogenic samples of the Tolillar Formation, the Diablo Formation contains the strongest input of volcanogenic material of all three successions (Fig. 4; Table 1). Relicts of altered glass shards, peperites and massive ne-grained silicied layers interpreted as recrystallized ashes are abundant in the Diablo Formation. XRD analysis reveals that the siltstones and shales have a similar qualitative mineralogical composition. In the heavy mineral fraction (Table 2), elongated idiomorphic zircons with gas bubble inclusions point to a volcanic origin (Kostov, 1973). The amount of amphibole increases markedly over the Tolillar Formation (Table 2). Neither metamorphic minerals nor heavy minerals indicating a mac source (e.g. Morton et al., 1992; Schfer & Dorr, 1997) were a observed.

Major element geochemistry


The geochemical analysis of sedimentary rocks is a valuable tool for provenance studies of matrixrich sandstones as long as the bulk composition is not strongly affected by diagenesis, metamorphism or other alteration processes (McLennan et al., 1993). Abundances and ratios of major elements thus need to be checked for mobility, especially during diagenesis (e.g. Boles & Franks,

1979; McLennan et al., 1980; McLennan, 2001). A good measure of the degree of chemical weathering can be obtained by calculation of the chemical index of alteration (CIA; Nesbitt & Young, 1982) using the molar proportions of Al2O3, CaO*, Na2O and K2O, where CaO* is CaO in silicates only. The resultant value (CIA) is a measure of the proportion of Al2O3 vs. the mobile oxides in the analysed samples typically representing the alteration of feldspars and volcanic glass to clay minerals. An ideal weathering trend for a rock of rhyolitic to dacitic composition is indicated by a solid arrow, subparallel to the line linking the CaO* + Na2O and Al2O3 apices of the triangular diagram in Fig. 5 (A-CN-K diagram after Nesbitt & Young, 1984; Fedo et al., 1995). However, whereas some samples follow this ideal trend, others deviate from it along a weathering trend towards an illite composition, possibly as a result of a metasomatic increase in K during diagenesis, caused by either the conversion of aluminous clay minerals to illite or transformation of plagioclase to K-feldspar (Fedo et al., 1995). Alternatively, this pattern may indicate mixing of a moderately weathered source with an unweathered one of different primary composition, or a secondary gain or loss especially of Na and K, and potentially Ca, in the silicate fraction, e.g. during albitization (McLennan et al., 1993). The average CIA value of all formations is  58 (SD 68), with broadly similar formation averages between 57 and 63 (Table 3). Variations resulting from different grain sizes are minor (Fig. 5). The quartz arenites of the Tolar Chico Formation have relatively low CIA values because of low Al2O3 concentrations, and do not seem to follow a clear weathering trend. Nesbitt et al. (1996) demonstrated that a probable in situ

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Provenance of sedimentary rocks, Puna Basin


Table 2. Abundances of the heavy mineral population in the Lower Ordovician sedimentary rocks of the southern Puna (the opaque fraction was dominated by pyrite and haematite). 1246 1792 1567 *After Boenick (1983): opaque minerals are not considered in the calculation of mean percentages. After Hubert (1962): percentage of zircon, tourmaline and rutile to total heavy minerals. Note the constant percentage of zircon in all three formations, and the change in the amount of idiomorphic grains. The number of apatite and amphibole grains increases from the Tolar Chico to the Diablo Formation, metamorphic heavy minerals were not found in the Diablo Formation. Thirty-eight samples with grain sizes between 60 and 250 lm were separated, and the fraction between 60 and 150 lm was analysed. Separation techniques and counting method according to Boenick (1983) and Mange & Maurer (1992). Zircon Zircon Zircon Zircon total rounded broken euhedral Tourmaline Rutile Monazite Apatite Amphibole Zoisite Sillimanite Epidote Others Opaque* ZTR Sum

1087

894

735

312 174

62 42

8 05

5 05

8 05

2 02

11 07

0 0

0 0

0 0

0 0

42 34

33 31

325 207

167 134

649

80 54

34 32

39 31

Fig. 5. (A) The relation CaO* + Na2OAl2O3K2O is plotted in combination with the chemical index of alteration (CIA) to show alteration trends (after Nesbitt & Young, 1984; Fedo et al., 1995). The CIA is calculated using molar proportions (Nesbitt & Young, 1982): CIA [Al2O3/(Al2O3 + CaO* + Na2O + K2O)]100; CaO*, only the CaO in silicates. An ideal alteration trend (straight line arrow) would be (sub)parallel to the ACN line (McLennan et al., 1990; Panahi & Young, 1997). K-metasomatism in the form of replacement of plagioclase by potassium feldspar greatly inuenced the alteration history (after Fedo et al., 1995). Kao, kaolinite; ill, illite; plag, plagioclase; ksp, potassium feldspar.

15 10

14 11

8 07

157 126

146 135

29 20

223 180

24 16

92 85

alteration of feldspar could produce a quartz-rich composition causing difculties in the interpretation of the CIA. Also, a few samples (e.g. A193, coarse grained greywacke; Table 3) have low concentrations of CaO (022%) and K2O (039%) and high Na2O values, which may reect albitization (e.g. Milliken, 1988). Major element abundances may in some cases reect the composition of the weathering prole mantling the source rocks rather than the bedrocks themselves (Nesbitt et al., 1996). The major element data presented here thus indicate that the rocks were affected by alteration to a minor degree. Provenance classication diagrams after Bhatia (1983) and Roser & Korsch (1986, 1988) using the major element chemistry show a wide spread for the three formations and cannot be used to interpret the tectonic setting successfully because of the mobility of K and Na, particularly during

323 218

45 42 78 110 464 650 172 240

27 30

313 340

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Tolar Chico Fm. Counts 920 Mean (%) 622 Tolillar Fm. Counts 714 Mean (%) 662 Diablo Fm. Counts 718 Mean (%) 578

580 630

165 230

460 640

93 130

49 39

1088

U. Zimmermann & H. Bahlburg

Table 3. Summary of selected major and trace element data for the Ordovician southern Puna Basin deposits. Tolillar Fm. (n 15/10) (major/trace element samples) Mean 573 7307 110 321 287 5580 1098 146 1359 3690 2680 471 420 12900 1403 873 089 067 018 683 435 1390 151 310 433 943 435 138 635 SD 77 568 120 160 156 2092 1340 053 414 1414 769 095 128 4082 356 224 008 028 003 448 172 275 054 142 172 261 227 136 475

Tolar Chico Fm. (n 9) Mean CIA SiO2 CaO Na2O K2O Cr Ni Ta Nb La Nd Sm Yb Zr Th Sc Ce/Ce* Eu/Eu* Sm/Nd LaN/YbN La/Sc Zr/Sc Th/Sc La/Th Th/U Zr/Th Cr/Th Cr/V Y/Ni 627 8776 023 021 203 8078 1176 067 618 2322 1878 291 171 19900 648 387 090 082 016 956 805 8180 228 359 522 3385 1493 319 137 SD 63 887 010 013 092 5011 1291 032 413 1228 986 157 108 5072 326 390 006 008 002 115 252 4047 082 038 388 832 992 269 078

Diablo Fm. (n 20/10) Mean 563 7294 156 331 217 4643 926 061 861 2660 2370 404 403 15300 1013 1281 107 079 017 409 235 1310 085 263 400 1589 593 052 025 SD 112 546 169 201 113 4152 1225 038 233 1175 955 163 141 3515 254 447 014 014 002 203 152 422 032 087 142 396 618 142 377 UCC* 66 42 39 34 83 44 1 12 30 26 45 22 190 107 14 107 066 017 505 214 1357 076 280 382 1776 776 078 050

Major elements in percentage, trace elements in p.p.m. Other trace elements were measured by INAA. Major and some trace elements were measured by XRF. SD, standard deviation (1 r) ; UCC after McLennan (2001). Further major (TiO2, Al2O3, Fe2O3, MnO and P2O5) and trace element data (REE, Sr, U, Rb, Cs, Ba, V, Hf, Y, Cu, Pb and Co) are available in Supplementary material, Table S2.

weathering of feldspar. This is consistent with the results of Bahlburg (1998) obtained from Ordovician rocks of the northern Puna.

Trace element geochemistry


Trace elements, such as the high eld strength elements Th, Sc and Zr and REE (rare earth elements), are particularly useful for provenance analysis as they are insoluble and usually immobile under surface conditions. On account of their typical behaviour during fractional crystallization, weathering and recycling, they preserve characteristics of the source rocks in the

sedimentary record (e.g. Taylor & McLennan, 1985; Bhatia & Crook, 1986; McLennan, 1989; McLennan et al., 1993; Roser et al., 1996). Trace elements such as Cr are useful in identifying accessory detrital components such as chromite, commonly derived from mac to ultramac sources including ophiolites, not readily recognized by petrography alone. The average Cr content of the upper continental crust is 83 p.p.m. (McLennan, 2001). Average Cr values of the studied formations decrease from the Tolar Chico Formations upper crustal value of  81 p.p.m. (n 9; SD 501) to  46 p.p.m. in the Diablo Formation (n 20; SD 415;

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Provenance of sedimentary rocks, Puna Basin Table 3). One sample from the Tolar Chico Formation and two samples from the Diablo Formation have Cr concentrations reaching 200 p.p.m. Cr-bearing minerals were not found in the studied heavy mineral concentrates (60150 lm), but may potentially be present as a minor constituent in the fraction < 60 lm. On the other hand, the relatively high average Cr/Th ratio (mean 149, SD 99) of the Tolar Chico Formation is mostly due to low Th concentrations. All the other samples have Cr/Th ratios lower than 78, the average upper continental crust (UCC) value (McLennan, 2001). Input from mac sources would also result in an enrichment of V and Ni. The respective values are much lower than in the upper continental crust leading to relatively high Cr/V and Y/Ni ratios in most of the samples (Table 3). This underlines the insignicance of a mac or ultramac contribution to the deposits (Floyd & Leveridge, 1987; McLennan et al., 1993). Another good tracer of mac source components is the compatible element Sc, particularly when compared with Th, which is incompatible and thus enriched in felsic rocks. Both elements are generally immobile under surface conditions and therefore preserve the characteristics of their source. Thus, the Th/Sc ratio is considered a robust provenance indicator (Taylor & McLennan, 1985; McLennan et al., 1990). The respective ratio in average upper continental crust is 079 (McLennan, 2001). From the Tolar Chico to the Diablo Formation, the average Th/Sc ratios decrease from 228 to 085 (Fig. 6, mean values from Table 3), close to or greater than the average crustal value. In the Tolar Chico Formation, both Th and Sc abundances are signicantly less than upper crustal values (Th/Sc mean 228; SD 082). This is

1089

almost certainly caused by quartz enrichment in the quartz arenites, producing a dilution effect also seen in the respective REE data. The Tolillar Formation shows a wide spread of values between 092 and 25 (mean 151; SD 054) resulting from the heterogeneous mineralogical composition. In the Diablo Formation, the average values (average 085; SD 032) obscure the fact that some samples have Th/Sc ratios between 056 and 070, which are lower than upper continental crustal values. Signicantly, these samples represent deposits intercalated with rhyolitic to andesitic lava ows (samples B44, B75, B86, B90, B96 and B130; Table 3). Th/Sc and Zr/Sc element ratios can reveal a compositional heterogeneity in the source(s), if the samples show Th/Sc and Zr/Sc values along the trend from mantle to upper continental crust compositions (McLennan et al., 1993; Fig. 6). Even though the data for the three studied formations cluster almost exclusively in the upper continental crust compositional eld, the lower Th/Sc values in the Diablo Formation reect a stronger input from a less evolved source. The Zr/Sc ratio is commonly used as a measure of the degree of sediment recycling leading to the enrichment of the stable mineral zircon in the deposits (McLennan, 1989; McLennan et al., 1993). The Zr/Sc ratio is highest in the Tolar Chico Formation (mean 818; SD 405) and far above the UCC value (Fig. 6; Table 3). Zr values range from 145 to 293 p.p.m. (average 199 p.p.m.) and scatter around the UCC average of 190 p.p.m. (Table 3; McLennan, 2001). The high Zr/Sc ratios are caused mainly by a Sc concentration far below the value of average upper crust (14 p.p.m.; McLennan, 2001). An average Zr/Th ratio of 3385 (SD 83), another measure of the degree of recycling, is well

Fig. 6. Plot of Th/Sc vs. Zr/Sc including the averages of the sedimentary rocks of the northern Puna for comparison. The values of the Tolar Chico Formation have to be interpreted carefully, as discussed in the text (modied after McLennan et al., 1993); , average; VS, volcanosedimentary successions; LTS, lower turbidite system; UTS, upper turbidite system (see also Figs 3 and 4).
2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

1090

U. Zimmermann & H. Bahlburg would have led to an increase in HREE (McLennan et al., 1993). This is supported by the heavy mineral analyses, which show nearly similar abundances of zircon in all three formations (Table 2), whereas concentrations of Zr and Hf are close to the average upper continental crustal value (Table 3). Eu/Eu* values lie between 069 and 097 (mean 082; SD 008), which reect the modest negative Eu anomaly. The REE patterns of the Tolillar and Diablo Formations show only small differences in the concentrations between ne- and coarse-grained rocks (Table 3; Fig. 7A). The samples of the Tolillar Formation show slight differences in the LREE and HREE concentrations. Two samples have slightly positive Eu anomalies with high Eu/ Eu* values (112 and 117) and reect a higher plagioclase concentration according to petrographic results (samples A296 and X27 in Table 1). Eu/Eu* values show a wide range between 044 and 117 (mean 067; SD 028). The LaN/YbN ratios are in general lower (mean 686; SD 448), with the exception of one sample and reect average UCC composition (505 after McLennan, 2001). Most samples of the Diablo Formation are slightly depleted in LREE and weakly enriched in HREE. The volcanic ash sample B44 of the Diablo Formation has a lower average abundance of LREE (Fig. 7A), which may have been caused by the loss of LREE during weathering of volcanic glass (e.g. Wood et al., 1976; Taylor & McLennan, 1985; Utzmann et al., 2002). One sample from the Diablo Formation has a slight positive Eu anomaly (Fig. 7A; Table 3, sample B90). The Eu/Eu* values of the Diablo Formation are generally higher (mean 079, SD 014) than in the Tolillar Formation. The LaN/YbN ratios decrease relative to the Tolillar Formation with all samples only slightly below average UCC (mean 409, SD 203; Table 3).

above the upper crustal average of 1776 and results from low Th concentrations. Th is present in many minerals but is commonly abundant in the heavy minerals monazite, zircon, titanite and in minerals of the epidote group such as epidote and allanite. Concentration of these heavy minerals during recycling would accordingly lead to an increase in Zr and Th abundances in the respective deposits. Low values of both Th and Zr therefore corroborate the evidence of the heavy mineral petrography, indicating that the mature quartz arenites of the Tolar Chico Formation do not show the zircon enrichment expected, given the maturity of the deposits. In contrast, the Zr/Sc and Zr/Th values of the Tolillar and Diablo Formations are very similar to the respective values of the upper crust (Table 3).

Rare earth elements


The REEs represent well-established provenance indicators (McLennan et al., 1990, 1993; McLennan, 2001). However, some mobility of REE may occur during weathering and diagenesis (Milodowski & Zalasiewicz, 1991; Zhao et al., 1992; Bock et al., 1994; McDaniel et al., 1994; Utzmann et al., 2002). The REE abundances and patterns of all three formations are characterized by an enrichment of LREE, pronounced negative Eu anomalies in almost all samples and relatively at HREE patterns. Most of the samples show patterns similar to the post-Archaean average Australian shale (PAAS) composite representing the postArchean upper continental crust (Fig. 7A and B; Taylor & McLennan, 1985). The quartz-rich rocks of the Tolar Chico Formation have patterns with similar shapes to PAAS but lower abundances due to a dilution effect by quartz (e.g. Cullers, 1995; Bock et al., 2000). This effect is borne out in Fig. 7A, which demonstrates that the Tolar Chico very negrained wackes have essentially the same REE patterns as PAAS, whereas the abundances in the quartz arenites are signicantly lower. The average LaN/YbN ratios (where the subscript N refers to chondrite-normalized abundances) of the samples decrease from 956 in the Tolar Chico Formation (SD 115) to 46 in the Diablo Formation (SD 203). The relatively high LaN/YbN ratio in the Tolar Chico Formation appears to be caused by a slight depletion of the HREEs. This indicates that the formation of these mature deposits has not been accompanied by an enrichment of heavy minerals, especially zircon, which

Implications of the trace and rare earth element ratios Comparison of the studied units with the average composition of the upper continental crust (McLennan, 2001) shows that the quartz arenites of the Tolar Chico Formation have lower values in relation to UCC, with the exception of Cr, Ba, Hf and Zr (Table 3). Cr appears in three samples in higher concentrations but, as shown above, Cr-bearing heavy minerals were not found in the heavy mineral spectra. In addition, the Cr/Th ratios are too low to point to an input from mac sources (Table 3). The low trace element concen-

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Provenance of sedimentary rocks, Puna Basin

1091

Fig. 7. (A) REE patterns of all samples of the three Lower Ordovician formations of the southern Puna. Stippled patterns represent the ne-grained samples in each formation. (B) REE element patterns of Ordovician volcanic and sedimentary rocks; pattern of PAAS, quartzites of the Cambrian Meson Group and siliciclastic deposits of the Puncoviscana Formation (Figs 1A and 3) for comparison; data from Bahlburg (1998) and Bock et al. (2000). Chondritic normalization of the sedimentary rocks and PAAS according to Taylor & McLennan (1985). The inset shows the REE spectra of crustally derived magmatic rocks of the Sierra Famatina (according to Pankhurst et al., 1998, g. 2).

trations may result from the high quartz content of the samples (e.g. McLennan et al., 1993; Cullers, 1995; Tables 1 and 3). Feldspar is rare in this formation as shown above (Table 1; Fig. 4A and B). This is consistent with low REE, Nb, Ta, Sc and Th contents typical of derivation from mostly granitoid sources (Cullers, 1995). The ne-grained

wackes of the same formation have higher concentrations of nearly all trace elements (Table 3), and of the REEs in particular (Fig. 7A). The negrained samples from the Tolar Chico Formation are considered to reect the composition of the source, a signal that is diluted in the quartz arenites by the high quartz content. Nearly all relevant

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

1092

U. Zimmermann & H. Bahlburg 1) and a probably young undifferentiated arc inuence (Eu/Eu*  100, Th/Sc < 10; McLennan et al., 1993). The Diablo Formation has slightly higher Eu/Eu* values (range 063107; mean 079; SD 014), but markedly lower Th/Sc ratios (mean 085; SD 032; Table 3). The REE patterns and the trace element ratios can be best explained by derivation from an evolved volcanic arc of largely UCC composition. This arc inuence is most dominant in the Diablo Formation, which is also consistent with the petrographic data. Trace element ratios such as La/Th, La/Sc, Zr/Sc and Th/Sc have been used successfully to discriminate tectonic settings (Bhatia & Crook, 1986). However, such an approach has to be used with caution because it has been shown that specic tectonic settings do not necessarily produce sedimentary rocks with unique geochemical signatures (McLennan et al., 1990; Bahlburg, 1998). La-Sc-Th and Th-Sc-Zr/10 ratios (Fig. 8) of the Tolar Chico Formation are typical of rifted margins. Those of the Tolillar Formation scatter over the active continental margin eld, whereas the Diablo Formation shows a continental island arc character.

trace element ratios (Sm/Nd, LaN/YbN, La/Sc, Zr/ Sc, Th/Sc and La/Th) show evolved values in relation to average continental crust (Table 3). The volcanogenic Tolillar and Diablo Formations most prominently display their overall upper continental crust composition in the trace element concentrations with slight enrichments in Yb and Y. Variations in both formations are pronounced (Table 3) and were most probably caused by the mixing of different sources. As shown in the Petrography section, the amount of volcanogenic material increased during the Arenig and from the Tolillar to the Diablo Formation. When the data are compared with average upper crustal values (not shown here), a marked negative Nb-Ta anomaly becomes evident in almost all samples but is most prominent in the Diablo Formation. Such a negative anomaly indicates the inuence of a volcanic arc in the source area (Hofmann, 1988, 1997). Despite the intraformational variations in the trace element ratios, there is a clear trend from the Tolillar to the Diablo Formation from a recycled sedimentary composition to one dominated by a continental arc source (Figs 4 and 8). This shift is observable using La-Th-Zr-Sc relations (see below) and Eu/Eu*. Most of the rocks in the Tolillar Formation show Eu/Eu* values between 044 and 063, and three samples have higher values. With very few exceptions, the Th/Sc ratios are uniformly 1 (mean 151, SD 054). According to McLennan et al. (1993), these data are in accordance with a dominant recycled sedimentary source (Eu/Eu*  060070; Th/Sc

Isotope geochemistry
Input of detrital material into a basin from different terranes with variable crustal compositions and histories can potentially be distinguished by analysing the Nd isotope systems of sedimentary rocks (e.g. McCulloch & Wasserburg, 1978; DePaolo et al., 1991; McLennan et al., 1993). The Tolar Chico Formation shows the most negative eNd(t sed) between )11 and )88 (mean )974; Table 4). In the Tolillar Formation, at the TremadocArenig boundary, the mean eNd(t sed) rises to )51. In the Diablo Formation, in turn, the mean eNd(t sed) value is only )195 (Table 4). Figure 9 plots Sm/Nd (the fractional deviation of the sample 147Sm/144Nd from a chondritic reference) of the southern Puna samples vs. eNd(t 470), where eNd(t 470) is eNd at 470 Ma, the sedimentation age. This allows a comparison of Sm/Nd and Nd isotope systematics at the time of sedimentation. The eNd values (Table 4) of the Tolar Chico, Diablo Formation and Falda Cienaga (Middle Ordovician rocks) are associated with Sm/Nd values between )043 and )031. However, two samples from the Tolillar Formation (A208 and A211) show clearly less negative Sm/Nd values ()021 and )0164 respectively) and may be the result of a resetting of the Nd isotopic

Fig. 8. (A and B) Tectonic discrimination diagrams for sandstones after Bhatia & Crook (1986). Note the loss of LREE, here La, in sample B44 in (A), where the same sample plots in the Th-Sc-Zr/10 triangle in the continental island arc eld similar to nearly all samples from the Diablo Fm. (B) A, oceanic island arc; B, continental island arc; C, active continental margin, including in (A) passive or rifted margin setting; D, passive or rifted margin.

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Table 4. Isotope geochemistry of the Ordovician sedimentary rocks and associated magmatic rocks.
147

Samples Rock type Age (Ma) Sm (p.p.m.) Nd (p.p.m.) 263 235 247 298 261 233 757 328 652 393 478 689 694 629 709 481 432 573 246 395 093 549 500 739 344 736 450 2023 2238 3290 2640 1838 2406 01262 01174 01224 01213 01206 01216 0512100 0512093 0512100 0512158 0512143 0512119 1299 2013 473 2561 2325 3265 1619 3276 2104 01143 01186 01193 01295 01299 01368 01286 01359 01266 0512270 0512268 0512159 0512386 0512348 0512374 0512385 0512369 0512320 0511913 0511897 0511785 0511982 0511943 0511948 0511985 0511946 0511925 0511709 0511733 0511725 0511782 0511773 0511744 )22 )25 )46 )08 )16 )15 )08 )16 )195 )63 )59 )61 )48 )51 )564 2887 1519 1750 4155 3350 3845 4282 2125 1993 2879 01364 01562 01652 01002 01252 00989 01001 01369 01310 01278 0512198 0512283 0512215 0512088 0512176 0512082 0512075 0512161 0512157 0512159 0511765 0511787 0511691 0511772 0511782 0511771 051176 0511726 0511741 0511755 )49 )44 )63 )48 )46 )49 )51 )56 )53 )510 )5 )45 )64 )49 )47 )50 )52 )57 )54 )520 )22 )25 )47 )09 )17 )16 )09 )17 )203 )63 )59 )61 )48 )51 )564 )0310 )0210 )0164 )0493 )0366 )0499 )0493 )0307 )0337 )0353 )0422 )0400 )0396 )0345 )0343 )0308 )0349 )0312 )0359 )0361 )0406 )0381 )0386 )0390 )0385 1114 1338 2044 870 986 867 890 1218 1117 1160 683 724 945 573 664 673 568 676 688 1165 1049 1104 972 993 1057 1408 1234 1297 1554 1422 1272 4023 1744 01129 01151 01151 01159 01109 01105 01138 01135 0511890 0511883 0511869 0511817 0511836 0511890 0511920 0511872 0511528 0511514 0511498 0511447 0511481 0511537 0511553 0511508 )94 )96 )99 )110 )103 )92 )88 )974 1360 1411 1436 1547 1425 1323 1320 1403 )96 )98 )101 )112 )105 )94 )90 )994 )0429 )0418 )0418 )0413 )0439 )0441 )0424 )0426

Sm/144Nd

143

Nd/144Nd

143

Nd/144Ndi eNd(t sed)* eNd(t 470 Ma) Sm/Nd TCHUR TDM TDM 1906 1925 1943 2013 1968 1896 1870 1931 1586 1554 1691 1579 1566 1583 1598 1641 1620 1602 1376 1400 1564 1269 1331 1324 1267 1327 1357 1679 1650 1662 1575 1593 1632 1839 1890 1910 2003 1883 1799 1812 1877 1796 2165 2733 1384 1614 1376 1400 1877 1755 1789 1305 1365 1544 1329 1402 1472 1318 1465 1400 1759 1615 1687 1577 1589 1645

Tolar Chico Fm. A6 qa A18 qa A76 qa A82 qa A88 qa A90 qa A92 g Mean

489 489 492 488 488 488 492

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Tolillar Fm. A194 g A208 a A211 a A299 g A303 g M5 g M7 g CU4 p CU6 a Mean

485 485 484 482 480 480 480 485 485

Provenance of sedimentary rocks, Puna Basin

Diablo Fm. B19 p B35 la B44 a B86 g B90 la B96 g B119 c B130 la Mean

477 478 478 476 476 476 475 475

Coquena QH45 VQ22 171 191 LN2 Mean

and Falda Cienaga Fm. g 473 423 are 468 435 are 468 666 g 473 530 are 468 367 484

1093

1094

Table 4. (Continued).
147

Samples Rock type Age (Ma) Sm (p.p.m.) Nd (p.p.m.) 553 767 373 01402 01570 01556 01960 08036 01758 0512824 0512866 0513171 0512507 0512507 0512389 0512341 0512395 69 60 70 )26 0512658 0512831 0512789 0512913 0512703 0512828 0512199 0512317 0512279 0512271 0510071 0512252 40 63 56 54 )376 50 3901 2830 3042 00857 01638 00741 0512284 0512390 0512412 0512017 0511879 0512181 )02 )28 30 )04 )30 28

Sm/144Nd

143

Nd/144Nd

143

Nd/144Ndi eNd(t sed)* eNd(t 470 Ma) Sm/Nd TCHUR TDM TDM )0566 )0171 )0625 488 1215 1153 1429 282 937 928 1893 1163 )0290 )54 867 )0205 )747 642 811 )0213 )564 6 896 )0008 ND 731 1715 3067 16 3477 )117 )0110 )1403 767 1190 )0282 )521 )0133 )1391 0281 1433 554 652 542 177 1077 662 950 89 733

U. Zimmermann & H. Bahlburg

Volcanic B13 B63 B66

rocks (Diablo Fm.) rhyolite 476 andesite 476 andesite 476

Plutonic rocks from Salar de Pocitos (CBOdC) Poc 5-5 gabbro 500 661 2851 A272 gabbro 500 295 1136 A220715 gabbro 500 270 1049 A4-6 gabbro 500 448 1382 A325 cumulate 500 529 398 A44 gabbro 500 164 564

Plutonic C15 C62 C137 829 6016 00833

rocks from the southern Sierra de Calalaste (SC) gabbro 468 742 3162 01419 gabbro 468 780 2751 01714 cumulate 468 067 160 02532

P450/44

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

*eNd(t sed) refers to the palaeontological controlled sedimentation age. Refers to a sedimentation age of 470 Ma, the age of the youngest sedimentation event discussed in the study, for comparisons with other formations in Figs 9, 10 and 11. Calculated according DePaolo et al. (1991). Calculated according DePaolo (1981). SC, southern Sierra Calalaste mac rocks; CBOdC, Complejo Basico Ojo de Colorado; qa, quartz arenite; g, greywacke; a, ash; are, arenite; la, litharenite; p, pelite; c, conglomerate; ND, not dened. Data for the magmatic rocks from Bock et al. (2000).

Provenance of sedimentary rocks, Puna Basin

1095

high glass content (e.g. Wood et al., 1976; Utzmann et al., 2002). In the Mid-Ordovician deposits, the inuence of younger material disappeared (Zimmermann et al., 2002; Falda Cienaga Formation, mean )564), and the formations returned to eNd(t sed) values typical of the Tolillar Formation and other Lower Palaeozoic sedimentary and magmatic rocks of this region (Figs 10 and 11; Tables 4 and 5; Bock et al., 2000).

Fig. 9. Plot of Sm/Nd vs. eNdt for samples from the Ordovician of the southern Puna, where eNdt is the isotopic composition at the approximate ages of deposition for each formation (see Fig. 3; Table 4). Note the differences between the arrays in the Tolillar Formation and the three other successions. This points to a redistribution of Nd isotopes at approximately the time of deposition (470 Ma).

system. Also, the Sm/Nd ratios of these two samples are much higher (023) than the mean of the Tolillar Formation (018; SD 003). The Sm/ Nd ratios of the remaining samples correspond to the mean of the formation and show PAAS-like REE patterns (Fig. 7A). Samples A208 and A211 were taken from feldspathic turbidites (A294, M5, M7) intercalated with quartz-rich siltstones (CU6, CU4) and pelites (A208 and A211). A large spread in Sm/Nd and a uniform isotopic composition can be noted at the time of sedimentation (470 Ma). Only the loss of a phase crystallized at the time of deposition will create a vertical array on a Sm/Nd vs. eNd(t) plot, whereas loss of older LREEenriched minerals will produce diagonal arrays (Bock et al., 1994). Fractionation of LREE during diagenesis can occur especially on the boundary of clay minerals during early diagenesis of unstable FeMn oxyhydroxides and volcanic phases (Milodowski & Zalasiewicz, 1991; Utzmann et al., 2002). Ce/Ce* anomalies are an indicator of an oxidizing environment, which could cause LREE mobility (McDaniel et al., 1994). Such anomalies were not observed, although A211 shows Ce/Ce* values of about 109 (Table 3). An early diagenetic rather than a weathering environment is favoured as the cause for the REE redistribution. The Diablo Formation is characterized by an average eNd(t sed) value of only )195 (Table 4), resembling isotopic data from the synsedimentary lava ows of the Diablo Formation, and consistent with a signicant input of less evolved material. Sample B44, interpreted as a volcanic ash, probably suffered the loss of LREE by weathering of

Implications of the isotope data The Tolar Chico Formation has the strongest negative eNd values at the time of deposition (Table 4; Figs 911). Calculation of model ages (TDM) from the Nd isotope data reveals that the Tolar Chico Formation has the oldest model ages, which cluster around values of 19 Ga. This indicates that at least one of the sources of this quartz arenite formation is considerably older than a large amount of the detrital material of the analysed exposed basement rocks of the Sierras Pampeanas and the Puncoviscana Formation in the neighbouring regions (Rapela et al., 1998; Bock et al., 2000; Figs 1A, 2, 3 and 10). As sedimentary rocks usually represent mixtures of several sources, a signicant part of the detritus must still be older than 19 Ga, i.e. early Palaeoproterozoic or even older. Potential source rocks or regions registering dated Palaeoproterozoic or even Archaean magmatic or metamorphic events do not occur in north-western Argentina but are present in the Arequipa Massif in southern Peru and northern most Chile (compilation in Bahlburg & Herve, 1997; Bock et al., 2000). They are also abundantly commonly recorded on the Brazilian shield (Sato, 1999) and occur in basement rocks of the Buenos Aires region of the Rio de la Plata Craton (see Figs 2 and 12B). Regions to the north of the Ordovician basin can be excluded as a source for the Ordovician rocks because of the sedimentary transport directions indicating a northward transport in the basin (see above; Bahlburg, 1990; Zimmermann, 1999; Zimmermann et al., 2002). The Tolillar Formation had negative eNd values at deposition ranging from )44 to )63. Calculated model ages (TDM) cluster around 16 Ga. Using the two-stage model of DePaolo (1981), the two disturbed samples (A211 and A208) show older, unrealistic values. With the exception of the Diablo Formation, the Arenig and Middle Ordovician formations of the southern and northern Puna have similar model ages to the basement rocks of the Pampeanas Terrane and the pre Ordovician sedimentary rocks (Meson Group and Puncoviscana Formation), which have TDM ages

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

1096

U. Zimmermann & H. Bahlburg

Fig. 10. Histogram of the Nd model ages (TDM) calculated according to the three-stage model of DePaolo et al. (1991). Included for comparison are data from Late Proterozoic to Cambrian sedimentary and magmatic rocks of Sierras Pampeanas and Sierra Famatina (Figs 1A, 2 and 3). The samples from the Diablo Formation are characterized by an input of slightly less fractionated, i.e. younger, material. Typical model ages of about 1518 Ga were found in all formations except the Tolar Chico and Diablo Formations and may indicate a western Gondwana crustal formation event presented by Sato (1999). Data for the Sierras Pampeanas from Rapela et al. (1998); for the Sierra Famatina from Pankhurst et al. (1998); for the Falda Cienaga Formation from Zimmermann et al. (2002); for the volcanic rocks of the Diablo Formation from Bock et al. (2000).

underlying formations and those of the entire Ordovician stratigraphy of the Puna. The input of less evolved material may have been derived from contemporary intermediate to felsic volcanic rocks, which have similar model ages (Table 4; Bock et al., 2000; Zimmermann, 2000). A co-variation of eNd data with Th/Sc ratios is consistent with dilution from young arc material (Fig. 11). Other potential source rocks with similar mixed Nd isotopic characteristics do not occur in the region.
Fig. 11. Th/Sc vs. eNd(t 470) plot. The results are consistent with those evaluated by petrological as well as geochemical methods (Figs 4 and 8). Diagram after McLennan et al. (1993); data for the Falda Cienaga Formation from Zimmermann et al. (2002).

that range from 14 to 18 Ga. More than 70% of these values fall in the range 14 to 16 Ga (Bahlburg, 1998; Pankhurst et al., 1998; Rapela et al., 1998; Bock et al., 2000; Lucassen et al., 2000; Zimmermann et al., 2002). This is consistent with the petrographic record of these formations, which shows a mixing of a minor volcanic component and a major amount of (recycled) detritus of cratonic provenance (Bahlburg, 1990, 1998; Zimmermann et al., 2002; this study). The model ages of the Diablo Formation range between 12 and 14 Ga (eNd(t sed) mean )195). They are signicantly younger than those of the

Mixing scenarios for the Tolillar and Diablo Formations Mixing models after DePaolo et al. (1991) have been used to model the probable sources for the Tolillar and Diablo Formations (Table 5). The debris constituting the Tolillar and Diablo Formations represents a mixture of an older metasedimentary source and a younger, magmatic one. Geochemical data for probable metasedimentary sources such as the Puncoviscana Formation (Late Proterozoic to Early Cambrian; Acenolaza et al., 1988) show comparable REE patterns to the Tolillar Formation (Bock et al., 2000; Zimmermann & van Staden, 2002). Therefore, the older source could in fact be represented by the metasedimentary rocks of the Puncoviscana Formation and mature rocks similar to sample A92 from the Tolar Chico Formation. The less evolved source was predominantly

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Provenance of sedimentary rocks, Puna Basin


Table 5. Mixing models after DePaolo et al. (1991) based on eNd A Source Tolar Chico Fm. (A92) Tolillar Fm. Puncoviscana Fm. Sierras Pampeanas Sierra Famatina Diablo Fm. Lava of Diablo Fm. CBOdC B Source 1 (s1) high fractionated TC (A92) PV C Source 1 (s1) high fractionated TC SF PV SP TC SF SP PV Source 2 (s2) low fractionated DBLAVA DBLAVA DBLAVA DBLAVA CBOdc CBOdc CBOdc CBOdc Diablo Fm. mix Average Average Average Average Average Average Average Average Beta 124 100 117 087 288 234 203 272 Alpha 093 082 090 080 072 048 044 064 Source 2 (s2) low fractionated SP SP Tolillar Fm. mix Average Average Beta 143 131 Alpha 093 090 Symbol TC TOL PV SP SF DB DBLAVA CBOdC
(t sed)

1097

values and Nd (p.p.m.) concentrations. eNd(t sed) )88 )51 )70 )49 )50 )20 )13 58 Nd (p.p.m.) 4023 2881 3803 2828 3263 2104 3255 1396

Age (approx.) Ma 490 485 560 520 500 476 476 472

% of source 2 93 90

% of source 2 93 82 90 80 72 48 44 64

A. The data represent the basis of the mixing calculations in B and C. The shale sample A92 was taken as representative for the Tolar Chico Formation because this ne-grained sample is considered to reect the source com position better than the quartz arenites of this formation (as discussed in the text). CBOdC, Complejo Basico Ojo de Colorado; unknown formation age. 472 Ma is the minimum intrusion age. B. Mixing models for the Tolillar Formation using as end-members the Tolar Chico Formation (sample A92) and the Puncoviscana Formation as old end-members. The younger end-member is represented by the magmatic rocks of the Sierras Pampeanas (Rapela et al., 1998). Beta(b) Nd(p.p.m)s1/Nd(p.p.m)s2; Alpha(a) b/(b + [(eNd(s2) ) eNd(mix))/ (eNd(mix) ) eNd(s1))]. C. Mixing models for the Diablo Formation. The Tolar Chico Formation, the Puncoviscana Formation and the magmatic rocks of the Sierras Pampeanas were selected as representative of the highly fractionated sources. These sources are mixed with the synsedimentary lava ows of the Diablo Formation and with mac and ultramac rocks (data in Table 4) exposed in the southern Puna (Zimmermann, 1999).

volcanic, as shown by the petrographic and geochemical data. With respect to the Tolillar Formation, the Ordovician volcanic rocks of the Puna are not possible source rocks because they are younger. Late Cambrian to very Early Ordovician volcanic and, to a lesser degree, plutonic rocks exposed in the southern Puna and to the south of the Puna in the Sierras Pampeanas and Sierra Famatina (Rapela et al., 1998; Table 5A) are more likely. This volcanic source, which was originally connected to the plutonic rocks, has been eroded away and may be preserved only as detrital grains in the Tolillar, and perhaps the Diablo, Formations. Deposits of the Tolillar Formation are therefore

modelled using the older low-grade to high-grade metamorphic source and a younger source that includes Late Cambrian to very Early Ordovician plutonic and volcanic rocks of the Sierras Pampeanas and Sierra Famatina (Rapela et al., 1998; Table 5A and B). The data are best modelled by a mix of Sierras Pampeanas-type metasedimentary basement material as well as Late Cambrian to very Early Ordovician felsic to intermediate magmatic debris (9095%) with 510% quartz-rich sedimentary material similar to the Tolillar Formation (Table 5B). This is consistent with the petrograpic as well as the geochemical data. A mixing model for the Diablo Formation (Table 5C) requires as the older source

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

1098

U. Zimmermann & H. Bahlburg analyses of individual quartz grains suggest a minor volcanic input of the order of 5%. Felsic synsedimentary volcanism is recorded in lower Tremadocian sedimentary rocks north of the study area (Las Vicunas Formation; Fig. 3). The constant element ratios in the quartz arenites reect either a homogeneous mixing or a strong inuence of physical sorting and/or weathering. The ne-grained rocks, in turn, preserve the geochemical and isotope geochemical characteristics of the source regions (Cullers, 1995). The chemical and isotopic geochemical signatures of the siltstones are consistent with a derivation from mostly (old) upper crustal sources (Figs 9 and 11). The most likely source region is represented by the Puncoviscana orogenic belt of Late Neoproterozoic to Early Cambrian age to the east of the Puna basin. This fold belt is composed of felsic plutonic, metasedimentary and subordinate volcanic rocks of upper crustal composition. Palaeocurrents are also consistent with this fold belt as a source region (Fig. 1A; Bahlburg, 1991; Zimmermann, 2000). Low-grade metaturbidites of the Puncoviscana Formation contain a signicant amount of metamorphic material and have generally younger Nd model ages and less negative eNd values than the Tolar Chico Formation (Bock et al., 2000; Lucassen et al., 2000). Preliminary geochemical and Nd isotope data on medium- to high-grade basement rocks show similar to slightly older model ages (Lucassen et al., 2000) than the low-grade metasedimentary rocks of the Puncoviscana Formation and the Lower Palaeozoic granitoids of the Sierras Pampeanas, which were used in the calculation of the mixing models. The Tolillar Formation received a larger input of detrital volcanic material of felsic and intermediate composition than the underlying Tolar Chico Formation and does not show signicant input from a metamorphic source. This is expressed in (1) higher abundances of volcanic quartz and lithoclasts and amphiboles; (2) a decrease in heavy minerals, metamorphic quartz and lithoclasts; (3) feldspar-rich arenites; and (4) geochemical characteristics pointing to an active continental margin (Fig. 8, McLennan et al., 1990, 1993). The geochemical and isotopic characteristics of the Tolillar Formation are similar to the Neoproterozoic Puncoviscana Formation (Bock et al., 2000; Zimmermann & van Staden, 2002), the Mid-Ordovician Falda Cienaga Formation (Zimmermann et al., 2002), the basement of the Pampeanas Terrane (Rapela et al.,

end-member rocks such as those of (1) the Tolar Chico Formation; (2) the metamorphic basement and plutonites of the Sierras Pampeanas; (3) the felsic to intermediate magmatic rocks of the Sierra Famatina to the south; or (4) the Puncoviscana Formation (Figs 1 and 2). For the other, less evolved and probably younger source, the Late Cambrian to very Early Ordovician eruptive rocks of the region could be considered and/or the mac to ultramac rocks of pre-Arenig age, including the Complejo Basico Ojo de Colorados (CBOdC) exposed in the southern Puna (Table 5A and C). However, an input of between 44% and 72% of this mac material is necessary to obtain the required isotopic values when combined with either of the older end-members (Table 5A and C). As shown above, neither the framework and heavy mineral petrography nor the trace element data including REE and Eu/Eu* values (Table 3; Figs 7, 8 and 11) indicate a signicant input of mac material. In view of this and the isotopic evidence, the absence of a signicant mac input for the mixing models is inferred. The best results are obtained when a basement source represented by the Sierras Pampeanas and/or the Puncoviscana Formation is mixed with the synsedimentary lava ows of the Diablo Formation. This mixture coincides very well with our petrographic data indicating a < 20% inuence of metasedimentary and magmatic basement material and 80% of felsic to intermediate volcanic detritus.

INTERPRETATION The Tolar Chico Formation is quartz rich and composed of mature recycled and well-rounded grains, a small fraction of angular quartz grains and rare alkali feldspar. Zircon abundances are lower than expected in quartz sandstones. Several features point to derivation of the quartz arenites from felsic metamorphic sources, including metasedimentary rocks and paragneisses, namely (1) the abundance of metamorphic quartz; (2) the presence of microcline and metamorphic lithoclasts; (3) the occurrence of metamorphic heavy minerals such as rutile, which is a widespread accessory mineral in metamorphic rocks, and tourmaline, which is abundant in granites and region and contact metamorphic rocks (Mange & Maurer, 1992); (4) the conspicuously low concentration of amphiboles; and (5) the fractionated geochemical characteristics comparable with average continental crustal compositions. CL

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Provenance of sedimentary rocks, Puna Basin 1998) and crustally derived plutonic rocks of the Famatinian volcanic arc (Pankhurst et al., 1998). This arc evolved on the basement of the Pampeanas Terrane (Pankhurst et al., 1998), today located south of the Puna (Fig. 1A, south of 27S; Fig. 12B). Also, the Tolillar Formation received most of its relatively immature detritus from nearby source regions by northward-directed axial palaeocurrents (present co-ordinates; Fig. 12B). A similar dispersal has been identied in the northern Puna Basin by Bahlburg (1990). The Diablo Formation contains the highest amount of detrital volcanic material (Fig. 4; Table 1), as well as, on average, the youngest detrital material (Figs 911; Table 4). The geochemical composition of the sedimentary rocks is

1099

similar to the synsedimentary volcanic rocks within the formation (Table 3). Nd isotope mixing models suggest that the latter may account for 80 85% of volcanic input in the Diablo Formation, assuming that the other mixing end-member is represented by the basement of the Pampeanas Terrane and/or the Puncoviscana Formation. The geochemical signature of the sedimentary rocks (Figs 8 and 11), including a negative Nb-Ta anomaly, points to a signicant inuence of material originating in a volcanic arc setting (e.g. Hofmann, 1988). The sedimentological features, the petrological, geochemical and isotope geochemical similarities of coeval sedimentary and volcanic rocks in this formation indicate that the Diablo Formation formed as a marine volcanic

Fig. 12. (A) Palaeotectonic reconstruction of the Ordovician active continental margin in the southern central Andes. No oceanic crust is formed in the Gondwana related retroarc basin. (B) Active margin setting during Late Arenig time (Diablo Formation). In the southern Puna, a retroarc basin evolved to the E of the NW- to SE-orientated PunaFamatinian volcanic arc (active since the lower Tremadoc) on continental crust. There is no geological evidence for allochthonous or even exotic crustal blocks (sketch modied after Rapela et al., 1998). AAT, Arequipa Antofalla Terrane; RPC, Rio de la Plata Craton; the white arrow shows the dominant sedimentation direction, and the shadowed oval indicates the intensity of deformation from high (darker part) to very low (white part).
2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

1100

U. Zimmermann & H. Bahlburg west to north-east (Mon & Hongn, 1991; Zimmermann, 2000; Fig. 12B). At the ArenigLower Llanvirn transition, the volcanic activity ended in the northern and southern Puna (Bahlburg, 1990; Bahlburg & Furlong, 1996; Moya, 1997; Zimmermann et al., 2002). Subsequent magmatic activity occurred in a large belt along the eastern border of the Puna in the Faja Eruptiva de la Puna Occidental (Fig. 1A, Early Palaeozoic plu tonics; Mendez et al., 1973). This activity was most probably related to Late Ordovician strike slip tectonics connected to the Ocloyic Orogeny (Bahlburg, 1990; Coira et al., 1999). During the Ocloyic Orogeny, the successions in the Puna were strongly deformed and subjected to great strain (Hongn & Mon, 1999). The interpretation of the plate tectonic evolution of the Puna basin is controversial and partly rests on the signicance of mac and ultramac rocks of assumed Ordovician age in the southern Puna (e.g. Ramos et al., 1986). These have been interpreted as oceanic crust ooring a basin, which became progressively oceanic southward. Accordingly, the end-Ordovician closure of the basin during the Ocloyic Orogeny should have resulted in the accretion of the western area of the Puna as a separate, potentially exotic terrane and in the preservation of the intervening ocean oor as an ophiolite in the southern Puna (Forsythe et al., 1993; Conti et al., 1996; Rapalini et al., 1999). However, the alleged ophiolites exposed in the southern Puna do not show the geochemical and petrological characteristics typical of oceanic crust (Nicolas, 1989; Zimmermann et al., 1999). The assumption of a southward-deepening and opening basin favoured by the interpretation of palaeomagnetic data is unrealistic as 90% of the ute mark palaeocurrent measurements (n 227) distributed over the entire basin and its stratigraphy indicate uniform axial transport towards the north (Bahlburg, 1990; Zimmermann, 2000; Zimmermann et al., 2002). Finally, the present study demonstrates that the detritus constituting the Ordovician sedimentary rocks of the Puna were probably supplied from neighbouring sources to the south and east of the basin. There are so far no indications of the input of exotic detrital material.

apron close to an eruption centre. Minor amounts of older crustal material were also transported into the depositional area. The deposits of the Tolillar and Diablo Formations are compositionally immature. The preservation of detrital material of plutonic, volcanic, sedimentary and, to a lesser extent, metamorphic origin in the same strata is indicative of only a minor weathering and sorting inuence acting on the detritus on its way from source to sink. In turn, this implies (1) relatively short transport paths and thus potentially marked relief; and (2) only minor intermediate storage of sediment. These features are typical of tectonically active basins at active continental margins or of strikeslip basins (e.g. Nilsen & Sylvester, 1995).

PALAEOTECTONIC EVOLUTION The collision of the Pampeanas Terrane with Gondwana between 535 and 515 Ma in the Lower to Middle Cambrian (Rapela et al., 1998; Keppie & Bahlburg, 1999) led to the deformation and metamorphism of the Puncoviscana Formation and medium- to high-grade metamorphic basement rocks. Subsequently, the siliciclastic and shallow-marine Meson Group was deposited unconformably above the metaturbidites of the Puncoviscana Formation in a small intracratonic basin during the Late Cambrian (Sanchez & Salty, 1990). During the Tremadoc, a subduction zone was initiated on the western margin of the Pampeanas Terrane, representing the western border of Gondwana (Fig. 12; Pankhurst et al., 1998). In the Puna Basin, the rst felsic volcanic rocks formed during the early Tremadoc in the Las Vicunas Formation (Salar de Rincon area; Fig. 1A, no. 5; Fig. 3; Moya et al., 1993). At the TremadocArenig boundary, volcanic input increased markedly in the Tolillar Formation. The overlying Diablo Formation, like the volcanosedimentary successions in the northern Puna (Huai tiquina and Aguada de la Perdz; Fig. 1A, no. 4) is dominated by volcaniclastic input and the deposition of primary volcanic material with a volcanic arc signature (this paper, e.g. Breitkreuz et al., 1989; Bahlburg, 1990). In the early Ordovician, the PunaFamatinian volcanic arc was built on the western border of the Pampeanas Terrane as represented by the Sierras Pampeanas (Fig. 12A and B), and a retroarc basin evolved to the east and north-east of this arc (Bahlburg, 1991; Zimmermann, 2000). The deformation of the basin ll increases from south-

CONCLUSIONS Petrographic, geochemical and isotopic data from the Tolar Chico through to the Diablo Formations

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Provenance of sedimentary rocks, Puna Basin dene a temporal compositional trend from a basement source to an arc volcanic source (Figs 8 and 11) in the southern Puna. This reects a change in provenance from an upper crustal source in the Tolar Chico Formation (Tremadoc), which was deposited at a rifted margin, to an overall andesitic arc composition in the Diablo Formation (Arenig). The amount of volcanigenic material delivered to the basin increased close to the TremadocArenig boundary. In the Tolillar Formation, the input of old basement material decreases correspondingly, but it preserves mineralogical, geochemical and isotopic evidence for a provenance in Precambrian to Cambrian (meta-) sedimentary successions (Meson Group and Puncoviscana Formation) as well as of the basement of the Pampeanas Terrane. The peak of active margin volcanism was reached during the Arenig in the Diablo Formation. It shows an input of slightly less evolved material mixed with detritus typical of the Tolillar and Puncoviscana Formations, the basement of the Pampeanas Terrane as well as the Famatinian magmatic arc. The immature volcanic debris with less negative eNd values was probably eroded from synsedimentary lava ows intercalated in the Diablo Formation, as it shows similar isotopic and trace element characteristics. The input of volcaniclastic material with its specic geochemical and isotopic composition was short lived and disappeared at the beginning of the Llanvirn when the volcanic sources of the Diablo Formation were recycled into the Mid Ordovician Coquena and Falda Cienaga Formations in the southern Puna and into the Puna Turbidite Complex of the northern Puna. This study shows that erosional debris with model ages over 20 Ga contributed signicantly to the detritus deposited in the Ordovician basin of north-western Argentina particularly as the Tolar Chico Formation of Tremadoc age. Input of mac material, implied by different palaeotectonic models, is neither indicated by the framework and heavy mineral compositions nor by geochemical and Nd isotope data. Average continental crust composition represents the provenance of most of the sedimentary rocks in the Ordovician Puna basin. Consequently, the orogenic processes at this active plate margin were dominated by recycling of pre-existing Proterozoic basement rocks of the Pampeanas Terrane. There are no indications of terrane suture zones, oceanic crust or the accretion of allochthonous or even exotic blocks. As there is no evidence of exotic sources, the Puna deposits reect the composition of the

1101

adjacent continental regions represented by the Sierras Pampeanas basement, i.e. the Pampeanas Terrane including the Puncoviscana fold belt.

ACKNOWLEDGEMENTS This study was funded by Deutsche Forschungsgemeinschaft grant Ba 1011/11-1 and Deutscher Akademischer Austauschdienst grant D/98/ 04324. We thank B. Bock for the geochemical data of the magmatic rocks, R. J. Pankhurst for his constructive comments on the isotopic data, as well as organizing perfect working conditions at the NIGL (Keyworth), Fernando Hongn for help in tectonic problems, and Cristina Moya as well as Jorg Maletz for the determination of the fossils. We also thank Diane McDaniel, Robert L. Cullers and Scott M. McLennan for their helpful and inspiring reviews. This paper is a contribution to IGCP projects 436 Pacic Gondwana margin and 453 Uniformitarianism revisited: a comparison between modern and ancient orogens.

SUPPLEMENTARY MATERIAL The following material is available from http:// www.blackwellpublishing.com/products/journals/ suppmat/sed/sed595/sed595sm.htm Table S1. Abundances of the framework minerals of the studied formations. All samples were counted according to the GazziDickinson method described by Ingersoll et al. (1984). is mean value. Table S2. Geochemical data of the Lower Ordovician sedimentary rocks. Major and some trace elements (1) were measured by XRF, other trace elements by INAA (2). SD, standard deviation (1 r); qu, quartz arenite; g, greywacke; a, ash; la, litharenite; p, pelite; c, conglomerate; phi, grain size; UCC * after McLennan (2001).

REFERENCES
Acenolaza, F.G. and Baldis, B. (1987) The Ordovician System of South America. Int. Union Geol. Sci., Spec. Publ., 22, 168. Acenolaza, F.G., Toselli, A.J. and Durand, F.R. (1975) Estra tigrafa y paleontologa de la region de Hombre Muerto, Provincia de Catamarca, Argentina. I. Congr. Argentino Paleontol. Bioestratigr., I, 109111. Acenolaza, F.G., Miller, H. and Toselli, A.J. (1988) The Puncoviscana Formation (Late PrecambrianEarly Cambrian). Sedimentology, tectonometamorphic history and age

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

1102

U. Zimmermann & H. Bahlburg


Ordovician to Tertiary age in the Wet Mountains region, Colorado, USA. Chem. Geol., 123, 107131. Dalla Salda, L., Cingolani, C. and Varela, R. (1992) Early Paleozoic orogenic belt of the Andes in southwestern South America: result of LaurentiaGondwana collision? Geology, 20, 617620. Dalziel, I.W.D. (1997) Overview: NeoproterozoicPaleozoic geography and tectonics: review, hypothesis, environmental speculation. Geol. Soc. Am. Bull., 109, 1642. DePaolo, D.J. (1981) Neodymium isotopes in the Colorado Front range and crust-mantle evolution in the Proterozoic. Nature, 291, 193196. DePaolo, D.J., Linn, A.M. and Schubert, G. (1991) The continental crustal age distribution: methods of determining mantle separation ages from Sm-Nd isotopic data and application to the southwestern United States. J. Geophys. Res., 96 (B2), 20712088. Dickinson, W.R. (1970) Interpreting detrital modes of greywacke and arkose. J. Sed. Petrol., 40, 695707. Dickinson, W.R. and Suczek, C.A. (1979) Plate tectonics and sandstone composition. AAPG Bull., 63, 21642182. Dickinson, W.R., Beard, L.S., Brakenridge, G.R., Erjavec, J.L., Ferguson, R.C., Inman, K.F., Knepp, R.A., Lindberg and Ryberg, P.T. (1983) Provenance of North American Phanerozoic sandstones in relation to tectonic setting. Geol. Soc. Am. Bull., 94, 222235. Fedo, C.M., Nesbitt, H.W. and Young, G.M. (1995) Unravelling the effects of potassium metasomatism in sedimentary rocks and paleosoils, with implications for paleoweathering conditions and provenance. Geology, 23, 921924. Floyd, P.A. and Leveridge, B.E. (1987) Tectonic environment of the Devonian Gramscatho basin, south Cornwall: Framework mode and geochemical evidence from turbidite sandstones. J. Geol. Soc. London, 144, 531542. Forsythe, R.D., Davidson, J., Mpodozis, C. and Jesinkey, C. (1993) Lower Paleozoic relative motion of the Arequipa block of Gondwana. Paleomagnetic evidence from Sierra de Almeda of Northern Chile. Tectonics, 12, 219236. Hofmann, A. (1988) Chemical differentiation of the Earth: the relationship between mantle, continental crust and oceanic crust. Earth Planet. Sci. Lett., 90, 297314. Hofmann, A. (1997) Mantle geochemistry: the message from oceanic volcanism. Nature, 385, 219229. Hongn, F.D. and Mon, R. (1999) La deformacion Ordovicca en gico el borde de la Puna. In: Relatorio 14 Congreso Geolo Argentino (Eds G. Gonzalez Bonorino, R. Omarini and J.G. Viramonte), pp. 212216. UNSa, Salta. Hubert, J.E. (1962) A zircon-tourmaline-rutile maturity index and the interdependence of the composition of heavy mineral assemblages with the gross composition and texture of sandstones. J. Sed. Petrol., 32, 440450. Ingersoll, R.V., Bullard, T.F., Ford, R.L., Grimm, J.P., Pickle, J.D. and Sares, S.W. (1984) The effect of grain size on detrital modes: a test of the GazziDickinson point-counting method. J. Sed. Petrol., 54, 103116. Keppie, J.D. and Bahlburg, H. (1999) Puncoviscana Formation of northwestern and central Argentina: Passive margin or foreland basin deposits? In: LaurentiaGondwana Connections Before Pangaea (Eds V.A. Ramos and J.D. Keppie), Geol. Soc. Am. Spec. Paper, 336, 139144. Kleine, T., Zimmermann, U., Mezger, K., Bahlburg, H. and Bock, B. (1999) Petrogenesis of a mac to felsic plutonic complex (Complejo Pocitos) and its host rocks in the southern Puna of NW Argentina. II Simposio Sudamericano pico, pp. 232234. de Geologa Isoto

of the oldest rocks of NW Argentina. In: The Southern Central Andes: Contributions to Structure and Evolution of an Active Continental Margin (Eds H. Bahlburg, C. Breitkreuz and P. Giese), Lect. Notes Earth Sci., 17, 2538. Allmendinger, R.W., Jordan, T.E., Kay, S.M. and Isacks, B.L. (1997) The evolution of the Altiplano-Puna plateau of the Central Andes. Annu. Rev. Earth Planet Sci., 25, 139174. Bahlburg, H. (1990) The Ordovician basin in the Puna of NW Argentina and N Chile: geodynamic evolution from back-arc to foreland basin. Geotekton. Forsch., 75, 1107. Bahlburg, H. (1991) The Ordovician back-arc to foreland successor basin in the ArgentineanChilean Puna: tectonosedimentary trends and sea-level changes. In: Sedimentation, Tectonics and Eustasy (Ed. D.I.W. Macdonald), Int. Assoc. Sedimentol. Spec. Publ., 12, 465484. Bahlburg, H. (1998) The geochemistry and provenance of Ordovician turbidites in the Argentine Puna. In: The ProtoAndean Margin of Gondwana (Eds R.J. Pankhurst and C.W. Rapela), Geol. Soc. London Spec. Publ., 142, 127142. Bahlburg, H. and Furlong, K.P. (1996) Lithospheric modelling of the Ordovician foreland basin in the Puna of northwestern Argentina: on the inuence of arc loading on foreland basin formation. Tectonophysics, 259, 245 258. Bahlburg, H. and Herve, F. (1997) Geodynamic evolution and tectonostratigraphic terranes of northern Argentina and northern Chile. Geol. Soc. Am. Bull., 109, 869884. Bhatia, M.R. (1983) Plate tectonics and geochemical composition of sandstones. J. Geol., 91, 611627. Bhatia, M.R. and Crook, K.A.W. (1986) Trace elements characteristics of graywackes and tectonic setting discrimination of sedimentary basins. Contrib. Min. Petrol., 92, 181193. Bock, B., McLennan, S.M. and Hanson, G.N. (1994) Rare earth element redistribution and its effects on the neodymium isotope system in the Austin Glen Member of the Normanskill Formation, New York, USA. Geochim. Cosmochim. Acta, 58, 52455253. Bock, B., Bahlburg, H., Worner, G. and Zimmermann, U. (2000) Ordovician arcs and terranes in NW-Argentina and N-Chile? Geochemical and isotope evidence. J. Geol., 108, 513535. Boenick, W. (1983) Schwermineralanalyse. Enke-Verlag, Stuttgart, 152 pp. Boles, J.R. and Franks, S.G. (1979) Clay diagenesis in Wilcox Sandstones of southwest Texas: implications of smectite diagenesis on sandstone cementation. J. Sed. Petrol., 49, 5570. Breitkreuz, C., Bahlburg, H., Delakowitz, B. and Pichowiak, S. (1989) Volcanic events in the Paleozoic central Andes. J. S. Am. Earth Sci., 2, 171189. Coira, B., Davidson, J., Mpodozis, C. and Ramos, V. (1982) Tectonic and magmatic evolution of the Andes of northern Argentina and Chile. Earth-Sci. Rev., 18, 303332. Coira, B., Kay, S.M., Perez, B., Woll, B., Hanning, M. and Flores, P. (1999) Magmatic Sources and tectonic Setting of Gondwana Ordovician Magmas, Northern Puna of Argentina and Chile. In: Laurentia-Gondwana Connections Before Pangaea (Eds V.A. Ramos and J.D. Keppie), Geol. Soc. Am. Spec. Paper, 336, 145170. Conti, C.M., Rapalini, A.E., Coira, B. and Koukharsky, M. (1996) Paleomagnetic evidence of an early Paleozoic rotated terrane in Northwest Argentina. A clue for Gondwana Laurentia interaction? Geology, 24, 953956. Cullers, R.L. (1995) The controls on the major- and traceelement evolution of shales, siltstones and sandstones of

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Provenance of sedimentary rocks, Puna Basin


Kostov, I. (1973) Zircon morphology as a crystallogenetic indicator. Kristallogr. Technica, 8, 1119. Krinsley, D.H. and Marshall, J.R. (1987) Sand grain textural analysis: an assessment. In: Clastic Particles Scanning Electron Microscope and Shape Analysis of Sedimentary and Volcanic Clasts (Ed. J.R. Marshall), pp. 215. Van Nostrand Reinhold, New York. Lucassen, F., Becchio, R., Wilke, H.G., Franz, G., Thirlwall, M.F., Viramonte, J.G. and Wemmer, K. (2000) ProterozoicPaleozoic development of the basement of the Central Andes (1826S) a mobile belt of the South American craton. J. S. Am. Earth Sci., 13, 697715. McCulloch, M.T. and Wasserburg, G.J. (1978) Sm-Nd and Rb-Sr chronology of continental crust formation. Science, 200, 10031011. McDaniel, D.K., Hemming, S.R., McLennan, S.M. and Hanson, G.N. (1994) Resetting of neodymium isotopes and redistribution of REEs during sedimentary processes. The Early Proterozoic Chelmsford Formation, Sudbury Basin, Ontario, Canada. Geochim. Cosmochim. Acta, 58, 931941. McLennan, S.M. (1989) Rare earth elements in sedimentary rocks: Inuence of provenance and sedimentary process. In: Geochemistry and Mineralogy of Rare Earth Elements (Eds B.R. Lipin and G.A. McKay), Mineral. Soc. Am. Rev. Mineral., 21, 169200. McLennan, S.M. (2001) Relationships between the trace element composition of sedimentary rocks and upper continental crust. Geochem. Geophys. Geosyst., 2, 2000GC000109. McLennan, S.M., Nance, W.B. and Taylor, S.R. (1980) Rareearth element-thorium correlations in sedimentary rocks, and the composition of the continental crust. Geochim. Cosmochim. Acta, 44, 18331839. McLennan, S.M., Taylor, S.R., McCulloch, M.T. and Maynard, J.B. (1990) Geochemical and Nd-Sr isotopic composition of deep-sea turbidites: Crustal evolution and plate tectonic associations. Geochim. Cosmochim. Acta, 54, 20152050. McLennan, S.M., Hemming, S., McDaniel, D.K. and Hanson, G.N. (1993) Geochemical approaches to sedimentation, provenance and tectonics. In: Processes Controlling the Composition of Clastic Sediments (Eds M.J. Johnsson and A. Basu), Geol. Soc. Am. Spec. Paper, 284, 2140. Mange, M.A. and Maurer, F.W. (1992) Heavy Minerals in Colour. Enke-Verlag, Stuttgart, 147 pp. Matter, A. and Ramseyer, K. (1985) Cathodoluminescence microscopy as a tool for provenance studies of sediments. In: Provenance of Arenites (Ed. G.G. Zuffa), pp. 191211. Kluwer Academic Publishers, Dordrecht. Mendez, V., Navarini, A., Plaza, D. and Viera, V. (1973) Faja gico Eruptiva de la Puna Oriental. V. Congreso Geolo Argentino, 4, 89100. Milliken, K.L. (1988) Loss of provenance information through subsurface diagenesis in Plio-Pleistocene sandstones, northern Gulf of Mexico. J. Sed. Petrol., 58, 9921002. Milodowski, A.E. and Zalasiewicz, J.A. (1991) Redistribution of rare earth elements during diagenesis of turbidite/ hemipelagite mudrock sequences of Llandovery age from Central Wales. In: Developments in Sedimentary Provenance Studies (Eds A.C. Morton, S.P. Todd and P.D.W. Haughton), Geol. Soc. Am. Spec. Publ., 57, 101 124. Mon, R. and Hongn, F. (1991) The structure of the Precambrian and Lower Paleozoic Basement of the Central Andes between 22 and 32 Lat. Geol. Rundsch., 80, 745758.

1103

Morton, A.C. and Hallsworth, C.R. (1999) Processes controlling the composition of heavy mineral assemblages in sandstones. In: Advanced Techniques in Provenance Analysis of Sedimentary Rocks (Eds H. Bahlburg and P.A. Floyd), Sed. Geol., 124, 329. Morton, A.C., Davies, J.R. and Waters, R.A. (1992) Heavy minerals as a guide to turbidite provenance in the Lower Paleozoic Southern Welsh Basin: a pilot study. Geol. Mag., 129, 573580. Moya, M.C. (1997) La Fase Tumbaya (Ordovcico Inferior) en gico los Andes del Norte Argentino. VIII Congreso Geolo Chileno, I, 185189. Moya, M.C., Malanca, S., Hongn, F.D. and Bahlburg, H. (1993) El Tremadoc temprano en la Puna Occidental Argentina. XII gico Argentino y II Congreso Exploracio n Congreso Geolo Hidrocarburos, II, 2030. Nesbitt, H.W. and Young, Y.M. (1982) Early Proterozoic climates and plate motions inferred from major element chemistry of lutites. Nature, 299, 715717. Nesbitt, H.W. and Young, Y.M. (1984) Prediction of some weathering trends of plutonic and volcanic rocks based on thermodynamic and kinetic considerations. Geochim. Cosmochim. Acta, 48, 15231534. Nesbitt, H.W., Young, G.M., McLennan, S.M. and Keays, R.R. (1996) Effects of chemical weathering and sorting on the petrogenesis of siliciclastic sediments, with implications for provenance studies. J. Geol., 104, 525542. Nicolas, A. (1989) Structures of Ophiolites and Dynamics of Oceanic Lithosphere. Kluwer Academic Publishers, Dordrecht, 367 pp. Nilsen, T.H. and Sylvester, A.G. (1995) Strike-slip basins. In: Tectonics of Sedimentary Basins (Eds C.J. Busby and R.V. Ingersoll), pp. 425457. Blackwell Science, Oxford. Panahi, A. and Young, G.M. (1997) A geochemical investigation into the provenance of the Neoproterozoic Port Askaig Tillite, Dalradian Supergroup, western Scotland. Precambrian Res., 85, 8196. Pankhurst, R.J., Rapela, C.W., Saavedra, J., Baldo, E., Dahlquist, J., Pascua, I. and Fanning, C.M. (1998) The Famatinian magmatic arc in the central Sierras Pampeanas: an Early to Mid-Ordovician continental arc on the Gondwana margin. In: The Proto-Andean Margin of Gondwana (Eds R.J. Pankhurst and C.W. Rapela), Geol. Soc. London Spec. Publ., 142, 181217. Ramos, V.A., Jordan, T.E., Allmendinger, R.W., Mpodozis, C., Kay, S.M., Cortes, J.M. and Palma, M. (1986) Paleozoic terranes of the central ArgentineChilean Andes. Tectonics, 5, 855880. Rapalini, A.E., Astini, R.A. and Conti, C.M. (1999) Paleomagnetic constraints on the evolution of Paleozoic suspect terranes from southern South America. In: Laurentia Gondwana Connections Before Pangaea (Eds V.A. Ramos and J.D. Keppie), Geol. Soc. Am. Spec. Paper, 336, 139144. Rapela, C.W., Coira, B., Toselli, A. and Saavedra, J. (1992) El magmatismo del Paleozoico en el Sudoeste de Gondwana. ico Inferior de Ibero-Ame rica (Eds J.G. Gutierrez In: Paleozo Marco, J. Saavedra and I. Rabano), pp. 2168. Universidad de Extremadura, Madrid. Rapela, C.W., Pankhurst, R.J., Casquet, C., Baldo, E., Saavedra, J., Galindo, C. and Fanning, C.M. (1998) The Pampean Orogeny of the southern proto-Andes: Cambrian continental collision in the Sierras de Cordoba. In: The Proto-Andean Margin of Gondwana (Eds R.J. Pankhurst and C.W. Rapela), Geol. Soc. London Spec. Publ., 142, 181217.

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

1104

U. Zimmermann & H. Bahlburg


basalts of eastern Iceland. Contrib. Mineral. Petrol., 55, 241 254. Zappettini, E.O., Blasco, G. and Villar, L.M. (1994) Geologa del extremo sur del Salar de Pocitos, Provincia de Salta, gico Chileno, I, Republica Argentina. VII Congreso Geolo 220224. Zeil, W. (1979) The Andes. A Geological Review. Gebr. Borntrger, Berlin, 260 pp. a Zhao, J.X., McCulloch, M.T. and Bennett, V.C. (1992) Sm-Nd and U-Pb zircon isotope constraints on the provenance of sediments from the Amadeus Basin, central Australia: evidence for REE fractionation. Geochim. Cosmochim. Acta, 56, 921940. Zimmermann, U. (1999) Sedimentpetrographische, geochemische und isotopengeochemische Methoden zur Bestimmung der Beziehung von Provenienz und Ablagerungsraum an aktiven Kontinentalrndern: Das ordovizische Back-Arca Becken in der Sud-Puna, Hochland im Nordwesten Argentiniens. PhD Thesis, University of Heidelberg, Germany, 282 pp. Zimmermann, U. (2000) The evolution of the Ordovician Southern Puna-Basin in NW Argentina a compilation. gico Chileno, I, 720725. IX Congreso Geolo Zimmermann, U. and van Staden, A. (2002) Neoproterozoic to Pre-Ordovician very-low to low-grade metasedimentary rocks from Sijan (Sierra de Ambato) and Campo Volcan gico (Puna) in north-western Argentina. XV Congreso Geolo Argentino, 2, 229234. Zimmermann, U., Mahlburg Kay, S. and Bahlburg, H. (1999) Petrography and geochemistry of southern Puna (NW Argentina) Pre-Late Ordovician gabbroic to ultra-mac units, intermediate plutonites and their host units: a guide to evolution of the western margin of Gondwana. XIV gico Argentino, 2, 143146. Congreso Geolo Zimmermann, U., Luna Tula, G., Marchioli, A., Narvaez, G., Olima, H. and Ramrez, A. (2002) Analisis de la proce dencia de la Foramcion Falda Cienaga (Ordovcico Medio, Puna Argentina) por petrografa sedimentaria, elementos trazas e isotopa de Nd. Asoc. Argentina Sedimentol. Rev., 9, 124.

Roser, B.P. and Korsch, R.J. (1986) Determination of tectonic setting of sandstone-mudstone suites using SiO2 and K2O/ Na2O ratio. J. Geol., 94, 635650. Roser, B.P. and Korsch, R.J. (1988) Provenance signatures of sandstone-mudstone suites determined using discrimination function analysis of major-element data. Chem. Geol., 67, 119139. Roser, B.P., Cooper, R.A., Nathan, S. and Tulloch, A.J. (1996) Reconnaissance sandstone geochemistry, provenance, and tectonic setting of the lower Paleozoic terranes of the West Coast and Nelson, New Zealand. NZ J. Geol. Geophys., 39, 116. Sanchez, M.C. and Salty, J.A. (1990) Litofacies del Grupo Meson (Cambrico) en el oeste del Valle de Lerma (Cordillera gico Argentino, II, Oriental Argentina). XII Congreso Geolo 129132. Sato, K. (1999) Superproduction evidence of the continental crust during Paleoproterozoic in South American Platform. II pico, pp. 361362. Simposio Sudamericano de Geologa Isoto Schfer, J. and Dorr, W. (1997) Heavy-Mineral analysis and a typology of detrital zircons: a new approach to provenance study (Saxothuringian ysch, Germany). J. Sed. Res., 67, 451461. Schneider, N. (1993) Das lumineszenzaktive Strukturinventar ttinger Arbeit. von Quarzphnokristen in Rhyolithen. Go a Geol. Palontol., 60, 181. a Suttner, L.J. and Basu, A. (1985) The effect of grain size on detrital modes: a test of the GazziDickinson point-counting method discussion. J. Sed. Petrol., 55, 616617. Taylor, S.R. and McLennan, S.M. (1985) The Continental Crust: its Composition and Evolution. Blackwell Scientic, Oxford, 312 pp. Utzmann, A., Hansteen, T.H. and Schmincke, H.U. (2002) Trace element mobility during sub-seaoor alteration of basaltic glass from Ocean Drilling Program site 953 (off Gran Canaria). Int. J. Earth Sci., 91, 661679. Warr, L.N. and Rice, A.H.N. (1994) Interlaboratory standardisation and calibration of clay mineral crystallinity and crystallite site data. J. Metam. Geol., 12, 141152. White, J.D.L. and Busby-Spera, C.J. (1987) Deep marine arc apron deposits and syndepositional magmatites in the Alisitos group at Punta Cono, Baja California, Mexico. Sedimentology, 34, 911927. Wood, D.A., Gibson, I.L. and Thompson, R.N. (1976) Element mobility during zeolite facies metamorphism of the Tertiary

Manuscript received 14 November 2000; revision accepted 3 February 2003.

2003 International Association of Sedimentologists, Sedimentology, 50, 10791104

Вам также может понравиться