BLOQUES Y SECTORES CUYA ALOCTONIA-AUTOCTONIA ES DISCUTIDA

PRECORDILLERA-BLOQUES DE SAN RAFAEL, LAS MATRAS Y CHADI LEUVU-PIE DE PALO (TERRENO COMPUESTO DE CUYANIA)

Y PATAGONIA

 

Precordillera

datacones nodules.gif (14695 bytes)

Dataciones de xenolitos Rincón Blanco (Precordillera)

En la Sierra de Pie de Palo, inmediatamente el este del Precordillera (imagenes satelitales y figuras), hay exposiciones extensas de basamento metamórfico que indican edades de cristalización de ~1.1 Ga y circones con edades metamórficas de ~1.0 Ga (Ramos et al., 1998). Al sur de Mendoza, en Ponón Trehué (Fig), una sucesión de calizas y rocas siliciclásticas del Ordovicico se apoyan sobre rocas metamórficas con edades circón U-Pb de ~1.2 Ga (Thomas et al., 2000). Principalmente en base a esta evidencia, se considera que el basamento de Cuyania es de edad Grenvilliana. Esto es consistente con el modelo de un origen Laurentiano porque el embayment ("la bahía") de Ouachita de donde se supone que el terrano Cuyania se separó, se ubica dentro de la Provincia Grenville a lo largo del margen del sur de Laurentia.

Una espesa sucesión carbonática cámbrica inferior a medio-ordovícica es particularmente distintiva en Precordillera y única en el Gondwana Occidental (Fig) y forma afloramientos largos, continuos norte- sur en los cinturones plegados y corridos orientales y centrales de la Precordillera y otros mas pequeños en Ponón Trehué y en la provincia de La Pampa. Su caracter poco profundo, con aguas cálidas, su sucesión estratigráfica y la afinidad de sus fósiles bentónicos lo hace similar a los estratos correlativos de Laurentia (Astini et al., 1995; Keller, 1999).

La sucesión típica incluye a las formaciones La Laja, Zonda, La Flecha, La Silla, y San Juan, con un límite Cámbrico-Ordovicico en Formación La Silla (ver Fig). En el Cerro Totora en la parte norte de Precordillera, la sucesión empieza con las fms Cerro Totora y Los Hornos que son facies lateral de la Fm La Laja (Bordonaro, 2003, ver fig.).

Con un intervalo espeso de yeso, limolitas rojas, arenisca, y dolomita, la Fm Cerro Totora se interpreta como sedimentitas continentales (fluviales, bajada y lacustre evaporítico) que se formó cuando Cuyania se separó del embayment de Ouachita en Laurentia (Thomas y Astini, 1999,; Thomas et al., 2001). La Fm Los Hornos Fm, así como La Laja y la sucesión de la Fm San Juan, se interpreta como un margen pasivo (Astini y Thomas, 1999) que se depositó durante la migración de Cuyania desde Laurentia.

La sucesión del calizas está cubierta por secuencias siliciclásticas del Ordovicico Medio a Superior que varían considerablemente, en litología, sucesión de facies, espesor, y edad (Fig). Según el modelo del bloque Laurentiano (Astini et al., 1995; Astini, 1998a,; Thomas y Astini, 2003), estas sucesiones representan una cuenca de foreland y cuña clastica sinorogénicas que se formaron como resultado del acercamiento de Cuyania contra el margen proto-andino de Gondwana. Estas sucesiones aparecen en el cinturón oriental y central de Precordillera, y junto con las sucesiones de calizas subyacentes componen las Tectofacies Oriental de Astini et al. (1995). A lo largo del borde occidental de las Tectofacies Oriental aparece un cinturón de facies que han sido colectivamente denominado como Fm Los Sombreros (Keller, 1999; Thomas y Astini, 2003) que está compuesta por olistolitos grandes dentro de una matriz pelítica que generalmente se correlaciona con Ordovícico medio.

Los olistolitos más grandes son bloques de calizas de edad Cámbrico y Ordovícico inferior que son similares o representan facies distal, de agua algo más profunda, de las suceciones de la plataforma Cámbrico a Ordovícica Medio (Fig anterior). Como se describió para el modelo Laurentiano, (Thomas y Astini, 2003 o Keller 1999, que propone una separación en el Ordovícico medio de Cuyania de Laurentia), la Fm Los Sombreros es considerada como depositada en el talud continental a lo largo del borde occidental del microcontinente de Cuyania, junto con los olistolitos generados por desplazamientos de gravedad a lo largo de escarpas de falla o erosión en cañones submarino a lo largo de ese talud o plataforma continental empinada.

La faja occidental de Precordillera están compuestas por estratos del Ordovicico Superior de las Fms Yerba Loca y Alcaparrosa y su equivalentes, que son turbiditas distales y pelitas con intrusivos máficos. Estas unidades, junto con la Fm Los Sombreros, fueron interpretadas por Keller (1999) como un contemporáneo lateral de las sucesiones proximales a distales del margen continental que va desde plataforma continental a facies de turbiditas.

LA EDAD Y AFINIDAD DEL BASAMENTO DE CUYANIA


Edad del basamento de Cuyania: Circones detríticos de las areniscas cámbricas y ordovícicas

Generalidades

La repetida asunción que el basamento de Cuyania es de edad Grenvilliana (Dalziel et al., 1996; Kay et al., 1996; Thomas y Astini, 1996; Dalziel, 1997; Ramos et al., 1998; Sato et al., 2000) es la línea primaria de evidencia que apoyaba la derivación de Cuyania del embayment de Ouachita en Laurentia. No sólo se basó en una evidencia geocronológica limitada, sino que no tiene en cuenta la distribución global del “Grenvilliano". La Provincia Grenville, uno de los orógenos más jóvenes de las fajas en el Escudo del Canadá , tiene su área tipo en el Canadá oriental. Allí hay un mosaico de terrenos geológicos formados por una serie de eventos orogénicos que se extendieron por casi 1 Ga de los cuales entre 1.71-1.60 Ga corresponden a la orogenia del Labrador y desde 1.08 a 0.98 Ga a la orogenia Grenvillian (Tollo et al., 2004).

A lo largo de los 4.500 km entre Labrador y México, la provincia Grenville registra: 1) un periodo temprano (1.3-1.2 Ga) de tectónica acrecionaria llamado orogenia Elzeviriana, 2) un periodo (1.18-1.08 Ga) de AMCG (anortosita-mangerita (granitoides portadores de piroxenos verdes, con composiciones ultramáfica a félsicas ~35-68 wt% SiO2)- charnockita-granito) y 3) un periodo final colisión continente-continente (1.08-0.98 Ga) a lo que varios autores se refieren a como la orogenia Grenvilliana (ej., Gower y Krogh, 2002) o el Pulso Ottawan de la Orogenia Grenvillian (ej., Karlstrom et al., 2001).

Aún mas, ya que varios orogenos colisionales contemperaneos se produjeron en varias paleoplacas durante el ensamble del supercontinente Rodinia, “el término orógeno Grenville se usa para referirse a
todas las áreas afectadas por orogénesis de tipo convergente durante el intervalo 1.3-1.0 Ga”
(Tollo et al., 2004, pág. 1), , incluso aquellos cinturones orogénicos que no involucraron colisión con Laurentia (vea Hoffman, 1991,: fig. 1A y figura del supercontinente Rodinia de capítulos anteriores).

Como consecuencia, las rocas con edades de 1.3 a 0.98 Ga son llamadas Grenvilliano, dándose a este término un sentido geocronológico. Para evitar confusión de las diferentes interpretaciones y para evitar la asunción que el término Grenvilliano quiere decir afinidad con Laurentia, se usa el término Mesoproterozoico para las dataciones de 1.6 a 1.0 Ga con sus eventos temprano y tardío (1.6 -1.3 Ga y 1.3-1.0 Ga, respectivamente).


Sólo en Ponón Trehué, cerca de San Rafael (ver Fig) los estratos de la plataforma ordovícica se apoyan sobre rocas del basamento. Areniscas que yacen sobre el basamento y que disponen de circones detríticos provenientes del mismo dan edades 1400-1100 Ma (pico en 1250 Ma) y así puede ser consideradas Grenvilliano en parte.

 

grenvilliano partes del mundo.JPG (130829 bytes)

pb-u.jpg (193644 bytes)

 

Relaciones 207Pb/204Pb vs 206Pb/204Pb y 208Pb/204Pb vs 206Pb/204Pb en rocas de Antártida, Precordillera, Islas Malvinas, Cratón de Arequipa, Escudos de Nubia (Africa) y Natal y Provincias Grenvillianas de Norte America y Europa Relaciones 207Pb/204Pb vs 206Pb/204Pb y 208Pb/204Pb vs 206Pb/204Pb en xenolitos de Precordillera

Sin embargo, Kay et al. (1996) no consideró la posibilidad de que esta signatura podría aparecer en otras rocas Mesoproterozoicas en otra parte en Gondwana. Seguidamente, Wareham et al. (1998) demostró una similitud íntima de isótopos de Pb y Nd, característicos de Laurentia en rocas del Mesoproterozoico en la Antártica Oriental y Occidental, en Natal (SE Africa), y en las Falkland/Malvinas , desafíando la asunción estricta de la afinidad Laurentiana del basamento de la Precordillera. Los defensores de la teoria Laurentiana (Astini y Rapalini, 2003,; Thomas y Astini, 2003,; Thomas et al., 2004) ha fallado en explicar los resultados de Wareham et al. (1998), que pueden ser consistentes con un origen Gondwánico para Cuyania.
Además, las roca del basamento de edad Grenvilliana de la Sierra de Pie de Palo que tienen la misma signatura isotópica de Pb que las rocas del basamento de Precordillera se consideran ahora autóctonas o parautoctonas a Gondwana basándose en los resultados de varias recientes investigaciones (Galindo et al., 2004; Baldo et al., 2005; McClelland et al., 2005; Rapela et al., 2005a)
.

Modelo parautoctono a la luz de los datos del basameno

Las edades Grenvilliana de circones detríticos dominan en muchas areniscas de Ordovícico de Precordillera, cuyas poblaciones se extenden entre 1600-1000 Ma (Finney et al., 2003; Gleason et al., 2007), indicando la presencia de Mesoproterozoico en muchas area del basamento de Cuyania. No obstante, la presencia de basamento Mesoproterozoico en Cuyania, incluyendo aquéllos limitados solo a edades Grenvilliana (1200-950 Ma) podría indicar una afinidad con Gondwana Occidental mas que con Laurentia.

Rocas con edades Grenvilliana componen la faja orogénica Sunsas, ubicada en el margen sudoeste del craton amazónico; ellos componen el cinturon Namaqua -Natal en el margen sur del craton de Kaapvaal, caracterizados por relaciones isotópicas de Pb bajo, y datos geocronológicos de circón recientemente adquiridos (Sims et al., 1998; Casquet et al., 2005; McClelland et al., 2005; Rapela al del et., 2005a, b) indican que la edad Grenvilliana son comunes en ambas Sierras Pampeanas occidentales y orientales de Argentina. Además, areniscas Neoproterozoicas depositado en el Cratón Río del craton de Plata tienen una cantidad significativa de circones detríticos de edad Mesoproterozoic a que incluyen un rango amplio de edades (1600-1000 Ma), encuentradas tambien en areniscas de la Precordillera y que indica la presencia de basamento de esa edad en el craton del Río de La Plata (Finney et al., 2006; Gaucher et al., 2006).


Arcos magmáticos con plutones tonalíticos a trondhjemiticos, la fuente potencial de la arenisca del Olistolito San Isidro, eran comunes en la parte del sur de Gondwana Occidental durante el Neoproterozoico a Cámbrico Temprano (Campos Neto y Figueiredo, 1995, Lira et al., 1997; Rapela et al., 1998; Llambías et al., 1998; Brito Neves et al., 1999; Basei et al., 2000, 2005, Veevers, 2003, Bossi y Gaucher, 2004). Plutonitas que cristalizaron a 1600-1500 Ma son comunes en Gondwana Occidental, particularmente en el sudoeste del craton amazónico (Tassinari y Macambira, 1999) y en el cratón Río de la Plata
(Gaucher et al., 2006) and could be the source of the distinctive early Mesoproterozoic zircons (1600-1500 Ma) in the Middle Cambrian and Middle Ordovician sandstones in the Precordillera.

Fuente potencial para todas las poblaciones de edades de circón en las areniscas de la Precordillera se pueden encontrar, por lo tanto, ya sea en las zonas que componen partes del Cratón del Río de la Plata adyacente a la Precordillera durante el Paleozoico temprano a neoproterozoica. En cambio, la posible fuente para las edades del Mesoproterozoico temprano, Neoproterozoica, y Neoproterozoica tardío a Cámbrico temprano no aparecen en aquellas partes de América del Norte que limitan con el Ouachita embayment y, en algunos casos, de la mayor parte de América del Norte. En consecuencia se cuestiona la validez de un microcontinente de origen Laurentiano microcontinent modelo es cuestionado, y un origen de Cuyania en el oeste de Gondwana, es decir, el modelo de parautochthonous Aceñolaza y Toselli (1988), Baldis et al. (1989), y Aceñolaza et al. (2002), se considera como una alternativa viable (Finney, 2005b, fig. 6).

 

Según el modelo parautoctono (Fig.), la migración de Cuyania comenzó durante el Ordovícico Medio a Superior a lo largo de una falla transformante de una ubicación en la margen sur de Gondwana occidental, y llegó a su actual posición en el borde occidental de la faja magmática del Famatina en el Devónico. Proponer este modelo alternativo a aquel de tipo Laurentiano requiere de un cuidadoso y detallado nuevo examen de todas las pruebas adicionales (estratigráficos, sedimentológicos, paleomagnético, y paleobiogeografico) que se utiliza para apoyar el modelo Laurentiano y de la compatibilidad de estas pruebas con el modelo parautoctono.

REEVALUACION DEL REGISTRO ESTRATIGRÁFICO DE CUYANIA


Significado de la sucesión estratigrafica de la plataforma calcárea del Cambrico-Ordovícico medio

La similitud entre las plataformas calcárea de margen pasivo de edad cámbrica-ordovícica media de la parte este de la Precordillera y de los Apalaches en Laurentia (Fig)

se cita frecuentemente como ina evidencia de la afinidad Laurentiana fr el terreno Cuyania (Astini et al., 1995; Dalziel et al., 1996; Thomas and Astini, 1996; Astini and Thomas, 1999; Keller, 1999; Thomas et al., 2004). Sin embargo estas similitudes no significan que los estratos de Cuyania se hallaren contiguos a aquellos de Laurentia. Ello requeriría que ellos no hubiesen sido desmembrados por rifting hasta el Ordovícico medio, como fue propuesto por Keller (1999). En adición, la sedimentación en una plataforma calcárea no se inició hasta el Cámbrico tardío en las zonas de Laurentia que rodea la bahía de Ouachita.

En el modelo de microcontinente Laurentiano, el rifting y la separación del terreno Cuyania ocurrió antes del Cámbrico Medio, por lo que las similitudes estratigráfica deben reflejar historias paleolatitud y de ascenso y descenso de nivel eustático que controla las historias de las facies sedimentarias y la acumulación de los estratos en Cuyania, mientras se separó y alejó de de Laurentia. Thomas y Astini (1996) propusieron que Cuyania migró desde el promontorio Alabama en Laurentia, en el Cámbrico Medio; en consecuencia su sucesión estratigráfica cámbrica sería probablemente más similar a la de los Apalaches del Sur. Sin embargo, existen diferencias sustanciales con las rocas clásticas de grano fino que domina gran parte del Cámbrico Medio y Superior de los estratos más australes de los Apalaches (Conasauga Grupo), en contraste con los estratos calcáreos predominantes en la Precordillera (Fig. 6; Resser, 1938; Palmer, 1971; Thomas, 1991).

In the parautochthonous model (Fig), Cuyania would have been at the same paleolatitude as Laurentia, experiencing the same eustatic events, and thus naturally could have been the site of carbonate sedimentation. Although southern West Gondwana was primarily the site of siliciclastic sedimentation during Cambrian and Ordovician time, following a long period of Brasiliano-Pan African orogenesis, carbonate sediments typical of a tropical climate did accumulate locally as recorded by the lowermost Cambrian Cerro Victoria Fm of Uruguay (Sprechmann et al., 2004; Gaucher et al., in press).

It is possible that Cuyania during Cambrian to Mid Ordovician time bore the same relationship to West Gondwana as Florida does today with North America, i.e. as a passive margin, carbonate platform free of siliciclastic sediment and extending into a tropical ocean.

 

Tectono-sedimentary significance of the Middle- Upper Ordovician stratigraphic record


According to the Laurentian model, the Middle-Upper Ordovician strata of Cuyania record a foreland basin and clastic wedge that formed as Cuyania docked with Gondwana (Astini et al., 1995; Thomas and Astini, 2003). The base of the Gualcamayo Fm, a black shale that overlies the San Juan Fm and records the drowning of the carbonate platform succession, is significantly older in the Guandacol region in the northern part of the Precordillera than at localities farther to the south (Thomas and Astini, 1996, 2003, among others). Thus, the basal contact is shown to be diachronous on correlation charts (Fig. ), becoming gradually younger between localities, and the Gualcamayo Fm is interpreted as representing deposition
in a subsiding, migrating peripheral foreland basin.

The coarse-grained, overlying siliciclastic successions (Las Vacas and Trapiche Fms in the north, the La Cantera Fm in the Villicum range to the south) are interpreted as coarser fractions of the clastic wedge, which arose from the collision of the Cuyania microcontinent with the proto- Andean margin of Gondwana and prograded westwards (Thomas and Astini, 2003).
Black shale also can be deposited in pull-apart or stike-slip basins along transform faults, just as has been the case with the organic-rich Monterey Fm of southern California - one of the world’s richest petroleum source rocks (Isaacs and Garrison, 1983).

Separate basins in the California borderland (e.g., Los Angeles, Santa Barbara, and Santa Maria) subsided to bathyal depths at approximately the same time, yet each had its own separate subsidence history, and the age of the Monterey Fm is somewhat different in each basin (Behl, 1999). This example presents a viable alternative explanation for the differences in age of the base of the Gualcamayo Fm in the Precordillera with sections at different localities recording different subsidence histories for different pull-apart basins.

Proponents of the Laurentian microcontinent model have not considered this alternative explanation.
Detailed study and evaluation of the Middle-Upper Ordovician siliciclastic strata overlying the Gualcamayo Fm in various sections in the Eastern Tectofacies of the Precordillera (Fig.) led Astini (1998a, b) and Keller (1999) to conclude that these strata were deposited in extensional basins (Astini et al., 1998a, figs. 7-9; Keller, 1999, fig. 58).

But this was not an original idea, having been proposed by many others, from Borrello (1969) to von Gosen et al. (1995), who have studied these rocks. These basins are characterized by considerable longitudinal and lateral basin asymmetry, including half-grabens, episodic rapid subsidence, abrupt lateral facies changes and local unconformities, greatly variable depositional processes within the same basin (pelagic settling, turbidity flows, mass flows, and gravity flows), marked contrasts in stratigraphy, facies geometry, and unconformities among different basins in the same region (Astini, 1998a, b).


Astini (1998a, 2002) has even used the term “yo-yo” tectonics to refer to the alternating up and down movement of crustal blocks as a result of extension. The Middle- Upper Ordovician stratigraphic successions are typically only 100-200 m thick; only in the Guandacol region does the succession approach 1000 m in thickness. Biostratigraphic correlation of these sections (Fig. ), plotted against an accurate time scale, demonstrates that the Middle-Upper Ordovician succession is dominated by major hiatuses with sedimentation occurring only in pulses of very short duration (Finney et al., 2005a). Sedimentation was episodic within each basin and nonsynchronous between basins
.

These are characteristics of strike-slip related basins along major transform faults (Christie-Blick and Biddle, 1985; Nilsen and Sylvester, 1995), and they are completely uncharacteristic of foreland basins (Busby and Ingersoll, 1995; Sinclair, 1997).
Olistoliths and boulder conglomerate of the Las Vacas Fm (Astini, 1998a, b) are particularly instructive. This thick, massive unit is part of a stratigraphic succession that includes the underlying shale of the Gualcamayo Fm and the overlying turbiditic siltstone/sandstone that dominates the Trapiche Fm (Astini, 1998b). Thus, this is not a coarsening, thickening upwards succession expected for a foreland basin.

The large rounded boulders of the Las Vacas conglomerate (clasts of meta-sandstone and quartzite and various plutonic and volcanic rocks) indicate erosion from mountainous terrane and transportation by rivers down steep gradients directly to the margins of the depositional basin. Most of the olistoliths are limestone blocks from the San Juan Fm; many are very large (> 1 m), and many are angular with flat surfaces representing bedding and fractures. And, everywhere, even where they are most abundant, the olistoliths are surrounded by rounded boulders of the conglomerate (Figs.).

These olistoliths were deposited with the boulders. It is clear that they were introduced into the stream of boulders only very near the site of deposition; otherwise that would have been pulverized by the boulders. Thus, the depositional basin of the Las Vacas Fm must have been adjacent to steep terrane with rivers that transported large boulders directly to the basin margin, and erosion and incorporation of limestone blocks must have occurred only as the river of boulders entered the depositional basin, the faulted margins of which were underlain by the San Juan Fm. The large, rounded boulders also are abundant in the Upper Member of the Gualcamayo Fm (Fig. ). They were delivered to the basin even during deposition of the shale that, according to the Laurentian microcontinent model, records the early subsidence of a peripheral foreland basin.

The Middle to Upper Ordovician strata were not deposited in a foreland basin; they represent a series of pull-apart or strike-slip basins with the Las Vacas having been deposited in a manner similar to that of fan-delta conglomerates in strike-slip basins, such as the Violin Breccia and Castaic Fm in the Ridge Basin of California (Link, 2003) and the San Onofre Breccia in the California continental borderland (Howell et al., 1974).


The Laurentian microcontinent model requires that the Middle-Upper Ordovician strata of the Precordillera be deposited in a foreland basin; thus, many very different attempts have been made to reconcile the clear evidence of strike-slip basins and related extension with the framework of a continent-continent collision. Initially, post-collisional relaxation and rebound were proposed (Astini et al, 1995; Astini, 1998a). More, recently thick-skinned thrusts were proposed to explain generation of the olistoliths of the Las Vacas Fm (Thomas and Astini, 2003; Astini and Davila, 2004), but without reconciling the co-occurrence of extreme compression required for the thrusts with the extensional regime of the adjacent basins in which the olistoliths were deposited, nor explaining the intimate depositional relationship of the olistoliths with the boulder conglomerate, nor explaining the occurrence of Las Vacas boulders in the Gualcamayo Fm. Most recently, Thomas and Astini (2005) proposed a third alternative, namely thin-skinned thrusts to explain the olistoliths. These hypotheses were proposed to explain the Middle-Upper Ordovician stratigraphy that is specific to the Guandacol area, a very small part of the Precordillera, but it is very different from correlative successions in the rest of the Precordillera and in the San Rafael region.
It is readily apparent that the Middle-Upper Ordovician stratigraphy of the Precordillera represents deposition in strike-slip basins. The Laurentian microcontinent model, on the other hand, requires that the stratigraphic record be interpreted in the context of a collisional orogenic setting. Different attempts to do so, however, have not been satisfactory.

The initial, most apparent interpretation of the Middle and Upper Ordovician stratigraphy, namely that it represents strike-slip related extensional basins (Borrello, 1969; von Gosen et al., 1995; Astini, 1998a, b; Keller, 1999; Gleason et al., 2007, and several others), is fully compatible with the parautochthonous model in which these strike-slip basins developed along a transform fault as Cuyania began its migration along the margin of West Gondwana. However, these kind of basins are incompatible with the Laurentian microcontinent model.


Significance of the Middle Ordovician K-bentonites

The occurrence of numerous K-bentonite beds in the upper San Juan Fm and overlying Gualcamayo Fm
of the Precordillera stratigraphic succession (Huff et al., 1998) is regularly cited as evidence for the Laurentian microcontinent model in which Cuyania was approaching and in proximity to the Famatinian magmatic arc in Mid Ordovician time (Astini, 1998a; Astini and Thomas, 1999; Thomas et al., 2002; Thomas and Astini, 2003; Astini and Rapalini, 2003; Fig. 9).


The age of the K-bentonite beds (~ 475-460 Ma) overlaps with that (~ 490-460 Ma) of Famatinian granitoids, and the geochemistry of the K-bentonites is consistent with their origin from subduction-related explosive volcanism (Huff et al., 1998; Fanning et al., 2004).

Thus, it is highly probable that the K-bentonites in the Precordillera succession represent ashes erupted from volcanoes of the Famatinian arc (Baldo et al., 2003; Fanning et al., 2004).

However, contrary to most interpretations, this does not support the Laurentian model in which Cuyania was approaching the Famatinian arc from the west (present coordinates). In recent paleogeographic reconstructions for Early and Mid Ordovician time (Dalziel, 1997, figs. 16; Cocks and Torsvik, 2002, figs. 4 and 5), the Famatinian arc faces to the east and is located between the equator and 30º S latitude. Assuming the same controls and general patterns of atmospheric circulation as those operating today, the prevailing winds in this paleolatitudinal belt, the southeast trade winds, would have blown and thus carried volcanic ashes to the northwest (Ordovician coordinates).

This direction is opposite that necessary to disperse the ashes to the Cuyania terrane if it was approaching from the east (Ordovician coordinates) as required by the Laurentian model
(Figs.).

Instead, the ashes would have been blown to the northwest to that part of Gondwana represented today by southeastern South America and southern Africa, where we place the Cuyania terrane in Mid Ordovician time. Confirmatory evidence for this interpretation is provided by the recognition of altered
pyroclastic beds in the Balcarce Fm of the Tandilia region, south of Buenos Aires (Dristas and Frisicale, 1987), which is correlated in part with the Lower Ordovician (Poiré et al., 2003; Rapela et al., 2005c). In citing the evidence from the K-bentonites, proponents of the Laurentian microcontinent model fail to ask
the question “Which direction was the wind blowing?” When that question is addressed, the occurrence of Middle Ordovician K-bentonite beds in the Precordillera is found to be consistent with the parautochthonous model. The paleolatitude used here to answer this question may be questioned. Regardless, the question does need to be asked.

REEVALUATION OF CAMBRIAN PALEOMAGNETIC DATA FOR CUYANIA
A paleomagnetically determined Cambrian paleolatitude of Cuyania is often cited as evidence for the origin of Cuyania from the Ouachita embayment of Laurentia. Rapalini and Astini (1998) established a paleomagnetic pole for red beds of the Lower Cambrian Cerro Totora Fm in Cuyania that indicated a paleolatitude of 20o S that was consistent with the paleolatitude of the Ouachita embayment of Laurentia and inconsistent with a Gondwanan location for Cuyania in its present position relative to South America.

However, they did not consider that Cuyania may have occupied a different position within Gondwana. A Cambrian paleolatitude of ~20º S places the Cuyania terrane on the southern margin of West Gondwana (present coordinates) using the Cambrian paleogeographic reconstructions of Hoffman (1991) and Dalziel (1997), and recent refinements in the early Paleozoic apparent polar wander path (APWP) for Gondwana (Meert, 2003; McElhinny et al., 2003) result in a nearequatorial position (~5-20º S) for the present southern tip of South America at ~530 Ma with a narrow Iapetus
Ocean separating it from Laurentia (McElhinny et al., 2003). This is consistent wit the paleogeographic reconstruction of Finney et al., (2005b, fig. 6; Fig. 5 herein), and thus is fully consistent with the parautochthonous model. The paleomagnetic data are consistent with a near equatorial location of Cuyania in the Cambrian.


REEVALUATION OF THE PALEOBIOGEOGRAPHIC RECORD OF CUYANIA


Changing affinity of Early and Mid Ordovician benthic and nektobenthic faunas

Most papers supporting the Laurentian microcontinent model (Astini et al., 1995; Thomas and Astini, 1996; Thomas and Astini, 2003; Thomas et al., 2004), or some variation on it that involves a Laurentian origin (Dalziel, 1997; Keller, 1999), invoke as evidence changes in affinities of Early and Mid Ordovician benthic faunas of the Precordillera. Typically, they include figures (e.g., Astini
et al., 1995, fig. 7; Thomas and Astini, 2003, fig. 2; Benedetto et al., 1995, fig. 1; Dalziel, 1997, fig. 7) showing that the percentage of genera of Laurentian affinity in the overall fauna decreases through the Lower and Middle Ordovician succession, while the percentages of endemic and Gondwanan genera increase.

Given the assumption that the larvae of benthic biota, such as brachiopods, can be dispersed across oceanic distances of only 1500 km, the changing composition of benthic faunas has been used, for example, to recognize an isolation stage for Cuyania when it was far removed from both Laurentia and Gondwana during late Early to early Mid Ordovician time, as well as earlier Laurentian and later pre-accretion and Gondwanan stages (Benedetto, 1998).
Thomas and Astini (1996, p. 756) went so far as to claim that from 495 Ma to 470 Ma (i.e. during Early to early Mid Ordovician time) “the Precordillera received no faunal migrants from either Laurentia or Gondwana, indicating ~ 1,000 km of open ocean separation from both Laurentia and Gondwana.”

More recently, Thomas et al. (2004) state: “Endemic faunas replaced the Cambrian Laurentian faunas by Early Ordovician time, indicating isolation of the Precordillera after rifting from Laurentia"

Brachiopods
It is necessary to challenge these interpretations of, and statements on, the paleobiogeography of benthic faunas of the Precordillera, which have become so widely repeated in the literature that they have become fundamental assumptions. The use of percentages gives a very different perspective to the data than does the use of actual numbers of genera. The only means to really check this is to use the percentages given for each brachiopod zone in fig. 7 of Astini et al. (1995), Fig. 12 herein, and the full list of genera for each zone found in table 1 of Benedetto (1998).

What one finds is that the percentage of Toquima- Table Head (or Laurentian) genera decreases in successive brachiopod zones through the Lower and lower Middle Ordovician: from 90% in the Archaeorthis zone to 60% in the Huacoella-Niquivilia zones to 65% in the Monorthis zone to 60% in the Ahtiella zone (Fig. 12). Yet, the number of genera varies greatly with the Ahtiella zone having nearly twice as many genera as any of the other zones.

If one uses the percentages and the total number of genera to determine the actual number of Laurentian genera appearing for the first time in each zone, one finds that that number increases dramatically in the Ahtiella zone (Table 1). These data conflict completely with the statement of Thomas and Astini (1996) regarding no faunal migrants reaching Cuyania from Laurentia in the Early to early Mid Ordovician. A large number of Laurentian brachiopod genera, 24 in fact, did reach Cuyania during this time, as well as an appreciable number of Laurentian trilobite genera (namely the genera Holia, Peltabellia, Uromystrum and Ectenonotus reported by Vaccari, 1994, p. 113).

These data also indicate that the number of Laurentian brachiopod genera that dispersed to Cuyania in the early Mid Ordovician is two to three times greater than the number that migrated during the time represented by any zone in the Lower Ordovician.

What does this mean? Obviously, the statement of Thomas et al. (2004) that endemic genera replace Laurentian genera in the Early Ordovician is not accurate given that the number of Laurentian genera actually increased. In addition, the data of Benedetto (1998) demonstrate that dispersal between Laurentia and Cuyania was far easier and more efficient and accomplished by more genera during the time, according to the Laurentian model, that Cuyania was nearing Gondwana and was farthest from Laurentia.
In subsequent papers (e.g., Benedetto, 1998; Benedetto et al., 1999) data from all the Lower Ordovician brachiopods zones (Archaeorthis to Monorthis) are grouped together as Arenig and compared to the data from the Ahtiella zone, a single zone taken to represent the Llanvirn. This, of course, presents a very different impression of faunal changes, with the Lower Ordovician faunas appearing to have far greater numbers of Laurentian genera in comparison to the Middle Ordovician faunas.

Regardless, the biogeography of the benthic faunas is compatible with the parautochthonous model. The primary control on the distribution of the Laurentian fauna, as it is for all benthic invertebrates, is water temperature, which, in turn, is related to paleolatitude. In addition, strong westward flowing (early Paleozoic coordinates) equatorial currents readily would have dispersed the larvae from Laurentia and to Cuyania, a carbonate platform located in the tropics on the southern margin of Gondwana

Given these paleolatitudes and current directions, Gondwanan taxa or taxa endemic to Cuyania would not have dispersed to Laurentia, a phenomenon noted and discussed by Benedetto (1998) and Benedetto et al. (1999).


Conodonts

Proponents of the Laurentian microcontinent model have also invoked the paleobiogeography of latest Cambrian to Mid Ordovician conodonts of Cuyania as supporting evidence. Again, the interpretations of this evidence are challenged. The latest Cambrian Clavohamulus hintzei of the North American Midcontinent conodont faunal province occurs in both Laurentia and Cuyania (Lehnert et al., 1997), and Astini (1998a) and Astini and Rapalini (2003) cite this as evidence to support the Laurentian microcontinent model.

However, the Midcontinent faunal province is not restricted to Laurentia; it is representative of shallow warm-water marine environments (Bergström, 1990); and C. hintzei, as well as other Midcontinent province conodonts, also occurs in Australia.

Thus, in the parautochthonous model, the occurrence of C. hintzei can be readily attributed to the low paleolatitude of Cuyania and to larval dispersal by oceanic currents. In upper Lower Ordovician strata of Cuyania, i.e. in the lower part of the San Juan Fm in the Precordillera and within Ponón Trehué Fm in the San Rafael area, conodont faunas of the North Atlantic province replace those of the Midcontinent province at a stratigraphic level corresponding to a global transgression (Lehnert et al., 1998). The North Atlantic conodont faunal province reflects coolwater and is developed in shallow-water platform settings at high paleolatitudes and in deeper-water outer shelf and slope settings at low paleolatitudes.

From the distributions of the Midcontinent and North Atlantic faunal provinces, Lehnert et al. (1998, 1999) concluded that Cuyania was derived from the southern margin of Laurentia at low paleolatitude and that by the late Mid to early Late Ordovician it was located at high paleolatitude, a considerable distance away from Laurentia and near to Gondwana - conclusions that were cited most recently by Astini and Rapalini (2003) as evidence of the Laurentian origin of the Precordillera.

This assertion is here challenged.
According to the Laurentian microcontinent model, Cuyania separated from Laurentia well before accumulation of strata with Midcontinent province conodonts, and although originally conceived as defining separate geographic provinces, the Midcontinent and North Atlantic conodont provinces are now interpreted as reflecting primarily differences in water temperature (Sweet and Bergström, 1984; Bergström, 1990).

Thus, the occurrence of Midcontinent conodonts in Cuyania reflects paleolatitude (expressed in water temperature) and oceanic larval dispersal and not necessarily any direct connection to Laurentia. Besides Cuyania and Laurentia, Midcontinent province conodonts are common in shallow-water sediments that accumulated at low paleolatitude in Australia in the Mid Ordovician (Webby et al., 2000), indicating that larval transport by oceanic currents could readily disperse Midcontinent conodonts from Laurentia to Cuyania on the southern margin of West Gondwana. The replacement of Midcontinent faunas by North Atlantic faunas correlates with a global sea-level rise and can be explained by a change in water temperature attributed to the sea-level rise and possibly also to the movement of Cuyania to higher paleolatitudes associated with the overall long-term migration of Gondwana (Fig. ).


Further complicating paleogeographic interpretations is the statistical analysis by Albanesi and Bergström (2002; in press) that demonstrates that conodont faunas of the Lower to Middle Ordovician successions of the Precordillera and the Marathon region of Texas become more similar upsection with conodont faunas of the lower Middle Ordovician of the Precordillera dominated by Laurentian taxa.

According to proponents of the Laurentian microcontinent model, these conodont faunas should become more dissimilar upsection as a result of the increasing distance and change in paleolatitudes between Laurentia and Cuyania. From this, it follows that conodont paleobiogeography must be used with caution in paleogeographic reconstructions.
Albanesi and Barnes (2000) invoked the Laurentian microcontinent model as a cause for major environmental perturbations and geographic isolation during deposition of the lower Gualcamayo Fm that provided the opportunity for allopatric speciation in the Paroistodus horridus to P. originalis conodont lineage by means of microevolution of endemic subspecies. In turn, Astini (1998a; Astini and Rapalini, 2003) used this interpretation, which is based on geographic isolation of intermediate subspecies, in support of the argument for the geographic isolation of the Precordillera.


However, given that Middle Ordovician conodont faunas of the Precordillera are diverse, that endemic species are few (and would be expected in diverse faunas), and that most species are globally widespread, the occurrence of two, short-lived endemic subspecies as evidence for drifting and isolation of the Cuyania terrane is surely overstated.

More surprising is the fact that the evolutionary lineage occurs in strata interpreted as foreland basin sediments recording the docking of Cuyania with Gondwana. Thus, on the one hand Astini (Astini and Rapalini, 2003) argues that the conodont lineage represents isolation of Cuyania, while on the other, he interprets the strata containing the conodont lineage to have been deposited during the collision of Cuyania with Gondwana (Astini et al., 1995; Thomas and Astini, 1996; and several others).

Neither Albanesi and Barnes (2000) nor Astini and Rapalini (2003) considered other models for producing the environmental perturbations and geographic isolation that may have influenced the micro-evolutionary event. The parautochthonous model with Cuyania as a carbonate platform on the margin of Gondwana, experiencing extension and subsidence of several pull-apart basins along a transform fault, surely would have.


Conclusion
Taken together, the evidence from the Early to early Mid Ordovician benthic faunas, primarily brachiopods and nektobenthic conodonts, is consistent with the paraautochthonus model but has several significant inconsistencies with the Laurentian microcontinent model, and in some instances these paleobiogeographic data have been misused or misrepresented.


Affinity of benthic shallow-water Cambrian trilobite faunas


The presence of shallow-water trilobites of Laurentian affinity in Cambrian strata in the Precordillera, particularly olenellid trilobites and the problematic fossil Salterella in upper Lower Cambrian strata, is the most compelling evidence for a Laurentian origin of Cuyania.

It is the ultimate argument used by proponents of the Laurentian microcontinent model to negate inconsistencies of their model with any other evidence (e.g., detrital zircon age populations). Nevertheless, Cambrian trilobite paleobiogeography is not inconsistent with the parautochthonous model. In fact, the abundance, diversity, and distribution of trilobites in Cambrian strata of the Precordillera are consistent with Cuyania having been a carbonate platform in tropical waters on the southern margin (present coordinates) of West Gondwana during the Cambrian.


According to the Laurentian microcontinent model (Thomas and Astini, 1996), Cuyania had migrated past the Alabama promontory and was completely separated from Laurentia by Mid Cambrian time and was well into the Iapetus Ocean by Late Cambrian time (Fig. ). The Laurentian model requires, therefore, that Mid and Late Cambrian shallow-water, Laurentian trilobites dispersed across oceanic waters in order to invade Cuyania.

Why couldn’t late Early Cambrian trilobites, largely olenellids, have done the same?


General Distribution


Cambrian trilobites are irregularly distributed through the carbonate stratigraphic successions of the Precordillera (Keller et al., 1994; Vaccari, 1994), which are composed of inner and outer platform carbonate facies with the former occurring in more or less continuous (autochthonous) sections exposed in the eastern and central Precordillera and the latter occurring only in olistoliths within the Los Sombreros and Estancia San Isidro formations and related strata along the western margin of the central Precordillera (Fig. ). Pelagic agnostids, abundant and diverse in outer platform facies, are cosmopolitan, but the benthic polymeroids of both inner and outer platform facies are mainly of Laurentian affinity.

In inner platform facies (the La Laja, Zonda, La Flecha and lower La Silla formations and, in the northernmost Precordillera, the Cerro Totora and Los Hornos formations), trilobites are common and diverse in only a few zonals intervals (Bordonaro, 2003), e.g. the Upper Cambrian Crepicephalus Zone (Fig. ).

Most zones are represented by few species (e.g. 4 species in the Olenellus Zone in the La Laja Fm, but only 1 for that zone in the Cerro Totora Fm; 2 in the Middle Cambrian Bolaspidella Zone; only one in the Upper Cambrian Aphelaspis Zone and two in the Upper Cambrian Saukia Zone).

Some zones (e.g. Cedaria Zone, and Dunderbergia to Prosaukia Zones) are not represented because coeval strata (e.g. the Zonda and Los Hornos formations and part of the the La Flecha Fm) lack trilobites; other zones (particularly Albertella and Plaguira-Poliella) are missing at the lower Middle Cambrian Hawke Bay hiatus (Bordonaro, 2003).

These gaps are similar to, but not identical to, the successions in the Appalachians, where some zones are not represented.


Upper Cambrian
In the Upper Cambrian La Flecha Fm, which bears the richest trilobite fauna of the Precordillera that includes ten Laurentian genera, many of the species are endemic to the Precordillera (35% according to Vaccari, 1994), and some of the genera are represented by species that are not known from the Appalachians, but instead from far removed areas, such as Montana (namely, Crepicephalus brevispinus and Pemphigaspis cagnasi). Other Laurentian genera are represented by few fragmentary specimens of uncertain identification (Coosella by one pygidium; Coosina by one pygidium; Kingstonia – one species by three pygidia and one unidentified species by three cranidia).


The La Flecha Fm, according to Vaccari (1994), which represents the nearly complete Upper Cambrian of the Precordillera, includes three trilobite zones: the Crepicephalus Zone with ten species representing nine genera (at least two of which are based on single fragmentary specimens not identifiable at species level); the Aphelaspis Zone with two species (one unidentified) representing two genera; the Saukia Zone with two species representing two genera. In an un-named unit of the outer platform facies near Mendoza, which according to Heredia (pers. comm., 2005) occurs in olistoliths in the Estancia San Isidro Fm, the lowest Upper Cambrian Cedaria Zone is represented by four or five different genera, only one of which is represented by an identifiable species, and that species is endemic (Bordonaro, 2003). In other olistoliths with outer platform facies, the Crepicephalus Zone is recognized by two identified species and two unidentified species representing four genera; the Elvinia Zone is recognized from four species (two Laurentian and two endemic) representing two genera; the Saukia Zone is based on endemic or unidentifiable species of three non-agnostid genera (Bordonaro, 2003).

In contrast, in the southern Applachians (Resser, 1938), trilobites are found only in the Nolichucky Fm of the lower Upper Cambrian. They represent only the Cedaria and Crepecephalus zones, which include, respectively, 17 species in 11 genera and more than 100 species in 16 genera. Higher trilobite zones are not represented in the southern Appalachians.

Similar large disparities in numbers of genera, relative to the Precordillera, have been reported for the Elvinia and Saukia zones of the central and northern Appalachians and in the more or less complete succession of Upper Cambrian zones of the Llano Uplift of Texas (Lochman-Balk and Wilson, 1958).


Middle Cambrian
Comparison between Middle Cambrian trilobite faunas of the Precordillera and Laurentia is similar to that of the Upper Cambrian with the faunas of the Precordillera being a small sample of those of eastern Laurentia and with several endemic species and genera. These are reviewed in descending stratigraphic order. Thirteen genera occur in the Bolaspidella Zone of eastern North America in primarily cratonic sequences of the Appalachians. Only two genera, both of Laurentian affinity but represented by unidentified species, have been reported from the La Laja Fm (Bordonaro, 2003).

In outer platform facies in olistoliths, this zone is represented by abundant, diverse agnostids and by five genera of polymeroids represented by endemic species (Borrello, 1971). The Bathyuriscus (or Ehmaniella) Zone is represented in the La Laja Fm by four genera with six species, all of which are endemic. In olistoliths, the coeval Oryctocephalus Zone includes five genera that also occur in Laurentia. However, two of these (Tonkinella and Oryctocephalus) occur on paleoplates that were distributed worldwide in tropical and temperate latitudinal belts during the Cambrian (Shergold, 1969; Gozalo et al., 2003). Coeval faunas from the southern Appalachians include 4 to 7 genera with 12 genera reported from the entire eastern United States.

The Glossopleura Zone is the lowest zone of the Middle Cambrian above the Hawke Bay Hiatus. It is found only in outer platform facies in olistoliths and is represented by five genera, two of which are endemic to the Precordillera.
Of the other three genera, one is known elsewhere only from the northwestern Mexico and is represented in Argentina by an endemic species, and a second is represented in the Precordillera by two endemic species (Borrello, 1971; Bordonaro, 2003). In contrast, the Glossopleura Zone in eastern North America contains more than 15 genera, and, in the southern Appalachians, it is represented by an average of 10-13 species in 8 to 10 genera. The Middle Cambrian trilobites of the La Laja Fm are currently under study by Osvaldo Bordonaro and Brian Pratt (pers. comm., 2005), who are discovering that a higher percentage of the species are Laurentian rather than endemic.

Nevertheless, the Middle Cambrian trilobite faunas of the Precordillera, as with those of the Upper Cambrian, are of lower diversity than those of the Appalachians and include a significant number of endemic taxa. This comparison to faunas of the Appalachians is considered essential for testing the Laurentian microcontinent model because Middle Cambrian strata accumulated only in narrow belts on the eastern and western margins of Laurentia and, according to the model, Cuyania migrated past the southern end of the eastern belt during the Mid Cambrian (Fig.)

preco_en_ouachitas.gif (81409 bytes) transferencia de Preco desde Laurentia.gif (44960 bytes) evolucion acresion preco-gondwana.gif (142555 bytes)
Ubicación del Terreno Precordillera en Laurentia Migración del Terreno Precordillera Acresión del Terreno Precordillera al margen de Gondwana


Lower Cambrian


Olenellid trilobites inhabited shallow seas in eastern North America and in Cuyania in the late Early Cambrian, and the olenellids of the Precordillera are genera considered to be of “Laurentian” affinity. However, in order to truly evaluate this relationship, one must first appreciate the taxonomic composition of the olenellids and their global distribution through their entire Early Cambrian history.
Olenellids are a large taxonomic group, the Suborder Olenellina (Table 2) within the Order Redlichiida and restricted to the Lower Cambrian, except for one genus. The Order Redlichiida also includes the Suborder Redlichiina with two superfamilies, the superfamily Redlichioidea largely restricted to the Lower Cambrian but with a few genera ranging into the lower Middle Cambrian, and the superfamily Paradoxidoidea that ranges from the upper Lower to the Middle Cambrian.
The olenellids characterize shallow shelf trilobite faunas of Laurentia, Siberia, Gondwana/ Perigondwana (limited to Portugal, Spain, Germany, and Morocco), Avalonia, and Baltica (Fig. ).

Redlichiid trilobites characterize shallow shelf trilobite faunas of Antarctica, Australia, the various paleoplates of China and East Asia, basically East Gondwana and Kazakhstan. Trilobites of both realms overlap in Morocco, southern Europe and Siberia (Pillola, 1990; McKerrow et al., 1992).
There also was paleobiogeographic differentiation within the olenellids with most genera restricted to one of five paleoplates: Laurentia, Siberia, Gondwana/Perigondwana, Siberia, Baltica, and Avalonia. Figure below 2 show the distribution of olenellid genera among these paleoplates. The argument for the Laurentian affinity of the Precordillera is based on the restriction to Laurentia of trilobites of the genus Olenellus and closely related genera. However, this assumption must be re-evaluated for the following reasons.

1. Some olenellids occur on more than one paleoplate.
Fallotaspis has for instance been recorded from Laurentia, Siberia, Gondwana, and Avalonia. Kjerulfia is known from Baltica, Avalonia, and Gondwana (Morocco). Holmia occurs in Baltica, Siberia, and Gondwana. Paranevadella and Nevadia occur in Laurentia and Siberia.


2. Much of the paleogeographic differentiation of olenellids reflects geologic time (stratigraphic position). Olenellus and the other closely related genera, naturally grouped together in the Family Olenellidae, evolved and diversified during the late Early Cambrian, a time in which few other olenellids existed elsewhere in the world (Fig. ), having largely gone extinct on the other paleoplates at the end of mid Early Cambrian, probably due to unfavorable environmental conditions. Thus, although genera of the Olenellidae evolved and diversified within Laurentia, their chances of dispersing to other paleoplates were minimal in light of the fact that conditions must not have been favorable on the other paleoplates for whatever reason.

The opposite appears to be the case in the Early Cambrian when olenellids of the Family Fallotaspidae were able to evolve and diversity on the Siberian and Gondwana/Perigondwana paleoplates, while conditions were unfavorable for olenellids in Laurentia, Avalonia, and Baltica. One of the earliest olenellids Fallotaspis, however, was able to disperse among Siberia and Gondwana/ Perigondwana and then later to Laurentia and Avalonia (Lieberman, 2002).

The mid Early Cambrian was a time of diversification of olenellids on all paleoplates and dispersal between them ( Fig.) . In the mid Early Cambrian, several genera occur on more than one paleoplate (Kjerulfia, Holmia, Nevadia, Paranevadella and possibly Andalusiana and Selindella). In addition, most families of mid Early Cambrian olenellids (Holmiidae, Archaeaspididae, Judomiidae, Nevadiidae) are composed of genera distributed among more than one paleoplate (Table 2).

Thus, when diversity of olenellids was highest and favorable conditions were present on most paleoplates, there was significant dispersal of olenellids between paleoplates.


3. Olenellids in the Precordillera are stratigraphically restricted, correlating with a very short interval in the upper Lower Cambrian – the middle and lower upper Olenellus zone (Fig. ). This correlation is based on the occurrence in the Precordillera of species of four genera/subgenera: Olenellus (Mesolenellus) based on one endemic species, Fremontella based on one endemic species, Bristolia based on specimens unidentifiable at species level, and Arcuolenellus based on one endemic species. The number of available specimens as reported in the literature is very small. The recognition of Fremontella is based on three small fragments of cephalons from the Villicum range (Borrello, 1963) and one small fragment of a cephalon from the Sierra de Zonda (Borrello, 1964), both in the La Laja Fm.

Bordonaro (1986) later described Olenellus (Mesolenellus) from the La Laja Fm from the Sierra de Zonda on the basis of seven listed fragmentary specimens of the species O.(M.) zondaensis and five very fragmentary specimens of an unidentified species. Also, he reported seven very small fragments as an unidentified species of Bristolia. In contrast, Arcuolenellus megafrontalis is based on a large number (~ 30) of very well preserved specimens from the Cerro Totora Fm (Vaccari, 1988), but these are from only a single, thin (~ 20 cm) shale interval at the very top of the formation and nowhere else in the section that is several tens of meters thick. One incomplete glabella from the Los Túneles
olistolith in the Río Jáchal section is referred to Olenellus? (Vaccari and Bordonaro, 1993). From this olistolith, Vaccari and Bordonaro (1993) also described a new, endemic species of the ptychoparid trilobite Sombrerella, which elsewhere occurs only in the Lower Cambrian of Sonora, Mexico.

Finally, Astini et al. (2004) report olenellids, unidentified at family, genus, or species level, occurring with Salterella in the Ancaucha olistolith. This is the total report of olenellids from the Precordillera.


4. Lower Cambrian strata and thus olenellids are absent in that part of Laurentia surrounding the Ouachita embayment. The nearest Lower Cambrian strata with trilobites are to the east in the southernmost Appalachians of Alabama and to the west in Sonora, Mexico, eastern California, and Nevada. In the Appalachians, the oldest trilobites are those of the upper Lower Cambrian, in particular the middle and upper parts of the Olenellus Zone; thus they are coeval with the olenellids of the Precordillera.


They occur from Alabama to western Newfoundland in a vertical facies succession that includes Antietam Quartzite, Shady Dolomite, and the Rome Fm in the southern Appalachians and comparable facies to the north (Resser, 1938; Palmer, 1971). A total of five species of Olenellus and one of Wanneria have been reported from numerous localities in the southern Appalachians. In the western Cordillera of North America, olenellid-bearing strata range from the upper “Fallotaspis” Zone to the top of the Lower Cambrian (Fig. ), and almost all of the Laurentian olenellid genera have been reported from this region with many genera represented by many species (e.g. Palmer and Halley, 1979). Of the “Laurentian” genera in the Precordillera, Fremontella is known only from Alabama; Arcuolenellus is known only from the western Cordillera; Bristolia is known from the southern western Cordillera and Greenland; Olenellus (Mesolenellus) is known from the western Cordillera and Greenland.

5. Given the information in 3 and 4, it follows that the olenellid fauna of the Precordillera is a very small sample, in terms of both diversity and abundance, of the coeval olenellid fauna of Laurentia. In addition, although Cuyania must have been in part adjacent to the southern Appalachians in late Early Cambrian time, according to the Laurentian microcontinent model, the fauna of the Precordillera is endemic at the species level and its genera are more representative of the western Cordillera.


6. The enigmatic fossil Salterella together with unidentified (or unidentifiable) olenellids in the Ancaucha olistolith is offered as convincing evidence of the Laurentian origin of Cuyania (Astini et al., 2004). As with genera of the Olenellidae, Salterella is known only from Laurentia, where it is widespread and occurs with Olenellus and associated genera. However, Salterella is most abundant and widespread in Laurentia in carbonate rocks, such as the Shady Dolomite of the southern Applachians and correlative carbonates to the north, that contain not only Olenellus but also archaeocyathids and the olenellid Wanneria (Fritz and Yochelson, 1988; Resser, 1938; Palmer, 1971). Therefore, it is notable that Wanneria and archaeocyathids do not occur in the Precordillera in coeval carbonate rocks i
n the La Laja Fm and in skeletal limestones with Salterella in the Ancaucha olistolith.


Given the reasons listed above, the location of Cuyania within southern Laurentia in the Early Cambrian is questioned. If it was, its outer margin would have been more-or-less continuous with Lower Cambrian strata of the southern Appalachians. Why are the olenellids of the Precordillera not more abundant and diverse? Why are the species endemic? Why are there more genera from the western Cordillera? Why are there no archaeocyathids and Wanneria with Salterella?


On the other hand, the Cambrian trilobite evidence is compatible with the parautochthonous model. On the southern margin of West Gondwana, Cuyania in the Early Cambrian would have been a carbonate platform in the tropics, west of Laurentia and directly within the path of westward flowing equatorial currents. Almost all of West Gondwana was mountainous terrain covered with siliciclastic sediment following the Brasiliano/Pan-African orogenies.

Given the complete absence of trilobite bearing Cambrian strata in all of West Gondwana except Morocco, far removed at high southern paleolatitude, the only way in which trilobite faunas could have populated Cuyania would have been by larval dispersal from Laurentia, and that larvae likely would have included “Laurentian” olenellids as well as Salterella. Several genera of olenellids were able to migrate across narrow oceans and between paleoplates in early Early and mid Early Cambrian time.

Thus, there is no reason to assume that some genera could not do so again in late Early Cambrian time.
Per Ahlberg (pers. comm., 2005), who has worked and published extensively on olenellids, supports this in a written communication quoted as follows: “Endemism is so pronounced among the olenellid trilobites and it is obvious that they could not disperse across wide oceans. Some genera, however, seem to occur on more than one paleocontinent. Fallotaspis has for instance been recorded from both Siberia and Gondwana, Kjerulfia is known from Baltica and Gondwana (Morocco), and Holmia occurs in Baltica and Siberia.

Thus, it appears likely that their larvae could disperse across narrow oceans, provided that the environment was roughly the same on both sides of the ocean.Your model with dispersal of Olenellus and Olenellus-like trilobites from Laurentia westwards across a narrow ocean to the Precordillera terrane seems not unlikely.”


It is significant that before the terrane hypothesis, trilobite experts (Palmer, 1972; Ross, 1975; Bordonaro and Banchig, 1995) attributed the paleobiogeography of Cambrian trilobites, including the occurrence of Laurentian trilobites in the Precordillera, to latitudinal or climatic belts and to dispersal by ocean currents. While some of these experts later invoked the terrane hypothesis, i.e. the Laurentian microcontinent model, it was not because they claimed that the alternative - dispersal by oceanic currents - was deficient.

Finally, Benedetto (2003) states that the inner-shelf Olenellid trilobite Realm in the Precordillera
cannot be explained by patterns of oceanic circulation, due to the fact that dispersion of benthic organisms across oceanic basins can only be achieved, at most, by a few eurytopic genera but never by the fauna as a whole. Of the eleven genera and subgenera of olenellids within the middle and lower upper parts of the Olenellus zone of Laurentia, only four occur in the Precordillera, generally in very low abundance and represented by endemic species. Although Salterella occurs in the Precordillera, the olenellid Wanneria and archaeocyathids that commonly occur with Salterella in Laurentia (Fritz and Yochelson, 1988) are absent.

Thus, the fauna of the Precordillera is a very small sample of the “Laurentian” fauna; a situation that, according to Benedetto (2003), is consistent with dispersal by oceanic circulation.


Conclusion
The distribution of trilobites in Middle and Upper Cambrian strata of the Precordillera follows the pattern established in the upper Lower Cambrian with the degree of affinities between faunas of the Precordillera and Laurentia varying with intensity of oceanic circulation, evolution of trilobite faunas and their dispersal abilities, evolution of facies within the carbonate platform of Cuyania, changes in sea level, and changes in the distance between and the relative paleolatitudes of Cuyania and Laurentia.
An objection to the parautochthonous model is the difference in trilobite faunas between Cuyania and East Gondwana, particularly Antarctica. Given that Antarctica was to the west of Cuyania in the Cambrian and that equatorial currents flow to the west, it is highly unlikely that redlichiid trilobites of East Gondwana would have dispersed eastwards to Cuyania.

Whether or not olenellid trilobites dispersed westward to Antarctica is difficult to determine because of their brief existence in Cuyania and the very limited and incomplete Cambrian fossil record for Antarctica (Palmer and Gatehouse, 1972; Palmer and Rowell, 1995). Some non-olenellid, benthic trilobites of Laurentian affinity did disperse to East Gondwana in the Early Cambrian, namely Bonnia and Kootenia, and the lower Upper Cambrian Aphelaspis Zone of Laurentia is recognized in the Minaret Fm of Antarctica by the occurrence of Laurentian species (Shergold and Webers, 1992).
At other times there may have been barriers to westward circulation and dispersal of “Laurentian” trilobites from Cuyania to Antarctica. Accordingly, the Cambrian trilobite record is compatible with the parautochthonous model for Cuyania, but there are some difficulties in reconciling it with the Laurentian microcontinent model because of the very limited abundance and diversity of Cambrian trilobites in the Precordillera relative to Laurentia.


Juxtaposition in the Mid Ordovician of Atlantic and Pacific province graptolite faunas by the Laurentian microcontinent model Although the paleobiogeography of benthic faunas is cited widely in econstructing the paleogeography of the Cuyania, the paleobiogeography of lanktonic graptolites is largely ignored. Is this because of an assumption that “shallow-water faunas provide an excellent biogeographic signal, whereas marginal faunas provide evidence of open ocean circulation but are not so diagnostic biogeographically” (Dalziel et al., 1996), or is it because the affinity of Mid Ordovician graptolite faunas in Cuyania are not easily reconciled with the Laurentian microcontinent model?

Graptolites are common and faunas are diverse in Middle and Upper Ordovician strata of the Precordillera (Maletz and Ortega, 1995). They represent the Pacific faunal province that was of global extent at low paleolatitude (Finney and Chen, 1990). In contrast, graptolites of the Atlantic faunal province, also of global extent but at high paleolatitude, occur in Lower and Middle Ordovician strata of the Famatinian belt and the Cordillera Oriental.

The sharp contrast between the two regions in faunas of the Mid Ordovician when provincialism was
at its greatest was noted by Finney and Chen (1990), Maletz and Ortega (1995) and Mitchell et al. (1997), among others. In the Laurentian microcontinent model, the Pacific province and Atlantic province faunas would have inhabited adjacent, but contiguous, areas of marine waters along the Gondwanan margin. Dalziel (1997) tried to attribute this juxtaposition to separation by oceanic currents, but in light of models of graptolite paleoecology (Finney and Berry, 1997), in particular the ability of graptolites to disperse globally, this distinction of globally distributed faunas in close geographic proximity could not have been maintained, a conclusion also favored by Maletz and Ortega (1995) and Mitchell et al. (1997). On the other hand, these graptolite distributions are compatible with, and readily explained by, the parautochthonous model. Through the Mid Ordovician, Cuyania was at relatively low paleolatitude within the realm of the Pacific graptolite faunal province, far removed from the Famatinian belt and the Cordillera Oriental, which were at substantially higher southern paleolatitude within the realm of the Atlantic province.

 

DISCUSSION: THE PARAUTOCHTHONOUS MODEL RECONSIDERED


All the evidence described above, including the occurrence of olenellid trilobites, is compatible with the
parautochthonous model (Aceñolaza et al., 2002; Finney et al., 2005b). Questions have been raised regarding two aspects of the model proposed by Finney et al. (2003, fig. 3; 2005b, fig. 6), namely the Cambrian location of Cuyania and its subsequent migration to its present position outboard of the Famatina belt. The Finney et al. (2005b) reconstruction (Fig. 5) is a schematic diagram based on the reconstruction of Gondwana by Hoffman (1991).

Located on the southern (present coordinates) margin of West Gondwana in the Cambrian, Cuyania would have been part of, or adjacent to, Mesoproterozoic and Neoproterozoic orogenic belts composing and bordering the Río de la Plata and Kalahari cratons - a location consistent with detrital zircon and paleobiogeographic evidence. Considering that Hoffman’s reconstruction may not portray accurately this continental margin, Cuyania is shown in a general location and as an oversized terrane to increase its visibility in Fig. 5. The constraints imposed by Hoffman’s reconstruction also require that Cuyania be shown migrating along a curved path from Cambrian to Devonian time and rotating in the process. However, given the proposal that it migrated along a transform fault, it is more likely that it followed a relatively straight path and did not rotate.


The width of the Iapetus Ocean is of considerable importance to the Laurentian microcontinent model and to the Finney et al. (2005b) version of the parautochthonous model (Fig. 5). The great width in the Laurentian microcontinent model (at least 2,000 kms) is based on interpretations of paleogeographic data that include an isolation stage for Cuyania, resulting from it being separated from both Laurentia and Gondwana by at least 1,000 kms in Early Ordovician time (Thomas and Astini, 1996). Here, the representation and interpretation of the paleobiogeographic data are challenged, and it is emonstrated that Laurentian faunal elements continued to migrate to Cuyania during the Early Ordovician and even
in the Mid and Late Ordovician. Thus, the parautochthonous model includes a relatively narrow Iapetus Ocean.
Evolving paleogeographic affinities of the benthic and pelagic faunas of Cuyania are attributed to its migration across paleolatitude initially during Cambrian to Mid Ordovician time as a part of Gondwana as the supercontinent itself migrated southwards (Fig. ) and from Mid- Ordovician time onwards by further relative southward migration along a transform fault(s).


The parautochthonous model is a hypothesis that appears to be most consistent with all available evidence. As with any hypothesis, there are still many aspects of the model to be further tested and checked and new evidence to be obtained and incorporated. No doubt, it will evolve further. Nevertheless, it has generated exciting, creative new research, new discoveries, and new ideas that would not have been possible if one is constrained by the Laurentian microcontinent model.

For example, Peralta (2005a, b) has realized and subsequently confirmed, that many of the large olistoliths composed of siliciclastic strata in the Los Sombreros and Rinconada formations were derived from the La Chilca, Los Espejos, Talacasto and Punta Negra formations of Silurian and early Devonian age, indicating that various parts of the Los Sombreros and Rinconada formations were deposited during early Devonian time. Peralta has interpreted the olistostromes to represent extensional pull-apart basins along major strike-slip faults. In addition, Peralta and Heredia (2005) have discovered that the Upper Ordovician Empozada Fm at San Isidro is overlain stratigraphically by a thick succession of shale and siltstone with olistoliths of shale that have Devonian land plant fossils, as well as abundant, huge olistoliths of Empozada Fm. The Empozada Fm stratigraphically overlies the Estancia San Isidro Fm that includes huge olistoliths composed largely of Middle and Upper Cambrian carbonate strata.

According to the Laurentian microcontinent model (Thomas and Astini, 2003), deposition of the carbonate olistoliths, those of both the Estancia San Isidro and Los Sombreros formations, occurred on the western continental slope and rise of the Cuyania microcontinent (Fig. ). However, the stratigraphic succession at San Isidro (Estancia San Isidro and Empozada formations) more likely accumulated in a strike-slip extensional basin because of 1) its great variety of sediment types representing a diversity of shallow and deep water depositional processes and 2) the Empozada
conglomerate bed composed of huge (1 m), rounded granitic boulders representing subaerial exposure and fluvial transport to the basin.

Now Peralta and Heredia (2005) have discovered that this extensional setting, possibly driven by movement along major strike-slip faults and operating from Mid to Late Ordovician time, was also active in Devonian time. Taking into account Peralta’s discoveries and re-interpretations of the Los Sombreros and Rinconada formations, it is apparent that huge strike-slip basins are present in the Precordillera, some of the most amazing examples in the world, and that they operated at least from Mid Ordovician to Devonian time.


If constrained by the Laurentian microcontinent model, they would not have been recognized.
Similarly, other discoveries have been made in the Cuyania terrane and the neighboring western Pampeanas ranges that are not clearly consistent with the Laurentian microcontinent model and perhaps also justify consideration of new models. These include: 1) The evidence of Rapela et al. (2005a), Galindo et al., (2004), and Baldo et al., (2005) that “the Western Sierras Pampeanas, which are considered part of the exotic Argentine Precordillera terrane of allegedly Laurentian derivation, could be autochthonous or parautochthonous to Gondwana.”
2) The suggestion that rocks between the Precordillera and the Famatina belt and previously considered part of Cuyania and of Laurentian origin are, instead, “a distinct crustal fragment of Gondwana or peri-Gondwana affinity…” (McClelland et al., 2005).
3) The description of Devonian age shear zones along the eastern margin of Cuyania (Roeske et al., 2005).
4) The discovery of Grenvillian age basement and overlying Neoproterozoic meta-sedimentary rocks in the Western Sierras Pampeanas (Casquet et al., 2005). It was obvious at the Gondwana 12 Conference in November 2005 (Pankhurst and Veiga, 2005) that there is much exciting research taking place on the rocks of the Cuyania terrane and the Sierras Pampeanas. It is a time for creative, innovative thinking and exciting new discoveries. This research should not be constrained in its starting assumptions, its methodological approach, or its data interpretation by any one model, whether it is the Laurentian microcontinent model or the parautochthonous model.


CONCLUSIONS
Although substantial, varied geological evidence is considered by many geologists to support the Laurentian microcontinent model of Thomas and Astini (1996), recently acquired U-Pb geochronology of detrital zircons from Cambrian and Ordovician sandstones in the Precordillera cannot be reasonably explained by a Laurentian affinity. Instead, these data are most consistent with a Gondwanan provenance and the parautochthonous model of Aceñolaza et al. (2002). Accordingly, it is incumbent that the many lines of evidence cited in support of the Laurentian model be re-examined critically for their consistency with the parautochthonous model. This assessment has revealed that:


1) potential basement rocks of Cuyania of Neoproterozoic to Early Cambrian age and Early Mesoproterozoic age are characteristic of Gondwana, rather than Laurentia;

2) Pb isotopic ratios of Grenvillian-age basement rocks are not only similar to those of Grenvillian base ment in Laurentia but also to those in other areas of West Gondwana;


3) the similarity of the Cambrian-Ordovician carbonate platform succession of Cuyania to that of Laurentia reflects similar paleolatitude and eustatic histories and not a direct connection;


4) the Middle-Upper Ordovician siliciclastic successions of Cuyania do not represent a peripheral foreland basin, but instead were deposited in strike-slip basins in a transform fault zone;


5) Middle Ordovician K-bentonites do not indicate that Cuyania was approaching the Famatina magmatic arc from the west (modern coordinates), but instead that it was located to the southeast;


6) the paleomagnetically determined Cambrian paleolatitude of Cuyania is consistent not only with the location of the Ouachita embayment of Laurentia but also with the southern margin of West Gondwana;


7) brachiopod and conodont faunas in lower Middle Ordovician strata of the Precordillera have many more genera in common with Laurentia than those in Lower Ordovician strata;


8) Cambrian trilobites faunas of Cuyania are of very limited abundance and diversity in comparison to correlative faunas of southeastern Laurentia; many species are endemic to Cuyania; olenellid trilobites considered to be restricted to Laurentia probably had the ability to disperse between paleoplates with similar environments; and


9) Mid Ordovician graptolites of the Precordillera on the one hand and of the Famatinian belt and Cordillera Oriental on the other belong to different oceanic provinces and likely did not live in close proximity.
Accordingly, the evidence used to support the Laurentian microcontinent model is found to be more consistent with the parautochthnous model, in which the Cuyania terrane was located on the southern margin of West Gondwana until the Mid Ordovician when it began migrating along a transform fault to its present position outboard of the Famatina magmatic belt in Devonian time.

arriba

 

PIE DE PALO

 

Sierra de Pie de Palo (provincia de San Juan)

La Sierra de Pie de Palo es un subsistema orográfico perteneciente a la faja occidental de las Sierras Pampeanas Noroccidentales. Alcanza una extensión de 71 km N-S y 35 km E-O. La altura media del cordón es de 3.000 msnm. Entre sus cerros se encuentra el Cerro Mogote Corralitos con sus 3.162 msnm el Cerro Las Pircas con 3.100 msnm.

La Sierra de Pie de Palo es uno de las más occidentales serranías de las Sierras Pampeanas. La sierra se eleva en el antepaís andino debido a la orogenia andina (Jordan et al, 1983; Ramos et al 2002). La sierra se encuentra entre los valles de Tulum y Bermejo en la provincia de San Juan.
Incluye en gran parte rocas metamorficas de medio a alto grado del Complejo Pie de Palo (Ramos y Vujovich, 2000). Rocas de bajo grado metamórfico pertenecen a la Secuencia Metasedimentaria Difunta Correa (Baldo et al., 1998) y el Grupo Caucete (Borrello, 1969) que estan expuestas en un sistema de fallas a lo largo del el flanco suroeste de la sierra. Granito intrusivo y pegmatitas están presentes en las zonas restringidas de la sierra.

Complejo Pie de Palo
Como fue definido originalmente, el Complejo Pie de Palo incluye un conjunto de esquisto, mármol, migmatita, gneis, leucogranito y rocas metaigneas máficas a ultramáficas (Stappenbeck , 1910; Schiller, 1912, Stieglitz, 1914, Dalla Salda y Varela, 1982 , 1984; Ramos y Vujovich , 2000) . Las unidades se pueden subdividir en varios unidades delimitadas por fallas de rumbo noreste. La unidad estructuralmente más baja consiste de rocas máficas, ultramáficas y rocas metamórficas en el flanco occidental de la sierra (Vujovich y Kay, 1998 ). En la porción noreste de la sierra, los mármoles cristalinos están estructuralmente intercaladas con gneis ricos en granate y biotita y anfibolitas ( Vujovich y Ramos , 2000 ) .
La edad Mesoproterozoico para el Complejo Pie de Palo fue originalmente asignada sobre la base de una edad discordante U/Pb 207Pb/206Pb de 1060 ± 20 Ma obtenida en circón de gneis en el área central de la sierra ( McDonough et al., 1993). Esta edad es consistente con edades Rb/Sr en roca total sobre metamórfica
y rocas ígneas en la porción central de la sierra (Varela y Dalla Salda, 1993; Rapela y Pankhurst, 1998). Estudios más recientes han arrojado mayores edades ígneas de ca. 1204 Ma, 1174 ± 43 Ma, y ca. 1169 Ma para pegmatita gabriode, leucogabros-diorita, y sills tonaliticos-granodiorita calco-alcalina, respectivamente (Vujovich et al. 2004). El granitoide (El Tigre granitoide) portador de granate y de dos micas emplazado en el Complejo Pie de Palo dio una edad U/Pb zircon es de 1104.8 ± 4.8 Ma (Morata et al., 2008).

 

Imágenes tomadas de M. Naipauer, G.I. Vujovich, C.A. Cingolani . W.C. McClelland, 2010. Detrital zircon analysis from the Neoproterozoic–Cambrian sedimentary cover (Cuyania terrane), Sierra de Pie de Palo, Argentina: Evidence of a rift and passive margin system? Journal of South American Earth Sciences 29: 306–326

T

R

 

Secuencia Metasedimentaria Difunta Correa
La secuencia Difunta Correa incluye facies de anfibolita, esquistos calcio-peliticos, cuarcitas, meta-arcosa, mármol y anfibolita expuestos a lo largo de las márgenes meridionales y orientales de la Sierra de Pie de Palo (Baldo et al., 1998). Edades U/Pb SHRIMP análisis sobre circones detríticos en para-anfibolitas intercaladas indican el máximo edad de deposición de 625 Ma (Rapela et al., 2005). Estudios isotópicos 87Sr/86Sr, C y O sobre las unidades carbonaticas Neoproterozoico sugieren una edad de depositación de la secuencia (ca. 720-580 Ma; Galindo et al. 2004). Edades de circones detríticos van desde 1032 a 1224 Ma con edades metamórficas de 460 Ma (Casquet et al. 2001). Un ortogneis milonítico dentro de la Secuencia metasedimentaria Difunta Correa en la porción suroeste de la sierra es similar a los magmatismo de tipo A de intraplaca y da la edad de cristalización U/Pb de 774 ± 6 Ma (Baldo et al., 2006).

 

Grupo Caucete

El Grupo Caucete incluye unidades de bajo grado metamorfico formado por rocas calcáreas y cuarcita en el lado occidental de la Sierra de Pie de Palo (Borrello, 1963, 1969) que están en contacto de falla con el basamento metamórfico mas antiguo (Schiller, 1912). La sección se compone de las formaciones El Quemado, La Paz, El Desecho y Angacos (Borrello, 1969; modificado por Vujovic, 2003). Dos componentes principales son reconocidos en el Grupo Caucete: uno de composición silicilastica (formaciones El Quemado y La Paz); la otra composición son calizas (formaciones El Desecho y Angacos). Las últimas dos formaciones se han correlacionado con unidades cámbricos basales de la sucesión de Precordillera oriental (e.g. formaciones Cerro Totora y La Laja, van Staal et al., 2002). Sin embargo, el origen de las formaciones El Quemado y La Paz y la correlación con las unidades no metamorfizadas de la Precordillera es difícil. La deformación penetrativa y y las facies de esquistos verdes (Ramos y Vujovic, 2000) del Grupo Caucete ha emnascarado las relaciones estratigráficas de la secuencia.

 

O

Las relaciones estructurales sugieren que la formación El Quemado fue imbricada con el Complejo Pie de Palo y posteriormente ambos fueron emplazados hacia el oeste sobre las formacion Angacos y El Desecho del Grupo Caucete (van Staal et al., 2002). La edad de la secuencia deposicional es confusa debido a la falta de fósiles diagnósticos, inciertas relaciones estratigráficas y deformación penetrativa. Una edad paleozoica inferior fue asignada basado en correlación con las unidades de la Precordillera (Schiller, 1912; Groeber, 1948). El Grupo Caucete se ha correlacionado específicamente con el Cámbrico tardío y Ordovícico temprano de calcareos de la Precordillera sobre la base de estudios isotópicos de carbono y oxígeno (Linares et al., 1982; Abbruzzi, 1994; Sial et al., 2001). Más recientemente, Galindo et al (2004) considera el Grupo Caucete como equivalente a las plataformas carbonáticas cámbricas de la Precordillera basada en datos 87Sr/86Sr, d13C y dO. Naipauer et al (2005a) estableció una edad Cámbrica (510 Ma) para la formación Angacos y lo correlaciona con los miembros inferiores de La formación la Laja en la Precordillera. Finalmente, posible ichnofossils descrito por Bordonaro et al (1992) pueden indicar la equivalencia con la Formacion Puncoviscana (Neoproterozoico– Cambrico inferior) en el noroeste de Argentina.

Formación El Quemado

Incluye unidades siliciclásticas expuestas en el flanco occidental de la Sierra de Pie de Palo que originalmente fueron mencionados como la Cuarcita El Quemado (Borrello, 1963, 1969). La sección está compuesta por metaareniscas ricas en cuarzo y feldespato, esquistos cuarzo–micáceos, areniscas de cuarzo pardo y sus equivalentes cataclasticos. La sección está bien expuesta de la Quebrada Agua del Conejo al norte (Ramos y Vujovic, 2000) a la Quebrada La Petaca al sur donde las rocas siliciclásticas son sustituidas por rocas calcáreas. El espesor original de las unidades dentro de la sección se desconoce debido al alto grado de plegamiento y fallamiento de la secuencia. Las metareniscas son predominantemente verde y amarillo, fino a grano medio con granos individuales de feldespato de hasta unos milimetros con cuarzo, feldespato, muscovita y biotita. Una laminación secundaria es común en la mayoría de las unidades y esta formada por deformación y recristalización. El protolito se interpreta como una arenisca inmadura basado en el alto contenido de mica y feldespato.

 

J

e

 

Formación de La Paz

Consiste de esquistos cuarzo-micaceos compuestos por moscovita, granate y albita variablemente milonitizados (Vujovic, 2003). Los mejores afloramientos se encuentran entre la Quebrada La Paz y la Quebrada Las Pirquitas, y estan ampliamente extendidos a la zona norte de la Sierra, en ña zona de Lomas Bayas. Entre la Quebrada El Molle y la Quebrada El Quemado, los estratos de la formación de La Paz se intercalan con capas de metaareniscas de la formación El Quemado, lo que sugiere un contacto transicional entre ambas unidades. Se observa un espesor máximo estructural de 250 m para la formación La Paz en la zona de El Quemado. Las capas individuales varían de unos pocos centímetros a 3 m de espesor. La unidad es comúnmente oscura en color, con color verdosos y grisáceos tonos dominantes: El tamaño de grano abarca desde muy fino a medio, con porpfiroclastos de albital que llegan a 2 mm. La formación La Paz difiere de la Fm. El Quemado Formation por la presencia de grnate, albita, epidoto, y filisilicatos en los esquistos micáceos. Se ha inferido la existencia de un protolito compuesto de pelitas volcanicásticas y areniscas. (Van Staal et al., 2002; Vujovich, 2003).

Formación El Desecho

Fue descrita originalmente como Formación Puntilla Blanca (van Staal et al., 2002), pero referido más adelante como la Formación El Desecho El (Vujovic, 2003), debido al hecho de que nombre anterior fue utilizada por Borrello (1969) para otra unidad del Grupo Caucete. La unidad incluye carbonatos y rocas dolomíticas, mármoles, esquistos calcáreos, metapelitas, metaareniscas calcáreas y metaconglomerados subordinados. El espesor es variable, desde unos pocos metros hasta 40 m en la Quebrada El Desecho. La formación varía de rojo, amarillo, negro y verde en color, y es un marcador útil que ayuda a definir las estructuras. Cuarcitas de grano fino a medio y metaareniscas calcáreas expuestos en la zona de La Olla, entre las Quebradas La Cruz y Pecan alcanza varios metros de espesor y fue seleccionada para el estudio de procedencia. Hay metaconglomerados con clastos redondeados que varían en tamaño de 5 a 30 cm de diámetro en el lado suroeste de la Lomas Bayas. Los clastos son principalmente graníticos y ocurren en una matriz calcárea.

Formación Angacos (Caliza Angacos: Borrello, 1969)

Se compone de caliza penetrativamente deformada, esquisto calcáreo y mármol. Los principales afloramientos se encuentran en las Quebradas El Gato y La Petaca pero se extienden hacia el sur donde Dalla Salda y Varela (1984) mencionaron que los mármoles y esquistos calcáreos son hasta 200 m de espesor en la Quebrada Ancha de la Punilla. Esquistos y esquistos calcáreos gris a negros, de grano fino se encuentran en la base y mármoles masivos ocurren en la parte superior de la Formación Angacos (Ramos y Vujovic, 2000). Las mineralogía va de calcita a dolomita. En la zona suroeste de las Quebradas El Gato, La Petaca y La Lichona las rocas muestran las laminaciones de 2 a 30 mm de espesor que se corresponden con la separación en capas de calcita - dolomita . Bajo el microscopio, cuarzo y capas orgánicas son reconocidas. Estas estructuras primarias fueron interpretadas como ritmitas por Naipauer et al., (2005c) dentro de una secuencia donde alternan calizas y areniscas calcáreas. (Ramos and Vujovich, 2000).

 

 

Ñ

w

Detrital zircon U–Pb geochronology
El Quemado Formation
A total of 121 zircons from sample QLPcz1 were analyzed, but results from 57 grains were rejected due to discordance of >20%. Concordant zircon ages define four main intervals: 1169– 1040 Ma (36%), 1289–1187 Ma (31%), 1350–1300 Ma (22%) and 1434–1391 Ma (8%). There are two single ages at 506 Ma and 1540 Ma. Zircon grains that produce the main peak at ca. 1150 Ma are long prismatic crystals with oscillatory zoning typical of plutonic and/or volcanic rocks. The peak at ca. 1220 Ma corresponds to zircons with rounded shapes and complex metamorphic textures. Peaks at ca. 1310 Ma and 1400 Ma are from zircons with oscillatory zoning characteristic of magmatic origin.


Zircon from sample QLPcz2 provided 135 analyses, 36 of which were rejected due to discordance. The remaining 99 grains define three dominant age ranges of 697–532 Ma (peak at ca. 550 Ma; 12%), 1228–1042 Ma (peak at ca. 1110 Ma; 43%), and 1492– 1273 Ma (peak at ca. 1360 Ma; 40%). Two grains give ages of ca. 1553 Ma and 1697 Ma. The youngest zircon group defines two peaks at ca. 550 Ma represented by 7% of the population (568–532 Ma) and ca. 640 Ma defined by 5% of the population (697–590 Ma).

The younger grains are typically large prismatic to moderately rounded crystals with fine oscillatory zoning characteristic of a magmatic origin. The peak at ca. 1110 Ma is represented by two groups: oscillatory zoned prismatic grains of probable volcanic origin that yield an age of ca. 1090 Ma and round grains of probable metamorphic origin with complex internal textures defined by variable luminescence, recrystallized rims, and ages as old as ca. 1200 Ma.The peak at ca. 1360 Ma is characterized by large prismatic zircons with oscillatory zoning of magmatic origin.


Sample QPir3 was collected from the same unit as the previous sample. Of the total 125 detrital zircons analyzed, 16 were rejected due to discordance. Concordant zircons define three major intervals:
1166–1070 Ma (46%), 1262–1188 Ma (40%), 1340–1274 Ma (11%) and 1476–1439 Ma (3%) with the most representative peaks at ca. 1070 and 1120 Ma. In addition, there are elongate prismatic grains with igneous zoning textures suggestive of volcanic sources define a peak at ca. 1070 Ma. The remaining grains from this sample display metamorphic textures. Subequant prismatic grains with metamorphic textures give peaks at ca. 1220 Ma and 1250 Ma. An older peak at ca. 1340 Ma is defined by metamorphic zircons and a few oscillatory zoned crystals of probable magmatic origin. Single zircon ages appear at ca. 1476 Ma, 1464 Ma, and 1439 Ma.

 

 

 

w

La Paz Formation
The zircon grains analyzed in the M4 sample were 125; 13 grains of discordant age (more than 20%) were rejected. The 112 grains with concordant age (<20%) fall in three main age brackets:
977–1166 Ma (77%), 1275–1171 Ma (15%), and 1490–1303 Ma (7%). The dominant peaks at ca. 1040, 1070 and 1145 Ma are defined by round zircons of probable metamorphic origin and subordinate prismatic grains with magmatic textures that give ages of ca. 1040 Ma. Small prismatic, idiomorphic to subidiomorphic, oscillatory zoned igneous grains define the main peak at 1145 Ma as well. Zircons with ages between 1171 Ma and 1275 Ma are of probable metamorphic origin whereas a peak at ca. 1360 Ma is defined by magmatic zircons. A small grain yielded a single age of 1925 Ma.


In sample M5, 130 detrital zircons were analyzed and 29 discarded. Concordant ages define five intervals: 1062–961 Ma (29%), 1118–1070 Ma (23%), 1171–1121 Ma (22%), 1257–1191 Ma (15%) and 1276–1357 (7%). The younger group is dominated by metamorphic zircons. However, prismatic oscillatory zoned zircons that define a peak at ca. 1035 Ma are likely of volcanic origin. Peaks at ca. 1080 Ma, 1135 Ma and 1220 Ma are formed by rounded and subrounded zircon grains, mostly of metamorphic origin. Finally, peaks at ca. 1300, 1360 and 1470 Ma are defined by large oscillatory zoned grains of igneous origin.

 

e

 

El Desecho Formation
A total of 118 zircon grains were analyzed from sample M8 and 16 were rejected due to discordance. The age produce six age intervals: 531–617 Ma (3%), 1186–1054 Ma (44%), 1293–1203 Ma (19%), 1460–1326 Ma (25%) 1527–1493 Ma (3%) and 1677– 1574 Ma (7%). The peak at ca. 550 Ma is defined by prismatic, subrounded oscillatory zoned grains with of probable igneous origin. The largest peak at ca. 1120 Ma is due to subrounded prismatic grains interpreted to be of metamorphic origin. Grains that yield a minor peak at ca. 1070 Ma are subrounded elongate prismatic grains with oscillatory zoning indicating magmatic origin. The peaks at ca. 1240 Ma and 1380 Ma are defined prismatic oscillatory zoned igneous zircons.
Finally, the peaks at ca. 1450 Ma and 1600 Ma consist of low luminescence grains of probable metamorphic origin.

Angacos Formation
Sample QLli1 provided 142 analyses, 33 of which are discordant. The remaining 109 concordant analyses are distributed in two major intervals: 1148–1050 Ma (peak at ca. 1114 Ma, 35%) and 1471–1313 Ma (dominant peaks at ca. 1373 Ma, 1400 Ma and 1450 Ma; 58%). Isolated ages are present in the interval 1312 to 1149 Ma (7%) as well as in the Paleoproterozoic.
Grains contributing to the peak at ca. 1114 Ma have complex internal textures characteristic of a metamorphic origin, but some grains with igneous textures are observed. The main peak at ca. 1373 Ma is defined by zircons with igneous textures with a minor proportion of grains displaying complex textures of probable metamorphic origin.

 

t

Principal populations and age components
The morphological and CL analysis of detrital zircons allowed definition of two populations composed of zircon grains with prismatic habit, oscillatory zoning, cores and inclusions but with different sizes. These features indicate a source area dominated by plutonic igneous rocks. Moreover, the presence of zircons with long prismatic habit indicates a probable third source with input from volcanic rocks. Finally, a population of rounded zircons with complex internal zoning probably indicates origin from metasedimentary rocks that were products of various sedimentary cycles.
These variations taken together with U/Pb (LA-ICP-MS) age determinations were used to characterize the main sources of sediment input for the Caucete Group.
Two Mesoproterozoic sources were identified. The oldest source with early Mesoproterozoic ages (ca. 1450– 1300 Ma) includes plutonic and volcanic rocks. A second source has metamorphic and subordinate igneous rocks of late Mesoproterozoic age (1300–1000 Ma; Grenvillian). Other subordinate zircons
with Neoproterozoic and Paleoproterozoic ages indicate different sources.


The early Mesoproterozoic source furnished igneous zircons with three main frequency peaks: ca. 1370, 1360 and 1310 Ma.
These zircons are present in all of the samples analyzed from the Caucete Group. It is the second dominant group in the El Quemado Formation, representing from 11% to 40% of the total population In contrast, these ages are subordinate in the La Paz Formation with peaks at ca. 1360 Ma (7%). The ca. 1380 Ma peak in the El Desecho Formation represents 25% of the total population, whereas the dominant ca. 1373 Ma peak in the Angacos Formation is defined by 58% of the analyzed grains.

This difference is interpreted to reflect a major change in source area for Caucete Group sediments. The late Mesoproterozoic metamorphic zircons are distributed in a range between 1293 and 997 Ma. This group is the most statistically significant in both the El Quemado Formation with 43–86% of the population and the La Paz Formation (89–92%).
This observation indicates that metamorphic basement of late Mesoproterozoic age was a constant source area during deposition of the La Paz and El Quemado Formations. The Mesoproterozoic‘‘Grenvillian” population is represented by 63% of the grains from the El Desecho Formation and decreases to 42% in the Angacos Formation.
This trend indicates that the Mesoproterozoic source area decreases in importance in the upper units of the Caucete Group.
A different Mesoproterozoic source is represented by a population of volcanic zircons with ages between ca. 1070 and 1030 Ma. This ‘‘Grenvillian” population is abundant in the La Paz Formation, as well as the El Quemado and El Desecho Formations. Magmatism in the source region of the Neoproterozoic igneous grains was probably contemporaneous with, or slightly older than, deposition of the Caucete Group.

Maximum depositional age
Detrital zircon data can provide limits on the maximum depositional age of sedimentary rocks as long as the data set is representative and statistically robust and there is sufficient knowledge of the geological setting and possible source regions (Andersen, 2005). This approach is complicated in metasedimentary rocks such as the Caucete Group due to the potential presence of metamorphic overgrowths of grains.


The youngest zircons in the El Quemado Formation exhibit a range of 238U/206Pb ages between ca. 492 and 506 Ma. These ages are problematic because they are close to the 490–450 Ma interval, the time of the Famatinian orogeny. Similar ages obtained in the Sierra de Pie de Palo have been interpreted as results of the Famatinian orogenic event (Ramos et al., 1998; Casquet et al., 2001; Vujovich et al., 2004). Accordingly, the dominant peak of 207Pb/206Pb ages at ca. 550 Ma defined by 10% of total zircon grains
analyzed from sample QLPcz2 is interpreted as the best estimate of the maximum depositional age of El Quemado Formation. This age may in fact be too young due to possible Pb loss due to Famatinian metamorphism. Several early Cambrian (ca. 531 Ma) zircons are present in the El Desecho Formation as well.

By comparison with the ca. 550 population of the El Quemado Formation, a similar maximum age of deposition is interpreted for the El Desecho Formation.

 

Gondwanan sources
Mesoproterozoic detrital ages of the Caucete Group can be compared with Gondwana basement ages, especially those of the Amazonian craton. Ages betwween ca. 1450 and 1300 Ma are widely represented in the geochronological province of Rondonia – San Ignacio in the southwestern Amazonian craton, where gneisses, migmatites and granulites ca. 1550 to 1300 Ma are exposed (Tassinari et al., 2000).

Moreover, ages of the ‘‘Grenvillian” zircons of the Caucete Group are comparable to those of the Sunsás belt, located in the southwestern Amazonian craton (Tassinari et al., 2000; Santos et al., 2008). Thus Gondwana is a possible source for Mesoproterozoic zircons of the Caucete Group.
However, significant differences exist between the Caucete Group and the Neoproterozoic–Cambrian metasedimentary rocks on the Gondwana margin. Gondwanan units in northern and central Argentina such as the Puncoviscana Formation and other units in the Sierras Pampeanas are characterized by an absence of ca. 1450–1300 Ma detrital zircon ages (see Rapela et al., 1998; Sims et al., 1998; Schwartz and Gromet, 2004; Adams et al., 2006; Escayola et al., 2007; Collo et al., 2009).

Similarly, grains of this age are not common in the Neoproterozoic–lower Paleozoic cover of the Rio de la Plata craton either (Rapela et al., 2007; Gaucher et al., 2008). Thus, the Gondwana margin is not a likely source region for the Caucete Group.

Laurentian sources
There are numerous p
otential source areas in Laurentia for detrital zircons of the Caucete Group. In particular, we examine sources close to Ouachita embayment: Yavapai and Mazatzal provinces (south-central to southwestern Laurentia), Granite-Rhyolite province (central and southern Laurentia), and the Grenville province exposed in the southern Appalachian region (eastern Ouachita Basin) and Llano Uplift-Van Horn-Franklin Mountain areas in the western Ouachita Basin.


Sources for older grains are wide spread in Laurentia. Paleoproterozoic metamorphic detrital zircons of the Caucete Group are compatible with the ca. 1800–1600 Ma Yavapai-Mazatzal province of south-central to southwestern Laurentia. Paleoproterozoic detrital ages observed in the Cambrian Cerro Totora Formation from the Precordillera have been attributed to a Laurentian source as well (Thomas et al., 2004). Early Mesoproterozoic ages from plutonic and volcanic detrital zircons of the Caucete Group coincide with the ages of the Granite-Rhyolite province in southern and central Laurentia (Muehlberger et al., 1967; Van Schmus et al., 1987).


The Provenance of late Mesoproterozoic or ‘‘Grenvilian” detrital zircons in the Caucete Group is straight forward since Grenvillian basement is well documented along the southern Laurentian margin.
The Llano Uplift (central Texas) and Van Horn-Franklin Mountain region (west Texas) contain exposures with ages of ca. 1320–1000 Ma (Roback, 1996; Mosher, 1998). Zircons with ca. 1300 Ma of the El Quemado, La Paz and El Desecho Formation are consistent with ca. 1330–1270 Ma magmatic rocks in the Llano Uplift (Roback, 1996). Igneous zircon ages ca. 1115, 1070 and 1030 Ma of the Caucete Group are compatible with ca. 1135–1070 Ma granite plutons in central Texas (Walker, 1992; Reed et al., 1995). The most representative outcrops of the southern Appalachian basement are in the Blue Ridge province. The ca. 1220–1160 Ma Caucete Group detrital zircon ages are comparable to the ca. 1190 Ma Grenvillian basement exposures in the southern Appalachians (Carrigan et al., 2003). Igneous zircon ages of ca. 1115, 1070 and 1030 Ma are compatible with the southern Appalachian basement as well (Bickford et al., 2000; Carrigan et al., 2003; Tollo et al., 2006).

Finally, the Neoproterozoic igneous zircons found in the El Quemado Formation (ca. 630–550 Ma) are interpreted to reflect magmatic activity in the source area. Rifting occurred along the southern and eastern Laurentian margins in Neoproterozoic to Cambrian times, in response to the opening of Iapetus and leading to terrane separation (Cawood et al., 2001; Tollo et al., 2004). The Neoproterozoic zircons from the El Quemado Formation are interpreted to have been derived source areas affected by rift-related magmatic activity at ca. 620–550 Ma. The ca. 531 Ma zircon ages of the El Desecho Formation may relate to a final magmatic stage (ca. 540–535 Ma) associated with the rift of the Cuyania terrane (Cawood et al., 2001).

Cuyania basement source
The igneous zircon populations with ages of ca. 1115 Ma, 1070 Ma and 1030 Ma in the Caucete Group, are broadly similar to those of the Mesoproterozoic Cuyania basement (1060 ± 20 Ma, McDonough et al., 1993; 1021 ± 12 Ma, Rb/Sr isochron, Pankhurst and Rapela, 1998; 1105 Ma, U/Pb SHRIMP age, Morata et al., 2008). Caucete Group detrital zircon ages of ca. 1220 and 1150 Ma clearly match ages in the crystalline basement as well (Vujovich et al., 2004).


The U/Pb detrital zircon peaks at ca. 1160–1150 Ma and 1080– 1050 Ma in the Neoproterozoic Difunta Correa Unit (Vujovich et al., 2004) are equivalent to peaks observed in the Caucete Group (ca. 1150 Ma, 1070 Ma and 1030 Ma). Zircon ages between 1100 and 1050 Ma from a para-amphibolite within the Difunta Correa Unit (Rapela et al., 2005) are similar to those in the Caucete Group as well. Finally, ca. 1220 Ma Caucete Group zircon ages match the age of the Las Matras pluton (1244 ± 42 Ma, Sato et al., 2004) and are similar to orthogneiss ages from the San Rafael block (1205 ± 1 and 1204 ± 2 Ma, Thomas et al., 2000) and other granitic rocks of the Cerro La Ventana Formation (1214.7 ± 6.5 Ma, Cingolani et al., 2005).


Mesoproterozoic ages are observed in the northern Cuyania terrane. The Juchi Orthogneiss of the Sierra de Umango gives an age of 1108 ± 13 Ma (Varela et al., 2003). Granitic mylonites and ultramylonites in the Jagüé area are similar in age (1118 ± 17 Ma, Martina et al., 2005). These crystallization ages are consistent with the ca. 1115 Ma igneous zircons in the Caucete Group. Moreover, detrital igneous zircon ages of ca. 1100–1080 Ma (Casquet et al., 2008) and metamorphic zircon ages of ca. 1220 Ma (Casquet et al., 2006)
observed in the Sierra de Maz are similar to detrital zircon ages from the Caucete Group.
Crustal xenoliths in Miocene subvolcanic rocks in the Argentine Precordillera (Leveratto, 1968) show the existence of metamorphic basement beneath the Cambro–Ordovician carbonate platform. Kay et al. (1996) compared the Pb isotope signature from those xenoliths of the Cuyania basement with rocks of Sierra Pie de Palo and of the Llano Uplift (Texas) in Laurentia.

The felsic and mafic xenoliths yielded U/Pb zircon ages about 1099 ± 3, 1102 ± 6, 1096 ± 3 Ma (Kay et al., 1996) comparable with ages between 1115 and 1070 Ma found in the Caucete Group and in the Grenvillian rocks of Texas In summary, detrital zircon ages from the Caucete Group are consistent with ages observed in Laurentia. Mesoproterozoic and older Laurentian sources are well represented in the Yavapai- Mazatzal province, Granite-Rhyolite province, and Grenville province.
Neoproterozoic to Cambrian source regions include areas along the southern and eastern Laurentian margin affected by rift-related magmatic activity. The Laurentian basement of Cuyania itself may have contributed detritus as well (e.g. Sierra de Pie de Palo, Precordilleran basement, and Sierras de Umango and Maz).

Stratigraphic relationships and depositional associations
Deposition of the Caucete Group occurred from the late Neoproterozoic to early Cambrian (ca. 550–515 Ma) as constrained by U/ Pb detrital zircon ages reported herein and by previous isotopic studies (Galindo et al., 2004; Naipauer et al., 2005a,b).


Field relationships, petrographic and geochronologic data show different sedimentary protoliths and tectonic settings for the metasedimentary rocks of the Caucete Group. The El Quemado Formation is attributed to a rift setting on the basis of the sedimentary protolith and the igneous zircon source that is close in age to the time of deposition. A similar tectonic setting has been proposed for the southern and eastern margins of Laurentia at approximately 550 Ma. The La Paz Formation, consisting of volcanogenic pelites interbedded with fine sandstones, was deposited in a more distal environment than that of the El Quemado Formation. A similar tectonic setting is inferred for the La Paz Formation on the basis of the interbedded and transitional boundary between the two units (Fig. 13). The El Desecho Formation is a heterogeneous sequence of pelites, calcareous and dolomitic marbles, calcareous sandstone, and subordinate conglomerates. This unit was correlated with the Cerro Totora Formation of the Precordillera and both are interpreted to reflect rift sedimentation (Fig. 13). Finally, the more homogeneous sedimentary protolith of Angacos Formation, composed mainly of limestone and marbles, is similar to the La Laja Formation (El Estero and Soldano Members Members) of the Precordillera (Fig. 13). The tectonic setting of sedimentation for these rocks inferred as a passive margin developed on the Cuyania terrane.


Accordingly, two events of rift-stage sedimentation are distinguished. The first is defined by the maximum depositional age of the El Quemado Formation (ca. 550 Ma) which interfingers with the La Paz Formation. The second stage is represented by the El Desecho Formation with a maximum depositional age of ca. 531 Ma. The late Mesoproterozoic (Grenvillian) detrital zircons in the El Quemado, La Paz and El Desecho Formations were probably
derived from a nearby positive feature such as a rift bulge. Older detrital components were likely sourced by more distal areas in the interior of the Laurentia craton such as the Granite-Rhyolite province and Yavapai-Mazatzal area.


The Angacos Formation is interpreted as a platform carbonate sequence developed at ca. 520 and 515 Ma on the passive margin of the Cuyania terrane (Fig. 13). The provenance pattern for this unit is different from those for the El Quemado, La Paz and El Desecho Formations. Neoproterozoic–Cambrian zircons are absent and late Mesoproterozoic zircon ages are subordinate, indicating that these sources were partially covered by the carbonate platform. The most important sediment source was the Granite-Rhyolite province (Fig. 13). The Angacos Formation is interpreted to have
developed upon Grenvillian basement and older units of the Caucete Group, only receiving more distally derived detrital components.

Tectonic model for the rift – passive margin transition

The tectonic evolution of Cuyania composite terrane that is most widely accepted initiates with rifting from the Ouachita embayment of Laurentia in the early Cambrian, development of a passive margin in latest early Cambrian to lower Ordovician, and collision with the Gondwana margin in the middle – late Ordovician
(e.g., Astini et al., 1995; Thomas and Astini, 1996; Ramos, 2004). Rift sedimentation in the Precordillera region is only recorded by the Cerro Totora Formation (Astini and Vaccari, 1996); strontium data from this unit are consistent with continental rifting during early Cambrian (Thomas et al., 2001).

q

The transition from the synrift deposition of the Cerro Totora Formation to the passive margin deposition is documented by the late Early Cambrian Los Hornos and La Laja Formations (Astini et al., 1996). The detrital zircon record of the Caucete Group is generally consistent with the classical rift and passive margin stages of Cuyania terrane. However, the El Quemado and La Paz Formations record older stages of rifting not observed elsewhere in the Precordillera region. This observation indicates initial separation of Cuyania
much earlier in the end of the Neoproterozoic (ca. 575–550 Ma), a view consistent with the age of rift-related magmatism in the western Precordillera (Davis et al., 2000). Also, the detrital zircons have demonstrated that the Cuyania basement have a portion of the Granite-Rhyolite province in the northern sector (Fig. 4). The major stages that were distinguished in the tectonic–stratigraphic evolution in the Caucete Group were as following.

Rift stage I (ca. 550 Ma)
Initial rifting in the Ouachita embayment (Fig. 15a) likely started during a second extensional pulse between 620 and 550 Ma along the Laurentia margin (Tollo et al., 2004) that resulted in opening of Iapetus and separation of Laurentia from Gondwana at ca. 570 Ma (Cawood et al., 2001). Deposition of the Caucete Group within the Ouachita basin is inferred to start prior to final separation of Cuyania from Laurentia. The northern portion of the Cuyania terrane is interpreted to be part of the Granite-Rhyolite province (Thomas et al., 2000).

The El Quemado and La Paz Formations reflect synrift sedimentation at ca. 545 Ma (Fig. 15b), prior to deposition of the Cerro Totora Formation. Provenance ages demonstrate a clear connection between the early sediments of the Caucete Group and the southern Laurentian margin (e.g., Granite- Rhyolite and Grenville provinces). The presence of igneous detrital zircons with ages close to 550 Ma allows the interpretation of riftrelated magmatic rocks in the source area (Fig. 15b).
9.2. Rift stage II (ca. 531 Ma) The southern Cuyania terrane is inferred to largely be underlain by Grenvillian basement (Fig. 16, see location of B–B0 in Fig. 14).

The second sedimentary stage of the Caucete Group is recorded by synrift deposits of the El Desecho Formation with a maximum depositional age of ca. 531 Ma. This unit probably developed during transition to a thermal subsidence stage. El Desecho Formation is correlative with the synrift deposits of Cerro Totora Formation exposed in the northern Precordillera (Astini et al., 1995) (Fig. 16a), but the units accumulated in different depocentres as indicated by the differences in their detrital zircon signature.
U/Pb detrital zircon ages of ca. 1600, 1380, 1240, 1120, 1070 and 550 Ma from the El Desecho Formation are similar to those from the El Quemado Formation.

Thus the basin configuration and source areas are interpreted to be similar and both record continued connection with Laurentia. Consistent with this hypothesis, detrital zircons from the Cerro Totora Formation define populations of ca. 1490–1300 Ma and ca. 1890–1640 Ma that are characteristic of the Laurentian craton interior (Thomas et al., 2004).

Passive margin stage (ca. 520 Ma)
The initial separation of the Cuyania terrane and generation of a passive margin occurred at approximately 520 Ma (Fig. 16b). The Angacos Formation was deposited during a third stage in Caucete Group sedimentation. The unit is comparable in age and depositional setting to the limestones of the Los Hornos and La Laja Formations of the Precordillera (Fig. 16b). The occurrence of siliciclastic interbeds in the carbonates indicates a proximal uplifted area in the Cuyania terrane. In addition, lower members of the La Laja Formation contain many quartz sandstone interbeds (Pereyra, 1987; Finney et al., 2005). These siliciclastic interlayers reflect deposition in a basin near an exhumed basement area (Gómez and Astini, 2006).

The zircon population at ca. 1118 Ma from the Angacos Formation shows that Grenvillian basement of Cuyania was partially exposed at the time of deposition. However, this source was subordinate to the early Mesoproterozoic (ca. 1450–1300 Ma) source area. The Grenvillian basement was partially overlapped by deposits of the El Quemado, La Paz and El Desecho-Cerro Totora Formations, and by subsequent passive margin sediments. Detrital zircons with ages around 1360 Ma are dominant in the early to middle Cambrian units of the Angacos Formation. Thomas et al. (2000), suggest that the northwest corner of the Precordillera consists of a small portion of the Granite-Rhyolite province, as shown in the sections of northern Cuyania (Figs. 14 and 15b).

This alternative is supported by data from basement clasts in conglomerate olistoliths in the western Precordillera, where zircon ages of ca. 1370 ± 2 and 1367 ± 5 Ma (Thomas et al., 2000) are similar to the ages from the Angacos Formation. Regardless of the basement correlations, deposition of the Angacos Formation likely records the post-rift phase as Cuyania started to drift through the Iapetus (Thomas and Astini, 1996). Since middle Cambrian to lower Ordovician ophiolitic rocks are not observed in Laurentia or Cuyania terrane (Dalziel, 1997; Keller, 1999), the presence of oceanic crust in the western sector of the Cuyania terrane is queried in Fig. 16b during the middle Cambrian to lower Ordovician. Nevertheless, the Angacos Formation and coeval units of the Precordillera were likely removed from Laurentian detrital sources.

Y

 

 

BLOQUE DE SAN RAFAEL, LAS MATRAS Y CHADI LEUVU

Afloramientos grenvilianos mendoza.JPG (116488 bytes)

geologia area ponon trewhue.JPG (609251 bytes)

 

Datacion La Ventana.JPG (63104 bytes)

 

Afloramientos del basamento Grenvilliano y Calizas en Mendoza Geología de detalle del sector de Ponón Trehue Datación Fm Cerro La Ventana

 

 

BIBLIOGRAFIA

Abruzzi, J.M., Kay. S.M.. Bickford, M.E., 1993. Implications for the nature of the Precordilleran basement from the geochemistry and age of Precambrian xenoliths in Miocene volcanic rocks, San juan province. XII Congreso Geológico Argentino y II Congreso de Exploración de Hidrocarburos, Actas, vol. 3, pp. 331—339.

Aceñolaza, F.G., Miller, H., Toselli, A.J., 2002. Proterozoic-Early Paleozoic evolution in western South America - a discussion. Tectonophysics, 354, 121-137.

Aceñolaza, F.G., Toselli, A.J., 1988. El Sistema de Famatina, Argentina: su interpretación como orógeno de margen continental activo. V Congreso Geologico Chileno, Santiago, Argentina, Actas, 1, A55-A67.

Albanesi, G.L., Barnes, C.R., 2000. Subspeciation within a punctuated equilibrium evolutionary event: Phylogenetic history of the Lower-Middle Ordovician Paroistodus originalis- P. horridus complex (Conodonta). Journal of Paleontology, 74, 492-502.

Albanesi, G.L., Bergström, S.M., 2002. Conodont paleobiogeographical co-evolution of the Argentine Precordillera and the Marathon area, Texas in the Ordovician Period. Eight International
Conodont Symposium held in Europe (ECOS VIII), Toulouse-Albi, Abstracts, Strata, Serie 1, p. 12.

Albanesi, G.L., Bergström, S.M., in press. Statistical analysis of Early-Middle Ordovician conodont paleobiogeography with special regard to the Argentine Precordillera. Geological Society of America, Special Paper.

Alonso, J.L., Rodríguez Fernández, L.R., García Sansegundo, J., Heredia, N., Farias, P., Gallastegui, J., 2005. Gondwanic and Andean structure in the Argentine Central Precordillera: The Río San Juan section revisited. 6th International Symposium on Andean Geodynamics, Paris, IRD Editions, 36-39.

Álvarez-Marrón, J., Rodríguez Fernández, L.R., Heredia, N., Busquets, P., Colombo, F., Brown, D., 2006. Neogene structures overprinting Palaeozoic thrust systems in the Andean Precordillera at 30o S latitude. Journal of the Geological Society, London, 163, 949-964.

Astini, R.A., 1998a. Stratigraphic evidence supporting the rifting, drifting and collision of the Laurentian Precordillera terrane of western Argentina. In: Pankhurst, R.J., Rapela, C.W. (eds.). The Proto- Andean Margin of Gondwana. Geological Society of London, Special Publication, 142, 11-33.

Astini, R.A., 1998b, El Conglomerado de Las Vacas y el Grupo Trapiche de la Precordillera: tectónica distensiva en el Ordovícico Tardío. Revista de Asociación Geológica Argentina, 53(4), 489-503

Astini, R.A., 2002. Los conglomerados basales del Ordovícico de Ponón Trehue (Mendoza) y su significado en la historia sedimentaria del terreno exótico de Precordillerra. Revista de Asociacíon Geologíca Argentina, 57(1), 19-34.

Astini, R.A., 2003. The Ordovician-Proto-Andean Basins. In: Benedetto, J.L. (ed.). Ordovician fossils of Argentina. Secretaría de Ciencia y Tecnología, Universidad Nacional de Córdoba, 1-74.

Astini, R.A., Dávila, F.M., 2004. Ordovician back arc foreland and Ocloyic thrust belt development on the western Gondwana margin as a response to Precordillera terrane accretion. Tectonics, 23, TC4008, 1-19.

Astini, R.A., Rapalini, A.E., 2003. Discussion of “Proterozoic- Early Paleozoic evolution in western South America - a discussion” in Tectonophysics, 354, 121-137 (2002). Tectonophysics, 366, 143-148.

Astini, R.A., Thomas, W.A., 1999. Origin and evolution of the Precordillera terrane of western Argentina: A drifted Laurentian orphan. In: Ramos, V.A., Keppie, J.D. (eds.) Laurentia Gondwana Connections before Pangea. Geological Society of America, Special Paper, 336, 1-20.

Astini, R., Benedetto, J., Vaccari, N., 1995. The early Palaeozoic evolution of the Argentine Precordillera as a Laurentian rifted, drifted and collided terrane, a geodynamic model, Geological Society of America Bulletin 107, 253—273.

Astini, R., Ramos, V.A., Benedetto, J., Vaccari, N., Cañas, F.L., 1996. La Precordillera: Un terreno exótico a Gondwana. XIII Congreso Geológico Argentino y III Congreso de Exploración de Hidrocarburos, Actas, vol. 5, pp. 293—324.

Astini, R.A., Thomas,W.A.,Yochelson, E.L., 2004. Salterella in the Argentine Precordillera: an Early Cambrian palaeobiogeographic indicator of Laurentian affinity. Palaeogeography, Palaeoclimatology, Palaeoecology, 213, 125-132.

Baldis, B.A., Peralta, S., Villegas, R., 1989. Esquematizaciones de una possible transcurrencia del terrane de Precordillera como fragmento continental procedente de áreas pampeanobonaerenses.
Tucumán, Argentina, Instituto Superior de Correlación Geológica, 5, 81-100.

Baldo, E.G., Fanning, C.M., Rapela, C.W., Pankhurst, R.J., Casquet, C., Galindo, C., 2003. U-Pb Shrimp dating of rhyolite volcanism in the Famatinian belt and K-bentonites in the Precordillera. In: Albanesi, G.L., Beresi, M.S., Peralta, S.H. (eds.). Ordovician from the Andes. INSUGEO, Serie Correlación
Geológica, 17, 185-189.

Baldo, E., Dahlquist, J., Rapela, C.W., Casquet, C., Pankhurst, R.J., Galindo, C., Fanning, C.M., 2005. Early Ordovician peraluminous magmatism in the Sierra de Pie de Palo, (Western Sierras Pampeanas): geotectonic implications. In: Pankhurst, R.J., Veiga, G.D. (eds.). Gondwana 12: Geological and Biological Heritage of Gondwana. Academia Nacional de Ciencias, Córdoba, Argentina, Abstracts, p. 57.

Basei, M.A.S., Siga Jr., O., Masquelin, H., Harara, O.M., Reis Neto, J.M., Preciozzi, F., 2000. The Dom Feliciano Belt of Brazil and Uruguay and its Foreland Domain, the Rio de La Plata Craton: Framework, Tectonic Evolution and correlation with similar provinces of southwestern Africa. In: Cordani, U.G., Milani, E.J., Thomaz Filho, A., Campos, D.A. (eds.). Tectonic Evolution of South America. 31st International Geological Congress, Río de Janeiro, 311-334.

Basei, M.A.S., Frimmel, H.E., Nutman, A.P., Preciozzi, F., Jacob, J., 2005. A connection between the Neoproterozoic Dom Feliciano (Barzil/Uruguay) and Gariep (Namibia/South Africa) orogenic belts - evidence from a reconnaissance provenance study. Precambrian Research, 139, 195-221.

Basei, M., Ramos, V.A., Vujovich, G.I., Poma, S.. 1998. El basamento metamórfico de la Cordillera Frontal de Mendoza: Nuevos datos geocronológicos e isotópicos. X Congreso Latinoamericano de Geología y VI Congreso Nacional de Geología Económica (Buenos Aires), Actas, vol. 2, pp. 412—417

 

Cingolani, C. A. and Varela, R. 1999. The San Rafael Block, Mendoza (Argentina): Rb-Sr isotopic age of basement rocks. II South American Symposium on Isotope Geology. Carlos Paz, Argentina

 

Behl, R.J., 1999. Since Bramlette (1946): The Miocene Monterey Formation of California revised. In: Sloan, D., Stout, D.L. (eds.). Classic Cordilleran Concepts: A View from California. Boulder, Colorado, Geological Society of America, Special Paper, 338, 301-313.

Benedetto, J.L., 1998. Early Paleozoic brachiopods and associated shelly faunas from western Gondwana: their bearing on the geodynamic history of the pre-Andean margin. In: Pankhurst, R.J., Rapela, C.W. (eds.). The Proto-Andean Margin of Gondwana. Geological Society of London, Special
Publication, 142, 57-83.

Benedetto, J.L., 2003. Paleobiogeography. In: Benedetto, J.L. (ed.). Ordovician Fossils of Argentina. Universidad Nacional de Córdoba, Argentina, Secretaría de Ciencia y Tecnología, 91-109.

Benedetto, J.L., Carrera, M., Sánchez, T.M., Vaccari, N.E., 1995. The evolution of faunal provincialism in the Argentine Precordillera during the Ordovician: New evidence and paleogeographic implications. In: Cooper, J.D., Droser, M.L., Finney, S.C. (eds.). Ordovician Odyssey: Short Papers for the Seventh International Symposium on the Ordovician System, Las Vegas, Nevada, USA, Pacific Section
Society for Sedimentary Geology (SEPM) Book n. 77, 181-184.

Benedetto, J.L., Sánchez, T.M., Carrera, M.G., Brussa, E.D., Salas, M.J., 1999. Paleontological constraints on successive paleogeographic positions of Precordillera terrane during the early Paleozoic. In: Ramos, V.A., Keppie, J.D. (eds.). Laurentia-Gondwana Connections before Pangea. Geological
Society of America, Special Paper, 336, 21-42.

Bergström, S.M., 1990. Relations between conodont provincialism and the changing palaeogeography during the Early Paleozoic. In: McKerrow,W.S., Scotese, C.R. (eds.). Palaeozoic Palaeogeograpy and Biogeography. The Geological Society Memoir, 12, 105-121.

Bond, G.C., Nickeson, P.A., Kominz, M.A., 1984. Breakup of a super-continent between 625 Ma and 555 Ma: new evidence and implications for continental histories. Earth and Planetary Science Letters, 70, 325-345.

Bordonaro, O., 1986. Bioestratigrafía del Cámbrico Inferior de San Juan. IV Congreso Argentino de Paleontología y Bioestratigrafía, Mendoza, Argentina, Actas, 1, 19-27.

Bordonaro, O., 2003. Review of the Cambrian Stratigraphy of the Argentine Precordillera. Geologica Acta, 1(1), 11-21.

Bordonaro, O., Banchig. A., 1995. Trilobites laurénticos en el Cámbrico de la Precordillera argentina. V Congreso Argentino de Paleontología y Bioestratigrafía, Trelew, Argentina, Actas, 59-65.

Bordonaro, O., Beresi, M., Keller, M., 1993. Reinterpretación estratigráfica del Cámbrico del área de San Isidro, Precordillera de Mendoza. XII Congreso Geológico Argentino, Mendoza, Argentina, Actas, 2, 12-19.

Borrello, A.V., 1963. Fremontella inopinata n. sp. del Cámbrico de la Argentina. Ameghiniana, 3, 51-55.

Borrello, A.V., 1964. Sobre la presencia del Cámbrico inferior Olenellidiano en la Sierra de Zonda, Precordillera de San Juan. Ameghiniana, 3, 313-318.

Borrello, A.V., 1969. Los geosinclinales de la Argentina. Dirección Nacional de Geología y Minería, Anales, XIV, 1-188.

Borrello, A., 1971. The Cambrian of South America. In: Holland, C. (ed.). Cambrian of the New World. Wiley Intersciences, 1, 385-438.

Bossi, J., Gaucher, C., 2004. The Cuchilla Dionisio Terrane, Uruguay: An Allochthonous Block Accreted in the Cambrian to SW-Gondwana. Gondwana Research, 7(3), 661-674.

Brito Neves, B.B., Campos Neto, M., Fuck, R.A., 1999. From Rodinia to Western Gondwana: an approach to the Brasiliano- Pan African Cycle and orogenic collage. Episodes, 22(3), 155-166.

Busby, C.J., Ingersoll. R.V., 1995. Tectonics of Sedimentary Basins. Cambridge, Massachusetts, Blackwell Science Inc., 579 pp.

Campos Neto, M.C., Figueiredo, M.C.H., 1995. The Rio Doce Orogeny, Southeastern Brazil. Journal of South American Earth Sciences, 8(2), 143-162.

Casquet, C., Pankhurst, R.J., Rapela, C.W., Fanning, C.M., Galindo, C., Baldo, E., Gonzalez-Casado, J.M., Dahlquist, J.M., Saavedra, J., 2005. The Maz suspect terrane: a new Proterozoic domain in the Western Sierras Pampeanas. In: Pankhurst, R.J., Veiga, G.D. (eds.). Gondwana 12: Geological
and Biological Heritage of Gondwana, Academia Nacional de Ciencias, Córdoba, Argentina, Abstracts, p. 92.

Christie-Blick, N., Biddle, K.T., 1985. Deformation and basin formation along strike slip faults. In: Biddle, K.T., Christie- Blick, N. (eds.). Strike-Slip Deformation, Basin Formation, and Sedimentation. Society of Economic Paleontologists and Mineralogists, Special Publication, 37, 1-34.

Cocks, L.R.M., Torsvik, T.H., 2002. Earth geography from 500 to 400 million years ago: a faunal and palaeomagnetic review. Journal of the Geological Society of London, 159, 631-644.

Dalla Salda, L. H., Cingolani, C., Varela, R., 1992a. Early Paleozoic orogenic belt of the Andes in southwestem South America: Result of Laurentia—Gondwana collision?. Geology 20, 617—620.

 

Dalla Salda, L. H., Dalziel, I., Cingolani, C., Varela, R., 1992b. Did the Taconic Appalachians continue into southern South America?. Geology 20, 1059—1062.

 

Dalla Salda, L. H., López de Luchi, M. G., Cingolani, C. A., Varela, R., 1998. Laurentia—Gondwana collision: The origin of the Famatinian—Appalachian Orogenic Belt (a review). In: Pankhurst, R. J., Rapela, C. W. (Eds.). The Proto—Andean Margin of Gondwana. Geological Society London, Special Publication, 142, 219—234.

Dalziel, I.W.D., 1997. Neoproterozoic-Paleozoic geography and tectonics: Review, hypothesis, environmental speculation. Geological Society of America Bulletin, 109(1), 16-42.

Dalziel, I.W.D., Dalla Salda, L., Cingolani, C., Palmer, P., 1996. The Argentine Precordillera: A Laurentian Terrane? GSA Today, 6, 16-18.

Dickerson. P. W., Keller, M., 1998. The Argentine Precordillera: Its odyssey from the Laurentian Ouachita margin towards the Sierras Pampeanas of Gondwana. In: Pankhurst, R.J., Rapela, C.W. (Eds.), The Proto— Andean Margin of Gondwana. Geological Society London, Special Publication, 142, 325—341.

Dristas, J.A., Frisicale, M.C., 1987. Rocas piroclasticas en el sector suroeste de las sierras septentrionales de la Provincia de Buenos Aires. Revista de la Asociación Argentina de Mineralogía, Petrología y Sedimentología, 18, 33-45.

Fanning, C.M., Pankhurst, R.J., Rapela, C.W., Baldo, E.G., Casquet, C., Galindo, C., 2004. K- bentonites in the Argentine Precordillera contemporaneous with rhyolite volcanism in the Famatinian Arc. Journal of the Geological Society, London, 161, 747-756.

Finney, S.C., Berry, W.B.N., 1997. New perspectives on graptolite distributions and their use as indicators of platform margin dynamics. Geology, 25, 919-922.

Finney, S.C., Chen, Xu, 1990. The relationship of Ordovician graptolite provincialism to palaeogeography. In: McKerrow, W.S., Scotese, C.R. (eds.). Palaeozoic Palaeogeograpy and Biogeography. The Geological Society Memoir, 12, 122-128.

Finney, S., Peralta, S., Heredia, S., 2005a. Where is the foreland basin at the Cuyania-Gondwana collisional zone? In: Pankhurst, R.J., Veiga, G.D. (eds.). Gondwana 12: Geological and Biological Heritage of Gondwana, Academia Nacional de Ciencias, Córdoba, Argentina, Abstracts, p. 156.

Finney, S., Peralta, S., Gehrels, G., Marsaglia, K., 2005b. The Early Paleozoic history of the Cuyania (greater Precordillera) terrane of western Argentina: evidence from geochronology of detrital zircons from Middle Cambrian sandstones. Geologica Acta, 3(4), 339-354.

Finney, S., Gleason, J., Gehrels, G., Peralta, S. and Aceñolaza, G. 2003. Early Gondwanan connection for the Argentine Precordillera terrane. Earth and Planetary Science Letters  205, 349-359.

Finney, S., Gleason, J., Gehrels, G., Peralta, S. and Aceñolaza, G., 2004. Corrigendum to "Early Gondwanan connection for the Argentine Precordillera terrane". [Earth and Planetary Science Letters  205, (2003) 349-359. Earth and Planetary Science Letters  219, 413.

Finney, S., Grove, M., Gehrels, G., Peralta, S., Heredia, S., 2005c. Early Mesoproterozoic zircons from conglomerate clasts and sandstones in Cambrian and Ordovician siliciclastic strata of the Argentine Precordillera. In: Pankhurst, R.J., Veiga, G.D. (eds.) Gondwana 12: Geological and Biological
Heritage of Gondwana, Academia Nacional de Ciencias, Córdoba, Argentina, Abstracts, p. 155.

Finney, S.C., Peralta, S.H., Heredia, S., Gehrels, G., McGraw, J., Gaucher, C., Poiré, D.G., 2006. Detrital-zircon geochronology of Cambrian-Devonian sandstones of the Cuyania (greater Precordillera) terrane of western Argentina and Neoproterozoic sandstones deposited on the Río de la Plata carton of Uruguay and eastern Argentina. Geological Society of America Abstracts with Programs, 38(7), p. 410.

The parautochthonous Gondwanan origin of the Cuyania (greater Precordillera) terrane of Argentina: A re-evaluation of evidence used to support an allochthonous Laurentian origin. Geologica Acta, Vol.5, Nº 2, 2007, 127-158.

Fortey, R.A., Owen, R.M., 1997. Evolutionary History. In: Whittington, H.B., Chatterton, B.D.E., Speyer, S.E., 14 others. Treatise on Invertebrate Paleontology, Part O Arthropoda 1, Trilobita, Revised, Volume 1: Introduction, Order Agnostida, Order Redlichiida, Boulder, Colorado and Lawrence, Kansas, The Geological Society of America and The University of Kansas, 249-288.

Fritz, W.H.,Yochelson, E.L., 1988. The status of Salterella as a Lower Cambrian index fossil. Canadian Journal of Earth Sciences, 25, 403-416.

Galindo, C., Casquet, C., Rapela, C., Pankhurst, R.J., Baldo, E., Saavedra, J., 2004. Sr, C, O isotope geochemistry and stratigraphy of Precambrian and lower Paleozoic carbonate sequences from the Western Sierras Pampeanas of Argentina: tectonic implications. Precambrian Research, 131, 1041-
1056.

Gaucher, C., Poiré, D.G., Finney, S.C., Valencia, V.A., Blanco, G., Pamoukaghlian, K., Peral, L.G., 2006. Detrital zircon ages in Neoproterozoic sedimentary units of the Río de la Plata craton: unravelling its complex geological evolution. Geological Society of America Abstracts with Programs, 38(7), 410.

Gaucher, C., Sial, A.N., Ferreira, V.P., Pimentel, M.M., Chiglino, L., Sprechmann, P., in press. Chemostratigraphy of the Cerro Victoria Formation (Lower Cambrian, Uruguay): evidence
for progressive stabilization across the Precambrian- Cambrian boundary. Chemical Geology.

Geyer, G., Elicki, O., 1995. The Lower Cambrian Trilobites from the Görlitz Synclinorium (Germany) - review and new results. Paläontologische Zeitschrift, 69, 87-119. Geyer, G., Palmer, A.R., 1995. Neltneriidae and Holmiidae (Trilobita) from Morocco and the problem of Early Cambrian intercontinental correlation. Journal of Paleontology, 69, 459-474.

Gleason, J.D., Finney, S.C., Gehrels, G.E., 2002. Paleotectonic Implications of a Mid- to Late Ordovician Provenance Shift, as Recorded in Sedimentary Strata of the Ouachita and Southern Appalachian Mountains. Journal of Geology, 110, 291-304.

Gleason, J.D., Finney, S.C., Peralta, S.H., Gehrels, G.E., Marsaglia, K.M., 2007. Zircon and whole rock Nd-Pb isotopic provenance of Middle and Upper Ordovician siliciclastic rocks, Argentine Precordillera. Sedimentology, 54(1), 107-136.

Gower, C.F., Krogh, T.E., 2002. A U-Pb Geochronological review of the Proterozoic history of the eastern Grenville Province. Canadian Journal of Earth Sciences, 39, 795-829.

Gozalo, R., Mayoral, E., Gámez Vintaned, J.A., Dies, M.E., Muñiz, F., 2003. A new occurrence of the genus Tonkinella in northern Spain and the Middle Cambrian intercontinental correlation. Geologica Acta, 1, 121-126.

Heredia, S., Beresi, M.S., 2004. La Formación Empozada y su relación estratigráfica con la Formación Estancia San Isidro (nom. nov.), Ordovícico de la Precordillera de Mendoza. Revista de la Asociación Geológica Argentina, 59(2), 178-192.

Hoffman, P.F., 1991. Did the Breakout of Laurentia Turn Gondwanaland Inside-Out? Science, 252, 1409-1412.

Hollingsworth, J.S., 2005. A trilobite fauna in a storm bed in the Poleta Formation (Dyeran, Lower Cambrian), western Nevada, U.S.A. Geosciences Journal, 9, 129-143.

Hollingsworth, J.S., 2006. Holmiidae (Trilobita: Olenellina) of the Montezuman Stage (Early Cambrian) in western Nevada. Journal of Paleontology, 80, 309-332.

Howell, D.G., Stuart, C.J., Platt, J.P., Hill, D.J., 1974. Possible Strike-Slip Faulting in the Southern California Borderland. Geology, 2(2), 93-98.

Huff, W.D. Bergström, S.M., Kolata, D.R., Cingolani, C.A., Astini, R.A., 1998. Ordovician K- bentonites in the Argentine Precordillera: relations to Gondwana margin evolution. In: Pankhurst, R.J., Rapela, C.W. (eds.). The Proto-Andean Margin of Gondwana. Geological Society of London, Special Publication, 142, 107-126

Isaacs, C.M., Garrison, R.E., 1983. Petroleum Generation and Occurrence in the Miocene Monterey Formation, California. Los Angeles, Pacific Section, Society of Economic Paleontologists and Mineralogists, 228 pp.

Karlstrom, K.E., Åhäll, K-I., Harlan, S.S., Williams, M.L., McLelland, J., Geissman J.W., 2001. Long lived (1.8 – 1.0 Ga) convergent orogen in southern Laurentia, its extensions to Australia and Baltica, and implications for refining Rodinia. Precambrian Research, 111, 5-30.

Kay, S.M., Orrell, S., Abbruzzi, J.M., 1996. Zircon and whole rock Nd-Pb isotopic evidence for a Grenville age and Laurentian origin for the basement of the Precordillera terrane in Argentina. The Journal of Geology, 104, 637-648.

Keller, M., 1999. Argentine Precordillera: Sedimentary and Plate Tectonic History of a Laurentian Crustal Fragment in South America. Geological Society of America, Special Paper, 341, 131 pp.

Keller, M., Bordonaro, O., Beresi, M., 1993. The Cambrian of San Isidro, Mendoza, Argentina: Facies and sedimentology at the platform slope transition. Neues Jahrbuch für Geologie und Paläontologie Monatshefte, 1993(6), 373-383.

Keller, M., Cañas, F., Lehnert, O., Vaccari, N.E., 1994. The Upper Cambrian and Lower Ordovician of the Precordillera (Western Argentina): Some stratigraphic reconsiderations. Newsletters in Stratigraphy, 31, 115-132.

Llambías, E. J., Melchor, R. N., Tickyj, H., Sato, A. M., 1996. Geología del Bloque del Chadileuvú. XIII Congreso Geológico Argentino y III Congreso de Exploración de Hidrocarburos, Actas, vol. 5, pp. 417—425.

Llambías, E.J., Sato, A.M., Suárez, A.O., Prozzi, C., 1998. The granitoids of the Sierra de San Luis. In: Pankhurst, R.J., Rapela, C.W. (eds.). The Proto-Andean Margin of Gondwana. Geological Society of London, Special Publication, 142, 325-341.

Lehnert, O., Keller, M., Bordonaro, O., 1998. Early Ordovician conodonts from the southern Cuyania terrane (Mendoza Province, Argentina). In: Szaniawski, H. (ed.). Proceedings of the Sixth European Conodont Symposium (ECOS VI). Palaeontologia Polonica, 58, 47-65.

Lehnert, O., Miller, J.F., Repetski, J.E., 1997. Paleogeographic significance of Clavohamulus hintzei Miller (Conodonta) and other Ibexian conodonts in an early Paleozoic carbonate platform facies of the Argentine Precordillera. Geological Society of America Bulletin, 109, 429-443.

Lehnert, O., Bergström, S.M., Keller, M., Bordonaro, O., 1999. Ordovician (Darriwilian-Caradocian) conodonts from the San Rafael Region, west-central Argentina: biostratigraphic, paleoecologic, and paleogeographic implications. Bollettino della Società Paleontologica Italiana, 37, 199- 214.

Lieberman, B.S., 1998. Cladistic analysis of the Early Cambrian olenelloid trilobites. Journal of Paleontology, 72, 59-78.

Lieberman, B.S., 2001. Phylogentic analysis of the Olenellina Walcott, 1890 (Trilobita, Cambrian). Journal of Paleontology, 75, 96-115.

Lieberman, B.S., 2002. Phylogenetic analysis of some basal early Cambrian trilobites, the biogeographic origins of the Eutrilobita, and the timing of the Cambrian radiation. Journal of Paleontology, 76, 692-708.

Link, M.H., 2003. Depositional systems and sedimentary facies of the Miocene-Pliocene Ridge Basin Group, Ridge Basin, southern California. In: Crowell, J.C. (ed.). Evolution of Ridge Basin, southern California: An interplay of sedimentation and tectonics. Boulder, Colorado, Geological Society of America, Special Paper, 367, 17-87.

Lira, R., Millone, H.A., Kirschbaum, A.M., Moreno, R.S., 1997. Calc-Alkaline Arc Granitoid Activity in the Sierra Norte- Ambargasta Ranges, Central Argentina. Journal of South American Earth Sciences, 10(2), 157-177.

Lochman-Balk, C., Wilson, J.L., 1958. Cambrian Biostratigraphy in North America. Journal of Paleontology, 32(2), 312-350.

McDonough, M. R., Ramos. V. A.. Isachsen. C. E., Bowring, S. A,. Vujovich, G., 1993. Edades preliminares de circones del basamento de la Sierra de Pie de Palo. Sierras Pampeanas occidentales de San Juan, sus implicancias para el snpercontinente proterozoico de Rodinia. XII Congreso Geológico Argentino y II Congreso de Exploración de Hidrocarburos, Actas.,vol. 3. pp. 340—342.

Maletz, J., Ortega, G., 1995. Ordovician graptolites of South America: palaeogeographic implications. In: Cooper, J.D., Droser, M.L., Finney, S.C. (eds.). Ordovician Odyssey: Short Papers for the Seventh International Symposium on the Ordovician System, Las Vegas, Nevada, USA, Pacific Section Society for Sedimentary Geology (SEPM) Book. 77, 189-192.

McClelland, W.C., Ellis, J.R., Roeske, S.M., Mulcahy, S.R., Vujovich, G.I., Naipauer, M., 2005. U-Pb SHRIMP igneous zircon ages and LA-ICPMS detrital zircon ages from metamorphic rocks between the Precordillera terrane and the Gondwana margin, Sierra de la Huerta to Pie de Palo, northwest
Argentina. In: Pankhurst, R.J., Veiga, G.D. (eds.). Gondwana 12: Geological and Biological Heritage of Gondwana, Academia Nacional de Ciencias, Córdoba, Argentina, Abstracts, p. 250.

McElhinny, M.W., Powell, C., Pisarevsky, S.A., 2003. Paleozoic terranes of eastern Australia and the drift history of Gondwana. Tectonophysics, 362, 41-65.

McKerrow, W.S., Scotese, C.R., Brasier, M.D., 1992. Early Cambrian continental reconstructions. Journal of the Geological Society, London, 149, 599-606.

Meert, J.G., 2003. A synopsis of events related to the assembly of eastern Gondwana. Tectonophysics, 362, 1-40.

Mitchell, C.E., Brussa, E.D., Astini, R.A., 1997. Biogeography of Middle and Upper Ordovician graptolites, Precordillera terrane, Argentina: plate tectonic implications. Geological Society of America, Abstracts with Programs, 29, A379.

Nilsen, T.H., Sylvester, A.G., 1995. Strike-Slip Basins. In: Busby, C.J., Ingersoll. R.V. (eds.). Tectonics of Sedimentary Basins. Cambridge, Massachusetts, Blackwell Science Inc., 425-456.

Palmer, A.R., 1971. The Cambrian of the Appalachians and eastern new England regions, eastern United States. In: Holland, C.H. (ed.). Cambrian of the New World. Wiley Intersciences, 1, 169-218.

Palmer. A.R., 1972. Problems of Cambrian Biogeography. 24th International Geological Congress, Montreal, Canada, Section 7 (Paleontology), Proccedings, 310-315.

Palmer, A.R., Gatehouse, C.G, 1972. Early and Middle Cambrian Trilobites from Antarctica. U.S. Geological Survey Professional Paper, 456-D, D1-D37.

Palmer, A.R., Halley, R.B., 1979. Physical Stratigraphy and Trilobite Biostratigraphy of the Carrara Formation (Lower and Middle Cambrian) in the Southern Great Basin. U.S. Geological Survey Professional Paper, 1047, 131 pp.

Palmer, A.R., Repina, L.N., 1993. Through a glass darkly: taxonomy, phylogeny, and biostratigraphy of the Olenellina, The University of Kansas Paleontological Contributions n. 3, 1-35.

Palmer, A.R., Repina, L.N., 1997. Introduction to Suborder Olenellina. In: Whittington, H.B., Chatterton, B.D.E., Speyer, S.E., and 14 others. Treatise on Invertebrate Paleontology, Part O Arthropoda 1, Trilobita, Revised, Volume 1: Introduction, Order Agnostida, Order Redlichiida. Boulder, Colorado
and Lawrence, Kansas, The Geological Society of America and The University of Kansas, 405-429.

Palmer, A.R., Rowell, A.J., 1995. Early Cambrian Trilobites from the Shackleton Limestone of the Central Transantarctic Mountains. The Paleontological Society Memoir 45, Journal of Paleontology, 69 (no. 6, suppl.), 28 pp.

Pankhurst, R.J., Rapela, C.W., 1998, The Proto-Andean Margin of Gondwana. Geological Society, London, Special Publications, 142, 383 pp.

Pankhurst. R. J.,. Rapela, C. W., Saavedra, J., Baldo, E., Dahlquist. J., Pascua, I., Fanning. C. M., 1998. The Famatinian magmatic arc in the central Sierras Pampeanas: An Early to mid-Ordovician continental arc on the Gondwana margin. ln: Pankhtrrst. R.J., Rapela, C.W. (Eds.), The Proto—Andean Margin of Gondwana. Geological Society London, Special Publication. 142, 343—367.

Pankhurst, R.J., Veiga, G.D., 2005. Gondwana 12: Geological and Biological Heritage of Gondwana. Academia Nacional de Ciencias, Córdoba, Argentina, Abstracts, 385 pp.

Peralta, S.H., 2005a. Formación Los Sombreros: un evento diastrófico extensional del Devónico (inferior? medio?) en la Precordillera argentina. XVI Congreso Geológico Argentino, La Plata, Actas, 4, 322-325.

Peralta, S.H., 2005b. The Lower Emsian?-Middle Devonian? Extensional basins of the Los Sombreros and Rinconada Formations: its tecto-sedimentary significance in the evolution of the Precordillera. In: Pankhurst, R.J., Veiga, G.D. (eds.). Gondwana 12: Geological and Biological Heritage of Gondwana, Academia Nacional de Ciencias, Córdoba, Argentina, Abstracts, p. 289.

Peralta, S.H., Heredia, S., 2005. Depósitos de olistostroma del Devónico (inferior?-medio?), Formación Los Sombreros, en la quebrada de San Isidro, Precordillera de Mendoza, Argentina. XVI Congreso Geológico Argentino, La Plata, Actas, 4, 326-331.

Pillola, G.L., 1990. Lithologie et Trilobites du Cambrien inférieur du SW de la Sardaigne (Italie): implications paléobiogéographiques. Comptes Rendus de l’Académie des Sciences, Paris, série II, 310, 321-328.

Poiré, D.G., Spalletti, L.A., Del Valle, A., 2003. The Cambrian- Ordovician siliciclastic platform of the Balcarce Formation (Tandilia System, Argentina): Facies, trace fossils, palaeoenvironments and sequence stratigraphy. Geologica Acta, 1(1), 41-60.

Ramos, V.A., 1995. Sudamérica: un mosaico de continentes y océanos. Ciencia Hoy, 6, 24-29.

Ramos, VA.. Basei, MAS., 1997. The basement of Chilenia: An exotic continental terrane to Gondwana during the Early Paleozoic. In: Terrane Dynamics-97 (Christchurch, New Zealand), Conference Abstracts, pp. 140—143.

Ramos, VA., Jordan, T., Allmendinger. R.W., Kay, S.M., Cortés, J.M., Palma, M.A., 1984. Chilenia: Un terreno alóctono en la evolución paleozoica de los Andes Centrales. IX Congreso Geológico Argentino, Actas, vol. 2, pp~. 84—106.

Ramos, V.A., Jordan, T.E., Allmendinger, R.W., Mpodozis, M.C., Kay, S.M., Cortes, J.M., Palma, M., 1986. Paleozoic terranes of the central Argentine-Chilean Andes. Tectonics, 5, 855-880.

Ramos, VA., Vujovich, G., Dallmever. R.D., 1996. Los klippes y ventanas tectónicas preándicas de la Sierra de Pie de Palo (San Juan): Edad e imiplicaciones tectontcas. XIII Congreso Geológico Argentino y III Congreso de Exploración de Hidrocarburos, Actas, vol. 5. pp. 377—391.

Ramos, V.A., Dallmeyer, R.D., Vujovich, G., 1998. Time constraints on the Early Palaeozoic docking of the Precordillera, central Argentina. In: Pankhurst, R.J., Rapela, C.W. (eds.). The Proto Andean Margin of Gondwana. Geological Society, London, Special Publications No. 142, 143-158.

Ramos, V.A., Keppie, J.D., 1999. Laurentia-Gondwana Connections before Pangaea: Geological Society of America, Special Paper, 336, 276 pp..

Rapalini, A.E., Astini, R.A., 1998. Paleomagnetic confirmation of the Laurentian origin of the Argentine Precordillera. Earth and Planetary Science Letters, 155, 1-14.

Rapela, C.W., Pankhurst, R.J., Casquet, C., Baldo, E., Saavedra, J., Galindo, C., Fanning, C.M., 1998. The Pampean Orogeny of the southern proto-Andes: Cambrian continental collision in the Sierras de Córdoba. In: Pankhurst, R.J., Rapela, C.W. (eds.). The Proto-Andean Margin of Gondwana. Geological
Society of London, Special Publication, 142, 181-218.

Rapela, C.W., Pankhurst, R.J.. Casquet. C., Baldo, E.. Saavedra, J., Galindo, C., l998a. Early evolution of the Proto—Andean margin of South Amenica. Geology 26, 707—710.

Rapela, C.W., Pankhurst, R.J., Casquet, C., Fanning, C.M., Galindo, C., Baldo, E., 2005a. Datacion U-Pb SHRIMP de circones detriticos en paranfibolitas neoproterozoicas de las secuencia Difunta Correa (Sierra Pampeanas Occidentales, Argentina), Geogaceta, 38, 227-230.


Rapela, C.W., Fanning, C.M., Baldo, E.G., Dahlquist, J.A., Pankhurst, R.J., Mura, J.A., 2005b. Coeval S- and I-type granites in the Sierra de Ancasti, Eastern Sierras Pampeanas, Argentina. In: Pankhurst, R.J., Veiga, G.D. (eds.). Gondwana 12: Geological and Biological Heritage of Gondwana, Academia Nacional de Ciencias, Córdoba, Argentina, Abstracts, p. 307.


Rapela, C.W., Fanning, C.M., Pankhurst, R.J., 2005c. The Rio de la Plata Craton: the search for its full extent. In: Pankhurst, R.J., Veiga, G.D. (eds.). Gondwana 12: Geological and Biological Heritage of Gondwana, Academia Nacional de Ciencias, Córdoba, Argentina, Abstracts, p. 308.


Resser, C.E., 1938. Cambrian System (Restricted) of the Southern Appalachians. Geological Society of Americam Special Paper, 15, 140 pp.


Roeske, S.M., McClelland, W.C., Cain, J.C. IV, Mulcahy, S.R., Vujovich, G.I., Iriondo, A., 2005. Paleozoic record of convergence and extension within the arc forearc transition of the Famatina arc, as recorded in western Sierra de la Huerta, Argentina. In: Pankhurst, R.J., Veiga, G.D., (eds.). Gondwana 12: Geological and Biological Heritage of Gondwana, Academia Nacional de Ciencias, Córdoba, Argentina,
Abstracts, p. 315.


Ross, G.M., Villeneuve, M., 2003. Provenance of the Mesoproterozoic (1.45 Ga) Belt basin (western North America): Another piece in the pre-Rodinia paleogeographic puzzle. Geological Society of America Bulletin, 115(10), 1191-1217.


Ross, R.J., Jr., 1975. Early Paleozoic trilobites, sedimentary facies, lithospheric plates, and ocean currents. Fossils and Strata, 4, 307-329.

Sato, A.M., Tickyj, H.. Llambías, E.J., 1998. Rb—Sr Grenvillian age from the Las Matras diorite, La Pampa province, Argentina. X Congreso Latinoamericano de Geología y VI Congreso Nacional de Geología Económica (Buenos Aires), Actas, vol. 2, p. 418.

Sato, A.M., Tickyj, H.. Llambías, EJ., Sato, K., 1999. Rh—Sr, Sm—Nd and K—Ar age constraints of the Grenvillian Las Matras pluton, central Argentina. II South American Symposium on Isotopc Geology (Villa Carlos Paz, Argentina), Actas, pp. 122—126.

Sato, A.M., Tickyj, H.. Llambías, EJ., Sato, K., 2000. The Las Matras tonalitic-trondhjemitic pluton, central Argentina: Grenvillian-age constrains, geochemical characteristics, and regional implications. Journ. South American earth Sciences, 13: 587-610.

Shergold, J.H., 1969. Oryctocephalidae (Trilobita Middle Cambrian) of Australia. Bureau of Mineral Resources, Geology and Geophysics Bulletin, 104, 66 pp.


Shergold, J.H.,Webers, G.F., 1992. Late Dresbachian (Idamean) and other trilobite faunas from the Heritage Range, Ellsworth Mountains, West Antarctica. In: Webers, G.F., Craddock, C., Splettsoesser, J.F. (eds.). Geology and Paleontology of the Ellsworth Mountains, West Antarctica. Geological
Society of America Memoir, 170, 125-168.


Sims, J.P., Ireland, T.R., Camacho, A., Lyons, P., Pieters, P.E., Skirrow, R.G., Stuart-Smith, P.G., Miró, R., 1998. U-Pb, Th-Pb and Ar-Ar geochronology from the southern Sierras Pampeanas, Argentina: implications for the Palaeozoic ectonic evolution of the western Gondwana margin. In: Pankhurst, R.J., Rapela, C.W. (eds.). The Proto-Andean Margin of Gondwana. Geological Society of London, Special
Publication, 142, 259-281.

Sinclair, H.D., 1997. Tectonostratigraphic model for underfilled peripheral foreland basins: An Alpine perspective. Geological Society of America Bulletin 109(3), 324-346.


Sprechmann, P., Gaucher, C., Blanco, G., Montaña, J., 2004. Stromatolitic and Trace Fossils Community of the Cerro Victoria Formation, Arroyo del Soldado Group (Lowermost Cambrian, Uruguay). Gondwana Research, 7(3), 753-766.


Sweet, W.C., Bergström, S.M., 1984. Conodont provinces and biofacies of the Late Ordovician. Geological Society of America Special Paper, 196, 69-87.


Tassinari, C.C.G., Macambira, M.J.B., 1999. Geochronological provinces of the Amazonian Craton. Episodes, 22(3), 174-182.

Thomas, W.A., 1991. The Appalachian-Ouachita rifted margin of southeastern North America. Geological Society of America Bulletin, 103, 415-431.

Thomas, W.A., Astini, R.A., 1996. The Argentine Precordillera: A traveler from the Ouachita embayment of North American Laurentia. Science, 273, 752-757.


Thomas,W.A., Astini, R.A., 1999. Simple-shear conjugate margins of the Argentine Precordillera and the Ouachita embayment of Laurentia. Geological Society of America Bulletin, 111, 1069-1079.

.

Thomas, W.A., Astini, R.A., 2003. Ordovician accretion of the Argentine Precordillera terrane to Gondwana: a review. Journal of South American Earth Sciences, 16, 67-79.


Thomas, W.A., Astini, R.A., 2005. Vestiges of an Ordovician Ocloyic west-vergent, thin-skinned thrust belt in the Argentine Precordillera, southern central Andes. In: Pankhurst, R.J., Veiga, G.D. (eds.). Gondwana 12: Geological and Biological Heritage of Gondwana, Academia Nacional de Ciencias,
Córdoba, Argentina, Abstracts, p. 350.


Thomas. W.A., Astini, R.A., Denison, R.E., 2001. Strontium Isotopes, Age, and Tectonic Setting of Cambrian Salinas along the Rift and Transform Margins of the Argentine Precordillera and Southern Laurentia. Journal of Geology, 109, 231-246.

Thomas, W.A., Astini, R.A., Bayona, G., 2002. Ordovician collision of the Argentina Precordillera with Gondwana, independent of Laurentian Taconic orogeny. Tectonophysics, 345, 131-152.


Thomas. W.A.. Astini, R.A., Mueller, P. A., Gehrels, G and Wooden, J., 2004. Transfer of the Argentina Precordillera terrane from Laurentia: Constraints from detritial-zircon geochronology. Geology, 32, 11: 965-968.


Thomas, W.A., Tucker, R.D., Astini, R.D., 2000. Rifting of the Argentine Precordillera from southern Laurentia: palinspastic restoration of basement provinces. Geological Society of America, Abstracts with Programs, 32(7), A-505.


Tollo, R.P., Corriveau, L., McLelland, J., Bartholomew, M.J., 2004. Proterozoic tectonic evolution of the Grenville orogen in North America: An introduction. In: Tollo, R.P., Corriveau, L., McLelland, J., Bartholomew, M.J. (eds.). Proterozoic evolution of the Grenville orogen in North America. Boulder, Colorado, Geological Society of America Memoir, 197, 1-18.

Tohvera, E. et al., 2002. Paleogeography of the Amazon craton at 1.2Ga: early Grenvillian collision with the Llano segment of Laurentia. Earth and Planetary Science Letters 199 (2002) 185-200


Vaccari, N.E., 1988. Primer hallazgo de trilobites del Cambrico inferior en la provencia de La Rioja (Precordillera septentrional). Asociacion Geologica Argentina Revista, 43(4), 558-561.


Vaccari, N.E., 1994. Las faunas de trilobites de las sucesiones carbonáticas del Cámbrico y Ordovícico temprano de la Precordillera
Precordillera Septentrional Argentina. Doctoral thesis. Universidad
Nacional de Córdoba, Argentina, 271 pp.


Vaccari, N.E., Bordonaro, O., 1993. Trilobites en los olistolitos cámbricos de la Formación Los Sombreros (Ordovícico), Precordillera de San Juan, Argentina. Ameghiniana, 30, 383-393.

Varela, R., Dalla Salda. L.H., 1992. Geocronología Rb—Sr de metamorfitas y granitoides del tercio sur de la Sierra de Pie de Palo, San Juan, Argentina. Revista de la Asociación Geológica Argentina 47, 271—275.

 

Varela, R., López de Luchi, M., Cingolani, C., Dalia Salda. L., 1996. Geocronología de gneises y granitoides de la Sierra de Umango, La Rioja: Implicancias tectónicas. XIII Congreso Geológico Argentino Argentino y III Congreso de Exploración de Hidrocarburos, Actas, vol. 3. pp. 5 19—527.

Veevers, J.J., 2003. Pan-African is Pan-Gondwanaland: Oblique convergence drives rotation during 650-500 Ma assembly. Geology, 31(6), 501-504.


Von Gosen, W., Buggisch, W., Lehnert, O., 1995. Evolution of the Early Paleozoic mélange at the eastern margin of the Argentine Precordillera. Journal of South American Earth Sciences, 8(3-4), 405-424.

von Gosen, W., Prozzi, C.. 1998. Structural evolution of the Sierra de San Luis (Eastern Sierras Pampeanas, Argentina): lmplications for the Proto—Andean Margin of Gondwana. ln: Pankhurst, R.J., Rapela. C.W. (Eds.), The Proto—Andean Margin of Gondwana. Geological Society London, Special Publication, 142, 235—258.

Vujovich, G., Kay, S.M., 1996. Evidencias geoquímicas del origen y ambiente geológico de las rocas metamórficas de composición máfica a intermedia de las Sierras Pampeanas Occidentales. XIII Congreso Geológico Argentino y III Congreso de Exploración de Hidrocarburos. Actas, vol. 5, pp. 273—291.

Vujovich, G.I., Kay, S.M., 1998. A Laurentian? Grenville-age oceanic arc/ back-arc terrane in the Sierra de Pie de Palo, Western Sierras Pampeanas, Argentina, In: Pankhurst, R.J., Rapela, C.W. (Eds.), The Proto—Andean Margin of Gondwana. Geological Society London, Special Puhlication, 142, 159—179.

 

Vujovich, G.I., Ramos, V.A., 1994. La faja de Angaco y su relación con las Sierras Pampeanas Occidentales. VII Congreso Geológico Chileno (Concepción, Chile), Actas, vol. 1, pp. 215—219.

Wareham, C.D., Pankhurst, R.J., Thomas, R.J., Storey, B.C., Grantham, G.H., Jacobs, J., Eglington, B.M., 1998. Pb, Nd, and Sr Isotope Mapping of Grenville- Age Crustal Provinces in Rodinia. The Journal of Geology, 106, 647-659.

Webby, B.D., Percival, I.G., Edgecombe, G.D., Cooper R.A., Vandenberg, A.H.M., Pickett, J.W., Pojeta, J., Jr., Playford, G., Winchester-Seeto, T., Young, G.C., Zhen, Y.-Y., Nicoll, R.S., Ross, J.R.P., Schallreuter, R., 2000. Ordovician palaeobiogeography of Australasia. Memoir of the Association of Australasian Palaeontologists No. 23, 63-126.

Wright, J. E. and Wyld, S. J. 2003. Appalachian, Gondwanan, Cordilleran interactions: A new geodynamic model for the Paleozoic tectonic evolution of the North American Cordillera. Paper No. 229-12. 2003 Seattle Annual Meeting (November 2–5, 2003) Geological Society of America Abstracts with Programs, Vol. 35, No. 6, September 2003, p. 557

volver arriba

 

 

 

PATAGONIA

Hipótesis de Ramos (1984) sobre aloctonia de Patgonia y configuración tectónica del sector norte de Patagonia

interaccion patagonia Gondwana.gif (42809 bytes) Hipotesis ramos.gif (38077 bytes) Estructura trascontinental.gif (109491 bytes) Estructura trascontinental 1.gif (125376 bytes) arcos norte Somoncura.gif (38197 bytes) Colision patagonia Gondwana.gif (42477 bytes)
Interacción TerrenoPanthalassan-Gondwana Hipótesis interacción Patagonia Gondwana Estructura transcontinental límite Patagonia Gondwana Estructura transcontinental límite Patagonia Gondwana Arcos del N de Patagonia relacionados con la interacción Patagonia Gondwana Interacción final entre  Patagonia Gondwana

Las figuras de arriba fueron tomadas de los siguientes trabajos, donde se expone la hipótesis de Ramos sobre la aloctonía de la Patagonia

Ramos, V., 1984. Patagonia: ¿Un continente Paleozoico a la deriva?. IX Congreso Geológico Argentino. Actas III: 311-325

Ramos, V., Cortes, J.M., 1984. Estructura e interpretación tectónica. In: Ramos, V. (Ed.), Relatorio de la geología y recursos naturales de la provincia de Río Negro. 9 Congreso Geológico Argentino I, pp. 317–346.

Ramos, V. A., 1988. Late Proterozoic-Early Paleozoic of South America-A collisional history. Episodes, 11 (3): 168-174.

volver arriba


 

Hipótesis de Pankhurst et al 2006

 

 

Hipótesis de Gregori-Kostadinoff

FIGURA 1 FIGURA 2 FIGURA 3 FIGURA 4 FIGURA 5 FIGURA 6
FIGURA 7 FIGURA 7B FIGURA 8A FIGURA 8B FIGURA 9 FIGURA 10
FIGURA 11 FIGURA  12A FIGURA 12B      

FIGURAS TOMADAS DE:

Kostadinoff, J., Gregori, D. A. and Raniolo, A., 2005. Configuración geofísica-geológica del sector Norte de la provincia de Río Negro. Revista de la Asociación Geológica Argentina. LX: 368-376).

Daniel A. Gregori , José Kostadinoff, Leonardo Strazzere, Ariel Raniolo, 2008. Tectonic significance and consequences of the Gondwanide orogeny in northern Patagonia, Argentina, Gondwana Research, 14: 429–450

4. Results
Three areas with positive gravimetric anomalies, four with negative anomalies and several features characterized by steep gravity gradients (Figs 7A and 8A, B) were recognized during this study.


4.1. Positive anomalies
4.1.1. Chimpay–Gobernador Duval High
Residual anomalies within the range −27 to −14 mGal, a relatively positive anomaly occur from Chimpay to Gobernador Duval and Puelén (Fig. 7). Refraction shots by the YPF (Yapeyú to RN5, Fig. 6A) indicate a thickness of more than 1000 m for the Mesozoic Neuquén Basin near Gobernador Duval. In the Chimpay–La Japonesa area this basin reaches a maximum thickness of 300 m. In order to satisfy the observed anomalies a density contrast of +0.32 g/cm3 between the Mesozoic sedimentary rocks and those of the basement (Table 1) is required.
No outcrops of basement rocks with such high density weredetected in the Neuquén Basin. However, Grenvillian (~1.0 Ga) basement composed of tonalite and trondhjemite were studied by Sato et al. (2000) in the Las Matras area (Fig 1), which is located 130 km north of Gobernador Duval. We interpret the Chimpay–Gobernador Duval High to be composed of basicintermediate rocks (seismic velocity 5400–6600 m/s) covered by sedimentary rocks deposited in the Jurassic–Cretaceous Neuquén Basin. Such igneous rocks could also explain positive magnetic anomalies, which range between 250 and 175 nT (Fig. 9) in this area.


4.1.2. Río Colorado High
Several relatively positive residual gravity anomalies (6, −8 and −11mGal) appears between PichíMahuida, Cerro LosViejos and Río Colorado (Kostadinoff and Llambías (2002) (Fig. 7A). These anomalies are considered to reflect Pampean basement, which is composed of small outcrops of amphibolite, gneiss, micaceous schist, phyllite and metasandstone of the Las Piedras Metamorphic Complex and granite and granodiorite of the Pichí Mahuida Group (Linares et al., 1980; Tickyj et al., 2002). The anomalies are due to relative uplift of blocks similar to those that occur in the Sierras Pampeanas de San Luis y Córdoba. TheNW– SE margin of RíoColoradoHigh is coincidentwith a steep gravity gradient (near 1 mGal/km) between Cerro Los Viejos and 38° S (Figs 7A and 8A, B). Fig. 7B shows the measured and the modeled anomalies in the Río Colorado High considering a density of 2.65 g/cm3 for the basement rocks. According to the geometries of the anomalies and the outcrops, the Río Colorado High is believed to represent a block-faulted system.
Refraction shots by YPF (numbers 20 to 25, Fig.6B) show the existence of basement rocks with high seismic velocity (5000 to 6000 m/s), which are interpreted as the Las Piedras Complex. This basement deepens to the south, reaching 700 m below the surface near the Huincul Fault zone. Magnetic anomalies are relatively low in this area and suggest the absence of paramagnetic rocks in the basement (Fig. 9).


4.1.3. Choele Choel High
The Choele Choel High (Kostadinoff et al., 2005) is a positive residual gravity anomaly of +10 and +15 mGal (Fig. 7A and B) with steep (2 mGal/km) gradients (Fig. 8A, B). The northern boundary is defined by the Huincul Fault Zone and lineament v“A” (Kostadinoff et al. 2005). The southern boundary (Figs 7A, B, 8 and 10) is a complex, arcuate system that is formed by 3 lineaments, which are defined by steep (1–2 mGal/km) gravity gradients (B, C and E: Fig. 10). The model of Fig. 7B requires a slighter increase in the basement rocks density to explain the
measured anomaly and as showed in this figure the Choele Choel High represent an uplifted block.
Refraction shots by YPF in this area (Puesto Segatore to profile number 8, Fig. 6C) indicate the existence of rocks with seismic velocities of 4500 to 5320 m/s which are covered by Mesozoic–Cenozoic rocks. Such rocks exist at depths between 200 and 600 m in the eastern and southeastern parts of the Choele Choel High (Profiles 8 to RN19) and to a depth of 1000 m when approaching lineament “A”.

Candidates that fulfil such seismic velocity are the phyllite, muscovite-bearing schist, and metasandstone of the Las Piedras Complex that are exposed in the Río ColoradoHigh. The seismic velocities may also be explained by phyllite, metasandstone, marble and pelite of the Nahuel Niyeu and El Jaguelito formations (Fig. 3), which are temporally correlated with the Las Piedras Complex, and crop out in the Nahuel Niyeu–Valcheta area (Caminos and Llambías, 1984).
If ultramafic rocks were emplaced in a Pampean basement of this density (Table 1) an anomaly of more than 100 mGals could be expected. Most ultramafic belts in suture zones show values between +50 and +120 mGals (Gibb and Thomas, 1976, Kunar et al., 2004, Williams et al., 2006). However, this region does not display anomalies of this magnitude and the seismic profiles also do not suggest the existence of rocks with seismic velocities
sufficient to be sourced from ultramafic rocks.
The magnetic survey (Fig. 9) shows scarce magnetic anomalies, with maximum values of +75 nTand minimum values of−75 nT. These data also preclude the existence of mafic– ultramafic rocks in this area and are more consistent with a magnetic source such as the non-magnetic rocks such as Pampean low-grade metamorphic rocks.


4.1.4. Viedma High
Gravity anomalies from the Viedma High (Figs. 7A and 10) vary between 1 and 15 mGal, (Figs. 7A and 10), suggesting a density contrast caused by rocks with similar densities to that which caused the Choele Choel High. The contact between both of these gravimetric highs is defined by NW–SE and NE–SW lineaments that follow narrow belts with steep gravity gradients (Figs 7A, 8 and 10). Unfortunately, there are no seismic profiles
located in this area, but outcrops in the nearby Salinas del Gualicho area (Fig. 3) belong to the Nahuel Niyeu Formation (Figs. 3 and 10). Minor outcrops of the Sierra Grande Formation and Gondwana igneous rocks are also present. Drill holes at O'Connor and Elvira (Kaasschieter, 1965) intersected lowgrade metamorphic rocks to a depth between 400 and 475 m. (Fig. 7).


The positive gravity anomalies associated with the Viedma High extend northwards (Kostadinoff and Labudía, 1991 and Kostadinoff, 1992) and link with positive gravity anomalies located to the west of Sierras Australes (Figs. 2, 3, 7A and 10).
The Algarrobo drill hole, and refraction data near Bahía Blanca (Estancia El Recuerdo and Puerto Belgrano, Bonorino et al, 1987), indicates seismic velocities similar to those that are interpreted to be the source of the Choele Choel High (i.e. lowgrade metamorphic rocks of the Las Piedras Metamorphic Complex).

The distribution of positive anomalies in the Viedma High and in the Choele Choel High supports the interpretation that a belt of Pampean rocks extends from the southern part of the Buenos Aires province to northern Patagonia (Figs. 3 and 10).


4.2. Negative anomalies

4.2.1. Neuquén Basin
The Mesozoic Neuquén Basin (Figs 2, 7A and 10) is characterized by gravity lows between −54 mGal and −32 mGal. Seismic data (Fig. 6D) and drill hole information indicate that the sedimentary sequences reach an approximate thickness of 3000 m. Negative gravity anomalies (−41 mGal) near Chelforó represent small depocenters, whose thicknesses decrease to 750 m near Chimpay (Fig. 7A). Magnetic anomalies in this area
vary between −125 and +125 nT (Fig. 9) and are attributed to the presence of Fe-oxides in the Cretaceous Neuquén Group (400 m thick) that covers this portion of the basin.


4.2.2. Carapacha Basin
The scarce outcrops of the Permian continental Carapacha Basin and the volcanic rocks near Puelches in La Pampa province (Fig. 4) are coincident with negative gravity anomalies (−40 to −30 mGal: Fig. 7A). The density contrast between these rocks and those of the Pampean basement is relatively low (Table 1 and Fig. 7B), which led to Kostadinoff et al. (2001) to estimate a thickness of 2700 m for this basin.

This is substantially more than the 900 m thickness estimated by Melchor (1999). Based on these results, Kostadinoff and Llambías (2002) considered the subsurface extent of the basin to be greater than that suggested its surface extent.
Therefore, the gravimetric minimums of −41, −31 and−40 mGal located northeast of Gobernador Duval, as well as those located west of Cerro Los Viejos and Pichí Mahuida (−41 and −36 mGal), possibly represent an extension of the Carapacha Basin (Figs. 7A, B and 10). In the Carapacha Basin, the Gondwana volcanic rocks unconformably overlying the Paleozoic sedimentary sequences in the Carapacha Basin and are interpreted
to contribute to the negative anomalies. The magnetic anomalies (−275, −225 and −125 nT: Fig. 9) are due to the low susceptibility of the Paleozoic sedimentary rocks and the Gondwana acidic volcanic rocks.


4.2.3. Estancia El Caldén Basin

A relatively small negative residual gravity anomaly (−24 mGal: Fig. 7A) is located near the Estancia El Caldén
(Kostadinoff et al., 2005), north of the Huincul Fault Zone. A density contrast of −0.2 g/cm3 between the Pampean basement and the Mesozoic sedimentary units suggests a sedimentary thickness of 1700 m (Figs. 7A and 10). Refraction data in this area (Fig. 6E) indicate a greater than 1000 m thick sedimentary
pile. The seismic data suggest velocities consistent with Lower Cretaceous and Upper Tertiary sedimentary successions, which cover the Paleozoic rocks.


4.2.4. Ramos Mexías anomaly
The Ramos Mexías anomaly is part of a continental-scale anomaly (500 km in diameter), located in the south-western part of the North Patagonian Massif (Blitzkow et al., 1994). The Ramos Mexías anomaly corresponds to a large area of several negative gravity anomalies (−46 mGal to −35 mGal, Figs. 7A and 10) located south of the lineament “B) (Fig. 10). Outcrops include Pampean metamorphic rocks and Gondwana intrusive and volcanic rocks. The metamorphic rocks belong to the Nahuel Niyeu Formation and include micaceous and amphibolite schist, marble, and amphibolites. The intrusive rocks (minor tonalite, granodiorite and granite) belong to the Gondwana Navarrete Plutonic Complex. The volcanic rocks (acidic ignimbrite) are part of the Treneta Volcanic Complex (Caminos et al.,
2001) and the Triassic volcaniclastic Los Menucos Complex (Cucchi et al., 1999).

4.2.5. Lagunas Dulces anomaly
The Lagunas Dulces anomaly is a negative gravity anomaly (−20 mGal) that is located north-eastwards of San
Antonio Oeste (Figs. 7A and 10). This anomaly is interpreted to be caused by the Lagunas Dulces Granite (Kostadinoff, 1992). A drill hole at this locality intersected granitic rocks (249±10 Ma: Kaasschieter, 1965) at a depth of 370 m. Because the geometry and magnitude of this anomaly is similar to those coincident with outcrops of the Gondwana Navarrete Granodiorite, we interpret the Lagunas Dulces anomaly to be caused by intrusion of Gondwana granitic bodies in the Pampean basement.


4.2.6. Bajo de Valcheta anomaly
The Bajo de Valcheta anomaly is a negative anomaly of the order of −10 mGal, located between the Viedma High and the Ramos Mexías anomaly. The area is coincident with outcrops of the low-grade metamorphic rocks of the Nahuel Niyeu Formation (Figs. 7A,B and 10). According to the model in Fig. 7B a density of 2.65 g/cm3 is required to satisfy the observed anomalies. No significant magnetic anomalies (Fig. 9) can be recognized at this
location.

4.2.7. Laguna del Zorro anomaly
The Laguna del Zorro gravity anomaly (Figs. 7A and 10) is located to the north-west of the Viedma High and is bounded by lineament “C” (Fig. 10). The geometry and magnitude of the anomaly is similar to the Lagunas Dulces anomaly and is therefore interpreted to be sourced from the same or similar Gondwana intrusive body.

 


4.3. Belts of steep gravity gradients
In order to validate the continuity of the Huincul Fault zone and other structural lineaments proposed by Turner and Baldis (1978), we focused on areas with steep gravity gradients in the gravity anomaly map. Such steep gradients are usually interpreted as representing fault structures (Nettleton, 1976), boundaries of rocks with contrasting petrophysical characteristics or sutures (Gibb and Thomas, 1976, Wellman 1978, Coward and Fairhead
1980, Gibb et al. 1983). Differences in density or thickness between blocks are often responsible for steep gradients (Gibb and Thomas, 1976, Thomas, 1983, Kunar et al. 2004, Williams et al., 2006). Values between 0.5 and 2 mGal/km, were obtained using the gradient and second derivate operator. A 3D image generated using Surfer 8™ (Fig. 8B), show several lineaments defined by belts with steep gravity gradients.
Table 2 indicates the name, location, gradient, any surface evidence and possible age of each lineament, while their location is shown in Fig. 10.

Three main lineament orientations directions can be recognized: lineaments A and C are NE–SW trending; lineaments B, D and E are NW–SE trending; and lineaments F, G and the Huincul Fault zone are E–W trending.
The E–W lineaments exhibit more complex geometries than the other lineaments. Most lineaments form the boundary of the Choele Choel High (e.g. A, B, C, E and the Huincul Fault zone).


4.3.1. Huincul Fault zone
Because the Huincul Fault zone (Kostadinoff et al., 2005)
crops out in the Neuquén Basin (Windhausen, 1914 and Keidel, 1925), its configuration and kinematics are relatively well constrained. The Huincul Fault Zone displays dextral strike-slip movement (Orchuela and Ploszkiewicz, 1984) and a zig-zag along strike pattern (Figs. 2 and 10). This pattern suggests that both transtensional and transpressional conditions occurred during fault movement, resulting in the contemporaneous development of positive and negative flower structures. This fault system was active from Upper Paleozoic–Triassic times, as recorded by continental depocenters associated with the transtensional subsidence (Vergani et al. 1995).
Its eastwards continuity was suggested by Turner and Baldis (1978), and established by Orchuela and Ploszkiewicz (1984) based on interpretation of seismic reflection data. The high gravimetric gradients between Neuquén and Cubanea (Figs. 7, 8B and 10) support the continuity of the Huincul Fault Zone until 63° 40´W, with maximum gradients (2 m Gal/km) between the Neuquén Basin and 64° 45´W–39° 20´S.
The section between Neuquén and Chimpay shows a smooth releasing bend (Fig. 10), that changes to a restraining bend in the Chimpay 63° 40´W section. The gravity anomalies are discontinuous along the fault zone. In the area between Chimpay and Cubanea, the differences in magnitude of the anomaly between the Choele Choel High and the Carapacha Basin and the Río Colorado High mark the trace of this fault zone. In the section between Chimpay and Neuquén, such differences in the anomalies are less evident, but high gravimetric gradients (2 mGal/km) are interpreted to reflect the location of the fault zone.


Magnetic anomalies are diffusely coincident with the interpreted fault zone between 5 and 10 km north of Chichinales and Choele Choel. These anomalies are low-amplitude (175 to 75 nT, Fig. 9) and are interpreted to be related to the Albian– Coniacian red beds of the Neuquén Group, which contain abundant Fe-oxides (hematite). The sediments have a magnetic susceptibility of 0.006500 SI (Telford et al., 1990), which is
sufficient to explain the magnitude of the magnetic anomalies.
These low-magnitude anomalies do not compared magnetic anomalies typically observed in sutures zones because suture zones are generally composed by serpentinite zones associated with sheared peridotites, which have a magnetic response N500 nT (Ferraccioli et al., 2002).

The Huincul Fault zone is therefore unlikely to be a suture. In the Neuquén–Chichinales section (Fig. 10) the
Neuquén Group is 1000 m thick, but in the cliffs located south of the Rio Negro, only the upper section of this unit crop outs. In contrast, the area north of the fault zone, Upper Cretaceous–Lower Tertiary marine Malargue Group crop out, suggesting more than 500 m of vertical displacement across a fault zone.

Due to the outline of the Huincul Fault Zone, the section between Plaza Huincul and Picún Leufú (Fig. 10) and between Neuquén and Chimpay represent domains of dextral transpressive shear (Orchuela and Ploszkiewicz, 1984), where basement rocks are exposed in the fault zone. In the section between
Chimpay and 64° 15´W, deformation appears to be transtensional based on the existence of small depocenters associated to the fault zone.


5. Mylonitic belts and lineaments with steep gravimetric and high magnetic gradients in northern Patagonia, southern La Pampa and Buenos Aires provinces
Northern Patagonia is characterized by the presence of several mylonitic belts and regional structural lineaments, whose geometries, deformational mechanisms and age mostly remains unclear (Fig. 11). The better-known of these lineaments and mylonites include:

a) Pangaré and La Seña mylonites (Gregori et al., 2000), are formed by two belts, each 1 km wide, 15 km long, and comprising granitic augen-mylonites, protomylonites, mylonites and ultramylonites, (Figs 10, 11 and 12A) with strikes between 310° and 330°. They exhibit sinistral movement associated with an east–west compression

b) Peynecura mylonitic belt (Llambías et al., 2002) is more than 500mwide, 15 kmlong, and has aNNW–SSE strike (Fig. 11). The deformation is interpreted to have occurred during the Early Triassic because is related to foliated granites of such age. No information about strain orientation is available.

c) The El Jaguelito Fault zone is a 50 km long NW–SE dextral lineament that appears in the eastern area of the North Patagonian Massif (Fig. 11). According to Ramos and Cortés (1984), deformation occurred during Upper Paleozoic times.

d) Peñas Blancas and La Laguna mylonites are described by Giacosa (1996). They are 10 km long, and are temporal related to the foliated Permian-aged Peñas Blancas and La Laguna Granites. Four more NW–SE mylonitic belts were recognized during regional mapping in this area, but information on strain orientation is unavailable. The timing of deformation is interpreted to be Gondwanan because the foliated granites were intruded at the same time as mylonitization occur.

e) The Nahuel Niyeu lineament system is a major structure that is 25 km wide, and extends for 65 km long (Figs. 10 and 11) between the Salinas Trapalcó–Laguna Curicó Lineament (Figs. 1, 2 and 12A) and the Somoncura Plateau (Fig. 2). Nahuel Niyeu lineament includes the Arroyo Seco, Barda de Lucho, Arroyo Salado, Laguna Negra, Quiroga and Ramos lineaments, the Tardugno, Musters and Huanteleo faults and the Nahuel Niyeu, Railer and Rana thrust sheets (Fig. 10).

In the La Pampa province, a mylonitic belt was described in Cerro de los Viejos (Figs. 11 and 12A) by Tickyj et al. (1997). A further four belts of steep gravity gradients which were interpreted as major structures were described by Kostadinoff et al. (2001).
These four belts include:
a) The Quehué lineaments is a 350 km long, 15 km wide. It is defined in the gravity data by a 2 mGal/km gradient which defines a structure interpreted to represent the boundaries of the Triassic Quehué basin.

b) The Valle Daza-Cuchillo Có lineament is a 310–320° trending belt that is 100 kmlong and characterised by a steep gradient in gravimetry data. This lineament is interpreted to represent the western boundary of the Rio Colorado High (Figs. 11 and 12A).

c) The Cerro los Viejos lineament extends in a NW-direction for more than 100 km from Cerro Los Viejos and is possibly related to the Cerro Los Viejos mylonite. According to Kostadinoff and Llambías (2002), the Cerro los Viejos lineament represents the contact between the Permian Carapacha Basin and the Río Colorado High. This lineament and its associated steep gravity gradient disappears in the central part of the La Pampa province, but is identified near Algarrobo del Aguila (Fig. 11) as a 50 km long, 15 km wide lineament with steep gravity gradient. This NW-trending structure is considered to be the boundary of a Triassic basin (Kostadinoff and Llambías, 2002).

d) The Macachín Rift includes Paleozoic and Mesozoic sedimentary sequences deposited in a NNW–SSE extensional basin (Figs. 11 and 12A). Outcrops of igneous and sedimentary Gondwanan rocks in the central part of the La Pampa province (Figs 2 and 4) follow WNWor NW trends, which suggest structural control.

In the Buenos Aires Province strike-slip faults located north and south have been used to explain the structure of the Sierras Australes. For example, Newton and Cingolani (1990) showed a series of WNW–ESE trending sinistral strike-slip faults near Pigue and Cerros Colorados (Figs. 11 and 12A), which are evident in gravity data (Alvarez, 2004).

In the Bahía Blanca area, Bonorino et al. (1987) showed the occurrence of several lineaments with steep gravity gradients south of Sierras Australes. Lineaments of importance include the Salitral de la Vidriera and Nueva Roma belts, which are strike-slip fault zones longer than 100 km in strike extent (Fig. 11).


6. Gondwana deformation in the Neuquén Basin, northern Patagonia and southern La Pampa and Buenos Aires provinces

Windhausen (1914), Keidel (1925) and Groeber (1929) described NE–SW structures transverse to the N–S Andean
Orogeny in the Neuquén Basin (Figs. 11 and 12A). According to these authors, as well as Herrero Ducloux (1946) and De Ferraris (1947) the stratigraphic nd structural evidences suggest folding during the Cenomanian or the Middle Jurassic.
They observed that the structures mimic the northwestern and northern border of the North Patagonian Massif and that they possibly formed during NW- and N-directed lateral translation of the massif. These structures were grouped in the “Dorsal de Huincul” (Huincul Fault zone) by De Ferraris (1947).
In the North Patagonian Massif, the Gondwana deformation is recorded by the Pangaré and La Seña mylonites. Clearly defined δ and σ type porphyroclasts of K-feldspar, and shaped lenses of Qz–K-feldspar, support sinistral movement. These belts and the associated foliated granite (Gregori et al., 2000) indicate maximum compression (σ1) in a WNW–ESE direction (Fig. 12A) during Middle to Upper Triassic (Saini-Eidukat
et al., in press).
In the El Jaguelito Fault zone, the fault pattern suggests a dextral transcurrence along N 310° (Fig. 12A) during Upper Paleozoic times (Ramos and Cortés, 1984). SE, S and SW vergence occurred along the Nahuel Niyeu lineament system (Fig. 12A), each of which was related to three phases of deformation, with principle shortening directions in NE–SW, NW–SE and E–W directions during Permo-Triassic times (von Gosen,
2003).
In the Sierra Grande area, the Pampean El Jaguelito Formation contains evidence which supports a D1 event that was generated byNW–SE or E–W shortening (Giacosa and Paredes, 2001, von Gosen, 2002). In the same area, the Famatinian Sierra Grande Formation is folded around open synclines and anticlines that have N–S trending axial traces (Fig. 12A). The Permian Laguna Medina granite intrudes the Sierra Grande Formation and records
a ductile E–W shortening deformation event, which was coincident with deformation of the Sierra Grande Formation (Rossello et al., 1997, von Gosen, 2002).
Transport directions from NE to SW during the Gondwana deformation was accommodated along mylonites within the Peñas Blancas and La Laguna Granites (Giacosa, 2001, von
Gosen, 2002).
In the North Patagonian Massif, the Sierra Grande Formation, the Laguna Medina Granite and foliated granites and mylonites in the Estancias La Seña and Pangaré area all indicate shortening in aWNW–ESE or E–Wdirection (Fig. 12A). The mylonites in the Peñas Blancas and La Laguna Granites and thrust sheets and faults in the Nahuel Niyeu area are related to a N–S or SW–NE shortening (Fig. 12A) (Giacosa, 2001, von Gosen, 2002).

In the La Pampa Province, the Carapacha Basin (Figs. 4, 11 and 12A) exhibits a NW–SE elongation, suggesting basin development occurred during NE–SWdirection extension. Folding of the sedimentary successions and inversion in the Carapacha Basin (Melchor, 1999) and mylonitization of the Cerro de los Viejos granite (Tickyj et al., 1997) dated as Late Permian (254±2 Ma), suggests that both areas were affected by the same deformational
phase (Fig. 12A) related to a NE–SW shortening.

In the Sierras Australes of Buenos Aires Province, the Pillahuincó Group (Carboniferous-Permian) was deformed
(Fig. 12A) between the late Early Permian and the early Late Permian (Tomezzoli and Vilas, 1996). Deformation generated a sigmoidally-shaped mountain chain. Folding and overthrusting occurred during NE vergence (von Gosen et al., 1990). This deformation was related to sinistral transpressive deformation (Sellés Martinez, 1989) along theWNW–ESE Pigué transcurrent fault system (Alvarez, 2004) and the Salitral de la Vidriera and
Nueva Roma transcurrent fault systems (Bonorino et al., 1987).
The deformation was interpreted to be related to transpressive movement of the “Patagonian block” relative to the South American craton during Permo-Triassic time (Martinez, 1980), although little supporting evidence was provided. Sellés Martínez (1989) considered the deformation in Sierras Australes to be due to a transpressive regime between the Patagonian block and the South American craton during the Late Permian oblique
subduction.
Detailed structural studies in the Sierras Australes define a sigmoidal belt, with fold trends rotating N60° clockwise, from WNW to NNW, a configuration attributable to dextral transpression (Cobbold et al., 1986, 1991, 1992). Inferred directions of principle shortening are oblique to the orogenic belt, and to the convergent southern margin of Gondwana and are interpreted to result from oblique (dextral) subduction beneath an Andean-type margin (Cobbold et al., 1992, Craddock et al., 1998).
Consequently, the sedimentary basins, the emplacement of the Gondwana intrusive and volcanic rocks and the regional structure north of the Colorado and Negro rivers are related to a first SW–NE extension (Fig. 4), which later switched to a shortening regime in a NE–SW or N–S direction (Fig. 12A).

In their structural analysis of northern Patagonia, Turner and Baldis (1978) identified several NW–SE lineaments in the area between Bajo del Gualicho and Golfo San Matías (Fig. 2). They suggested variable compressive stresses in the North Patagonian Massif including: west-shortening directed (Río Limay area); northeast-directed shortening (Salinas del Gualicho); eastdirected shortening (Sierra Grande); and southwest-directed shortening (El Cuy). The variation in stress directions is attributed to counterclockwise rotation of the North Patagonian
Massif (Fig. 12A).


7. Deformation in the southern area of the Gondwana supercontinent
The first detailed reconstruction of the relative position of Africa and South America during the Upper Paleozoic was given by du Toit (1927). The continuity of an orogenic system was extended from South America to Australia, passing just south of the Permo-Triassic fold belts, introducing the name of Samfrau Orogenic Zone (du Toit, 1937). This belt is nearly equivalent to what we understand today as the Gondwanide Orogenic Belt (du Toit, 1937).
In the South African segment, the arrangement of the Gondwanide Orogenic Belt and the associated basins is strongly conditioned by the configuration of older basements blocks (Kaapvaal, ≥2.5 Ga, Namaqua-Natal, ≈1 Ga and Gariep, Saldania and Mozambique, 1–0.5 Ga). The Ordovician–Devonian Cape Basin follows a main W–E trend along the Saldania province with branches to the NWand NE. The Cape Fold Belt includes southern and western tectonic domains (Söhnge, 1983), separated by the Cape Syntaxis (Fig. 12b), across which the structural strike, and the intensity of deformation changes markedly (de Beer, 1995). The southern domain extends from the eastern margin of the Cape Syntaxis to Port Elizabeth and is characterized by east–west-trending, north-verging structures including bedding- parallel thrusts and north-verging folds (Fig. 12B).

No evidence for dextral shearing was found in this domain. In the western domain, strata is mildly deformed (de Beer, 1992). Folds and faults trend northwest to north (de Beer, 1990). North-trending slickensides developed on thrust planes (Ransome and de Wit, 1992) and stretching lineations (Cobbold et al., 1992) indicate a component of strike-parallel displacement.
The pattern of folds, faults and lineations are consistent with overall dextral strike-slip deformation (Cobbold et al., 1992). The Cape Syntaxis is a northeast-trending structural domain that separates the southern and western domains (de Beer, 1995). Based on a reconnaissance of the structure of the syntaxis and adjacent domains, Ransome and de Wit (1992) proposed a model in which the continental crust to the west of Cape Syntaxis was rotated clockwise during syntaxis formation, while crust immediately east is rotated counter-clockwise.

In the Malvinas (Falkland) Islands, Late Permian fold and thrust belt structures form the eastward continuation of the Cape Fold Belt when restored into a pre-Gondwana break-up configuration.
Detailed structural studies by Curtis and Hyam (1998) and Hyam (1998) suggest that these structures developed
in a complex strain system of combined NW–SE shortening and dextral shear.

In Antarctica, the Gondwanide Orogen extends through the Ellsworth–Whitmore Mountain block and the Pensacola Mountains (Curtis and Storey, 1996). During Gondwanide orogenesis the Ellsworth–Whitmore Mountain block is assumed to have occupied a site intermediate between the more easterly Falkland Islands and the more westerly Pensacola Mountains (Curtis and Storey, 1996). Structural analyses (fold orientation, shortening directions and strain partitioning) indicate contemporaneous development of gently plunging foreland-verging folds and
asymmetric, steeply plunging folds symptomatic of deformation within a highly oblique dextral transpressive belt (Curtis, 1997, 1998, Craddock et al., 1998).
Dextral transpressional structures are evident throughout much of the Gondwanide Belt, suggesting a component of dextral transpressional during the evolution of this orogenic belt. Johnston (2000) indicated that the east–west-trending domain of the Cape Fold Belt ends to the west and east against the northwest-trending South American and Antarctic portions of the Gondwanide Orogen respectively.


8. Tectonic model
The diverse shortening directions and lateral crustal translations during Gondwana time north and south of the interpreted boundary between northern Patagonia and Gondwana South America cannot be explained as the result of homogeneous stress regime (i.e. NE–SW compression due to the subduction and collision of Patagonia with the southern margin of Gondwana South America) as speculated by Ramos (1984), and von Gosen (2002, 2003).
All lineaments plotted in Fig. 11 (those both related and unrelated to mylonites, and the belts with steep gravity and magnetic gradients, presumably active during Gondwana times) suggest a lineament network that defines a block system (Fig. 12A).
This collage of continental crustal blocks and its movement was the result of a tectonic event that produced crustal fragmentation.
The geometry of individual blocks controlled the blockmovement and therefore the development of mylonite belts and other lineaments. In order to explain the variable stress regime in which zone of compressional and extensional tectonism were active at the same time over such a large area, we require movement of crustal
blocks along arcuate strike-slip faults (Fig. 12A).
This style of deformation is common during collisional orogeny, where frontal zones are deformed during indentor
tectonics, where the lateral terminations of the indentor involve extensive escape tectonics (Ratschbacher et al., 1991). The size and shape of the continental fragments involved, the orientation of the indentor, and the collision vector, influence the extent and geometry of indentation and later escape.
Modern examples of large-scale indentation-escape orogens are found in the Himalaya-Alpine belt. These models include the collision of the Arabian and the European plates (west-directed escape) and the east- and southeast-directed escape of southeast Asia following the collision of India and Asia (Tapponnier and Molnar, 1976). The transcurrent shear zones separating elongate crustal slices are the most important large-scale structural features
in these Himalaya–Alpine models.
Indentation-escape tectonics is widely observed and accepted in younger orogens, but has not been recognized to the same extent in older mountain belts, probably because the structures associated with such a tectonic regime are less obvious. In addition, subsequent orogenic eventsmay obscure the presence of an ancient indentation-escape regime (Ratschbacher et al., 1991).
Several examples of Precambrian indentation-escape tectonics have been described. In southern Africa, the proto-Kalahari craton acted as a southwest-directed indenter into juvenile Mesoproterozoic island-arc terranes at ca. 1100 Ma (Jacobs et al., 1993).

The associated escape tectonics regime in the orogen is shown by extensive craton-parallel ductile shear zones and the development of pull-apart basins in the preserved African foreland (Koras– Sinclair basins). Other example of indenter-escape tectonic has been reported in Australia, where the Yilgara craton indented
the Pilbara–Gawler craton during the late Paleoproterozoic Capricorn orogeny (Krapez, 1999). Another example is the East African–Antarctic orogen, which resulted from the collision of various parts of proto-East and West Gondwana during late Neoproterozoic–early Paleozoic time (between 650 and 500Ma).
The southern part of this Himalayan-type orogen can be interpreted in terms of a lateral-escape tectonic model. Small microplates (the Falkland, Ellsworth–Haag, and Filchner blocks) probably represent shear-zone-bounded blocks produced by tectonic translation during lateral escape, similar to those currently evolving in Southeast Asia.
Assuming amodel of indentation-escape tectonics for northern Patagonia, the belts with steep gravity gradients described in this study represent transcurrent shear zones separating shear-zone bounded blocks (Figs. 11 and 12A). The block situated between lineaments C, D and E has possibly acted as a northwest-directed
indenter associated with west-directed movements of the area located south of the Huincul Fault zone. Several small blocks were displaced west and southwest, possibly the western sector of theChoeleChoel High and the areas located west of the lineAand southwest of line B, as recorded in the La Seña–Pangaré mylonites
and Nahuel Niyeu system (Fig. 12A).
The eastern sector of the Choele Choel High, between the Huincul Fault zone and lineament E was displaced to the east and southeast during lateral escape. In areas of transpression, such as along the Huincul Fault Zone between Estancia El Caldén and Chimpay (Fig. 10), rocks were faulted upwards to form positive
flower structures. Such a configuration can explain the uplift of the Choele Choel High against the Carapacha Basin and accounts for the change of detritus composition as detected in the Carapacha and Claromecó Basins (Rossello et al, 1997, Melchor, 1999).
It seems probable that some blocks in the North Patagonian Massif, located south of the Huincul Fault Zone, rotated counterclockwise during Permo-Triassic time (Fig. 12A). It is doubtful that the deformation can be explained in terms of the collision of two major crustal blocks, principally because there is no conclusive
evidence for such a collision.
The development of the Gondwanide deformation in northern Patagonia–Sierras Australes–La Pampa province in a location far from the active convergent margin of Gondwana is attributable to two factors: a) dextral margin-parallel translation of a crustal block or blocks (Fig. 12A and B) outboard of the Gondwanide
Orogen; and b) the oblique orientation of this area relative to the east–west trend (present coordinates) of the
orogen in the Cape Fold Belt, as shown in Fig. 12B.
Coincidences in timing of deformation and the orientation of principal stress directions, point to an evolutionary history that is similar to other portions of the Gondwanide Orogen (de Wit and Ransome, 1992; Cobbold et al., 1991, Vaughan and Pankhurst, 2008).
In the southern Canadian Rocky Mountain fold and thrust belt, Late Cretaceous and Tertiary deformation developed ~1500 km inboard of the active convergent margin (McMechan and Thompson, 1989). Oblique (dextral) subduction of oceanic plates beneath western North America resulted in the development of major dextral strike-slip faults within the overriding North American crust. The development of this fold and thrust belt 1500 km inboard of the cordilleran margin in response to oblique subduction-driven, margin-parallel dextral translation
was considered as an analogue for development of the Cape Fold Belt (Johnston, 2000). According to the evidence presented here, this model can also be applied to northern Patagonia. However, the relatively extensive width of this active margin is at present not fully understood.


9. Conclusions
North–South belts of Pampean and Famatinian rocks that cross the supposed boundary between the northern Patagonia and central Argentina regions support the presence of a common continental crust in both areas. No geological or geophysical evidences for oceanic crust formed by mafic– ultramafic rocks between both areas have been identified. The gravity and magnetic anomalies over the supposed boundary are incompatible with a suture composed of such rocks.
The continental crust in the studied area was sliced by displacements along the dextral Huincul Fault Zone and
several minor lineaments that represent strike-slip faults along the boundaries of several continental blocks. South of the Huincul Fault Zone, the indentation of these blocks, related to the translation and simultaneous counterclockwise rotation along a NW-direction produced escape of minor slices in W, SW, E, SE, and N directions. Such movements are recorded in Gondwana rocks and mylonitic belts around northern
Patagonia.
Transpression in the northern part of the massif possibly produced positive flower structures, e.g.: the Choele Choel High or those recognized in the Neuquén Basin.

The balance of evidence suggests that the Gondwanide Orogeny in the northern Patagonia area is due to the transtensive– transpressive deformation. The dextral strike-slip component detected in the northern Patagonia–Sierras Australes, Cape Fold Belt (Africa), and in the Ellsworth Whitmore Mountains block (Antarctica), may reflect margin-parallel strike-slip movements induced by oblique subduction (Cobbold et al., 1991, 1992) beneath the convergent southern margin of Gondwana. volver arriba

Reinterpretacion de Ramos (2008) sobre aloctonía de Patagonia y configuración tectónica de Patagonia

Mas información en: Victor A. Ramos, 2008. Patagonia: A paleozoic continent adrift? Journal of South American Earth Sciences. 26: 235–251

volver arriba comparar con la hipotesis de Gregori-Kostadinoff

BIBLIOGRAFIA RELACIONADA A LA CONFIGURACIONDE UNIDADES Y A LA ALOCTONIA-AUTOCTONIA DE PATAGONIA

  • Almeida, F.F.M.de, Hasui, Y., Brito Neves, B.B., 1976. The Upper Precambrian of South America. Universidade de Sao Paulo, Instituto de Geociencias, Boletim, vol. 7, pp. 45–80.
  • Andreis, R.R., Cladera, G., 1992. Las epiclastitas pérmicas de la cuenca Sauce Grande (Sierras Australes, Buenos Aires, Argentina), Parte II: Emplazamiento tectónico de las áreas de aporte. 4 Reunión Argentina de Sedimentología, La Plata, Actas 1, pp. 135–142.
  • Andreis, R.R., Archangelsky, S., González, C.R., López Gamundi, O., Sabattini, N., Aceñolaza, G., Azcuy, C.L., Cortiñas, J., Cuerda, A., Cúneo, R., 1987. Cuenca Tepuel– Genoa. In: Archangelsky, S. (Ed.), El Sistema Carbonífero de la República Argentina. Academia Nacional de Ciencias, Publicación Especial, pp. 169–196.
  • Andreis, R.R., Iñiguez, A.M., Lluch, J.L., Rodríguez, S., 1989. Cuenca paleozoica de Ventania, Sierras Australes, provincia de Buenos Aires. In: Chebli, G., Spalletti, L. (Eds.), Cuencas Sedimentarias. Universidad Nacional de Tucumán, Serie Correlación Geológica, Tucumán, vol. 6, pp. 265–298.
  • Armstrong, R.A., de Wit, M.J., Reid, D., York, D., Zartman, R., 1998. Cape Town’s Table Mountain reveals rapid Pan-African uplift of its basement rock. Journal of African Earth Sciences 27 (1), 10–11.
  • Augustsson, C., Bahlburg, H., 2008. Provenance of late Palaeozoic metasediments of the Patagonian proto-Pacific margin (southernmost Chile and Argentina). International Journal of Earth Sciences (Geologische Rundschau) 97, 71–88.
  • Augustsson, C., Münker, C., Bahlburg, H., Fanning, C.M., 2006. Provenance of Late Palaeozoic metasediments of the SW South American Gondwana margin: a combined U–Pb and Hf-isotope study of single detrital zircons. Journal of the Geological Society 163, 983–995.
  • Bangert, B., Stollhofen, H., Lorenz, V., Armstrong, R., 1999. The geochronology and significance of ashfall tuffs in the glaciogenic Carboniferous–Permian Dwyka Group of Namibia and South Africa. Journal of African Earth Sciences 29, 33–50.
  • Basei, M.A.S., Brito Neves, B.B., Varela, R., Teixeira, W., Siga Jr., G., Sato, A.M., Cingolani, C.A., 1999. Isotopic dating on the crystalline basement rocks of the Bariloche Region, Río Negro, Argentina. 2o Simposio sudamericano de geología isotópica (Carlos Paz), Servicio Geológico Minero Argentino, Anales, vol. 34, pp. 15–18.
  • Basei, M.A., Varela, R., Passarelli, C., Siga Jr., O., Cingolani, C., Sato, A., Gonzalez, P.D., 2005. The crystalline basement in the north of Patagonia: isotopic ages and regional characteristics. In: Pankhurst, R., Veiga, G. (Eds.), Gondwana 12: Geological and Biological Heritage of Gondwana, Abstracts, Academia Nacional de Ciencias, Córdoba, p. 62.
  • Bellosi, E.S., Jalfin, G.A., 1989. Cuencas neopaleozoicas de la Patagonia extraandina e Islas Malvinas. In: Chebli, G., Spalletti, L. (Eds.), Cuencas Sedimentarias. Universidad Nacional de Tucumán, Serie Correlación Geológica, Tucumán, vol. 6, pp. 379–394.
  • Buggisch, W., 1987. Stratigraphy and very low grade metamorphism of the Sierras Australes of the Province of Buenos Aires, Argentina and implications in Gondwana correlations. Zentralblatt Mineralogie Geologie Paläontologie 1, 819–837.
  • Burckhardt, C., 1902. Traces géologiques d’un ancien continent pacifique. Revista del Museo de La Plata 10, 179–192.
  • Cagnoni, M.C., Linares, E., Ostera, H.A., Parica, C.A., Remesal, M.B., 1993. Caracterización geoquímica de los metasedimentos de la Formación Nahuel Niyeu: Implicancias sobre su proveniencia y marco tectónico. 12 Congreso Geológico Argentino y 2 Congreso de Exploración de Hidrocarburos, Actas 1, pp. 281–288.
  • Caminos, R., Llambías, E., 1984. El basamento cristalino. In: Ramos,V.A. (Ed.), Geología y recursos naturales de la provincia de Río Negro. 9 Congreso Geológico Argentino, Relatorio, Buenos Aires, pp. 37–63.
  • Caminos, R., Llambías, E.J., Rapela, C.W., Parica, C.A., 1988. Late Paleozoic Early Triassic magmatic activity of Argentina and the significance of new Rb/Sr ages from Northern Patagonia. Journal of South American Earth Sciences 1 (2), 137– 146.
  • Cawood, P.A., 2005. Terra Australis orogen: Rodinia breakup and development of the Pacific and Iapetus margins of Gondwana during the Neoproterozoic and Paleozoic. Earth-Science Reviews 69, 249–279.
  • Cerredo, M.E., López de Luchi, M.G., 1998. Mamil Choique granitoids, southwestern North Patagonian Massif, Argentina: magmatism and metamorphism associated with a polyphasic evolution. Journal of South American Earth Sciences 11 (5), 499–515.
  • Charrier, R., Pinto, L., Pía Rodríguez, M., 2007. Tectonostratigraphic evolution of the Andean Orogen in Chile. In: Moreno, T., Gibbons, W. (Eds.), The Geology of Chile. The Geological Society, London, pp. 1–114.
  • Chernicoff, J., Zapettini, E., 2004. Geophysical evidence for terrane boundaries in South-Central Argentina. Gondwana Research 7 (4), 1105–1117.
  • Chotin, P., Giret, A., 1979. Analysis of northern patagonian transverse structure (Chile, Argentina – 38 to 42 s.l.) from landsat documents. 7 Congreso Geológico Argentino (Neuquén), Buenos Aires, Actas 2, pp. 197–202.
  • Cingolani, C., Varela, R., 1973. Examen geocronológico por el método rubidio estroncio de las rocas ígneas de las Sierras Australes bonaerenses. 5 Congreso Geológico Argentino, Buenos Aires, Actas 1, pp. 349–371.
  • Cingolani, C., Varela, R., 1976. Investigaciones geológicas y geocronológicas en el extremo sur de la isla Gran Malvina, sector cabo Belgrano (Cabo Meredith), Islas Malvinas. 6 Congreso Geológico Argentino, Buenos Aires, Actas 1, pp. 457–474.
  • Compton, J.S., 2004. The Rocks and Mountains of Cape Town. Double Storey Books, Cape Town, 112 pp.
  • Coney, P.J., Jones, D.L., Monger, J.W.H., 1980. Cordilleran suspect terranes. Nature 288, 329–333.
  • Cortiñas, J.S., 1996. La cuenca de Somuncurá – Cañadón Asfalto: sus límites, ciclos evolutivos del relleno sedimentario y posibilidades exploratorias. 13 Congreso Geológico Argentino y 3 Congreso de Exploración de Hidrocarburos, Buenos Aires, Actas I, pp. 147–163.
  • Cortiñas, J.C., Arbe, H.A., 1982. Facies y paleoambientes sedimentarios del grupo Río Genoa, Pérmico Inferior de la región de Nueva Lubecka, provincia de Chubut. Revista de la Asociación Geológica Argentina 37 (3), 300–312.
  • Dalla Salda, L., Cingolani, C., Varela, R., 1990. The origin of Patagonia. Comunicaciones, Una revista de geología andina 41, 55–61.
  • Dalla Salda, L.H., Varela, R., Cingolani, C.A., 1992a. Los granitoides de Chasicó Mencué, Macizo Norpatagónico, Río Negro, Su importancia geotectónica. Revista de la Asociación Geológica Argentina 46 (1–4), 189–200.
  • Dalla Salda, L., Cingolani, C.A., Varela, R., 1992b. El basamento cristalino de la región nordpatagónica de los lagos Gutiérrez, Mascardi y Guillelmo, provincia de Río Negro. Revista de la Asociación Geológica Argentina 46 (3–4), 263–276.
  • Dalla Salda, L., Varela, R., Cingolani, C.A., Aragón, E., 1994. The Río Chico Paleozoic crystalline complex and the evolution of Northern Patagonia. Journal South American Earth Sciences 7 (3–4), 377–386.
  • De Wit, M., 1977. The evolution of the Scotia Arc as a Key to the reconstruction of southwestern Gondwanaland. Tectonophysics 37 (1–3), 53–81.
    De Wit, M., Armstrong, R., Bowring, S., Alexander, J., Branch, T., Decker, J., Ghosh, J., Moore, S., Lindeque, A., Stankiewicz, J., Rakatolosofo, N., 2007. Chrono-, chemical-, seismic-, electrical- and tectono-stratigraphy across parts of the Cape Fold Belt – Karoo Basin of South Africa: new foundations for correlations across the South Atlantic. In: 1 Workshop: Problems in Western Gondwana Geology (Gramado), Extended Abstracts, Porto Alegre, pp. 34–41.
  • Dimieri, L., Delpino, S., Turienzo, M., 2005. Estructura de las Sierras Australes de Buenos Aires. In: de Barrio, R., Etcheverry, O., Caballé, M.F., Llambías E. (Eds.), Geología y recursos minerales de la Provincia de Buenos Aires. 16 Congreso Geológico Argentino, Relatorio, pp. 101–118.
  • du Toit, A., 1927. A geological comparison of South America with South Africa. With a paleontological contribution by F. Cowper Reed. Carnegie Institute of Washington, Washington, Publication 381, 158 p.
  • Fígari, E., 2005. Estructura y evolución geológica de la Cuenca de Cañadón Asfalto, provincia del Chubut. Universidad de Buenos Aires, Buenos Aires, Ph.D. thesis, unpublished, 177 pp.
  • Fitzgerald, M.G., Mitchum, R.M., Uliana, M.A., Biddle, K.T., 1990. Evolution of the San Jorge Basin, Argentina. American Association of Petroleum Geologists, Bulletin 74 (6), 879–920.
  • Forsythe, R., 1982. The Late Paleozoic to Early Mesozoic evolution of Southern South America: a plate tectonic interpretation. Journal Geological Society 139, 671– 682.
  • Fortey, R., Pankhurst, R.J., Hervé, F., 1992. Devonian trilobites at Buill, Chile (42S). Revista Geológica de Chile 19 (2), 133–144.
  • Franzese, J.R., Spalletti, L.A., 2001. Late Triassic continental extension in southwestern Gondwana: tectonic segmentation and pre-break-up rifting. Journal of South American Earth Sciences 14, 257–270.
  • Frutos, J., Tobar, A., 1975. Evolution of the southwestern continental margin of South America. In: Campbell, K.S.W. (Ed.), III International Gondwana Symposium (Canberra), Gondwana Geology, Australian National University Press, pp. 565–578.
  • Galeazzi, J.S., 1996. Cuenca de Malvinas. In: Ramos V.A., Turic, M.A. (Eds.), Geología y Recursos Naturales de la Plataforma Continental Argentina. 13 Congreso Geológico Argentino y 3 Congreso de Exploración de Hidrocarburos, Buenos Aires, Relatorio 15, pp. 273–309.
  • Gallagher, J.J., 1990. Andean chronotectonics. In: Ericksen, G.E., Cañas Pinochet, M.T., Reinemud, J.A. (Eds.), Geology of the Andes and its Relation to Hydrocarbon and Mineral Resources, Circumpacific Council for Energy and
    Mineral Resources, Houston, Earth Sciences Series, vol. 11, pp. 23–38.
  • Ghidella, M.E., Paterlini, C.M., Kovacs, L.C., Rodríguez, G.A., 1995. Magnetic anomalies on the Argentina Continental Shelf. In: Fourth International Congress of the Brazilian Geophysical Society and First Latin American
    Geophysical Conference, Río de Janeiro, Expanded Abstracts, vol. 1, pp. 269– 272.
  • Ghidella, M.E., Yañez, G., LaBrecque, H.L., 2002. Revised tectonic implications for the magnetic anomalies of the Western Weddell Sea. Tectonophysics 347 (1-3), 65– 86.
  • Ghidella, M.E., Lawver, L.A., Marenssi, S., Gahagan, L.M., 2007. Plate kinematic models for Antarctica during Gondwana break-up: a review. Revista de la Asociación Geológica Argentina 62 (4), 636–646.
  • Giacosa, R.E., Márquez, M.M., 2002. El Basamento Paleozoico de la Cordillera Patagónica. In: Haller, M.J. (Ed.), Geología y Recursos Naturales de Santa Cruz, Relatorio, vol. 1(3), pp. 45–56.
  • González, P., Poiré, D., Varela, R., 2002. Hallazgo de trazas fósiles en la Formación El Jagüelito y su relación con la edad de las metasedimentitas, Macizo Norpatagónico Oriental, provincia de Río Negro. Revista de la Asociación
    Geológica Argentina 57 (1), 35–44.
  • Von Gosen, W., Buggisch, W., 1989. Tectonic evolution of the Sierras Australes fold and thrust belt (Buenos Aires Province, Argentina): an outline. Zentralblatt für Geologie und Paläontologie Teil I(5/6), Stuttgart, pp. 947–958.
  • Halpern, M., 1968. Ages of antarctic and argentine rocks bearing on continental drift. Earth and Planetary Science Letters 5, 159–167.
  • Harrington, H.J., 1955. The Permian Eurydesma fauna of eastern Argentina. Journal of Paleontology 29 (1), 112–128.
  • Harrington, H.J., 1962. Paleogeographic development of South America. American Association of Petroleum Geologists, Bulletin 46 (10), 1773–1814.
  • Heredia, N., Alonso, J.L., Busquets, P., Colombo, F., Farías, P., Gallastegui, G., Gallastegui, J., García Sansegundo, J., Giacosa, R.E., Montes, M., Nozal F., Ramos, V., Rodríguez Fernández, L.R., 2006. El orógeno gondwánico entre los Andes Centrales (30S) y la Península Antártica (65S): Evolución y marco geotectónico. 11 Congreso Geológico Chileno (Antofagasta), Actas 1, pp. 251–254.
  • Hervé, F., Fanning, C.M., Pankhurst, R.J., 2003. Detrital zircon age patterns and provenance of the metamorphic complexes of southern Chile. Journal of South American Earth Sciences 16, 107–123.
  • Hervé, F., Fanning, C.M., Yaxley, G.M., Pankhurst, R.J., 2006. U–Pb and Lu–Hf isotopric constraints on the provenance of Permian detritus in metasedimentary rocks of Southern Chile and Livingston Island, Antarctica.
    11 Congreso geológico Chilenao (Antofagasta), Actas 1, pp. 259–260.
  • Hervé, F., Fanning, C.M., Mpodozis, C., Pankhurst, R.J., 2008. Aspects of the Phanerozoic evolution of southern Patagonia as suggested by detrital zircon age patterns. 6 South American Symposium on Isotope Geology, Abstracts, S.C. Bariloche, p. 3.
    Homovc, J.F., Constantini, L.A., 2001. Hydrocarbon exploration potential within intraplate shear-related depocenters, Deseado and San Julián basins, southern Argentina. American Association of Petroleum Geologists, Bulletin 85 (10), 1795–1816.
  • Introcaso, A., 2003. Significativa descompensación isostática en la Cuenca del Colorado (República Argentina). Revista de la Asociación Geológica Argentina 58 (3), 474–478.
  • Keidel, J., 1925. Sobre el desarrollo paleogeográfico de las grandes unidades geológicas de la Argentina. Sociedad Argentina de Estudios Geográficos GAEA, Anales 4, 251–312.
  • Kostadinoff, J., Gregori, D.A., Raniolo, A., 2005. Configuración geofísica-geológica del sector norte de la provincia de Río Negro. Asociación Geológica Argentina, revista 60 (2), 368–376.
  • Leanza, A.F., 1958. Geología Regional. In La Argentina, Suma de Geografía, Editorial Peuser 1 (3), 217–349. Buenos Aires.
  • Lesta, P.J., 1968. Estratigrafía de la Cuenca del Golfo San Jorge. 3 Jornadas Geológicas Argentinas, Buenos Aires, Actas 1, pp. 251–289.
  • Lesta, P., Ferello, R., Chebli, G., 1980. Chubut Extraandino. In: Segundo Simposio Geología Regional Argentina, Academia Nacional de Ciencias, Córdoba 2, pp. 1306–1387.
  • Linares, E., González, R.R., 1990. Catalogo de edades radimétricas de la República Argentina 1957-1987. Asociación Geológica Argentina, Publicaciones Especiales Serie B, Didáctica y Complementaria, vol. 19, 628 pp.
  • Lizuaín, A., 1981. Características y edad del plutonismo en los alrededores de Lago Puelo, Prov. del Chubut. 8 Congreso Geológico Argentino, Buenos Aires, Actas 3, pp. 607–616.
  • Llambías, E.J., Rapela, C.W., 1985. Geología de los complejos eruptivos de La Esperanza, provincia de Río Negro. Revista de la Asociación Geológica Argentina 39 (3–4), 220–243.
  • Llambías, E.J., Caminos, R. Rapela, C.W., 1984. Las plutonitas y vulcanitas del ciclo eruptivo Gondwánico. In: Ramos,V.A. (Ed.), Geología y recursos naturales de la provincia de Río Negro, 9 Congreso Geológico Argentino (Bariloche), Relatorio, Buenos Aires, p. 85–118.
  • Llambías, E.J., Varela, R., Basei, M., Sato, A.M., 2002. Deformación dúctil y metamorfismo neopaleozoico en Yaminué y su relación con la fase orogénica San Rafael. 15 Congreso Geológico Argentino, Actas 3, pp. 123–128.
  • López de Lucchi, M.G., Cerredo, M.E., 2008. Geochemistry of the Mamil Choique granitoids at Rio Chico, Río Negro, Argentina: Late Paleozoic crustal melting in the North Patagonian Massif. Journal of South American Earth Sciences 25, 526–546.
  • López de Luchi, M.G., Spikermann, J.P., Strelin, J.A., Morelli, J., 1992. Geología y petrología de los plutones de la Tapera de Burgos, Arroyo El Rápido y Cerro Caquel, Departamento Languineo, Provincia del Chubut. Revista de la
    Asociación Geológica Argentina 47 (1), 87–98.
  • López Gamundi, O., 2006. Permian plate margin volcanism and tuffs in adjacent basins of west Gondwana: age constraints and common characteristics. Journal of South American Earth Sciences 22, 227–238.
  • López Gamundi, O.R., Breitkreuz, C., 1997. Carboniferous to Triassic evolution of the Panthalassan margin in southern South America. In: López Dickins, J.M. (Ed.), Late Paleozoic and Early Mesozoic Events and Their Global Correlation. Cambridge University Press, pp. 8–19.
  • López Gamundi, O., Limarino, C.O., 1985. Facies de abanico submarino en el grupo Tepuel (Paleozoico superior), provincia del Chubut. Revista de la Asociación Geológica Argentina 39 (3–4), 251–261.
  • López Gamundi, O.R., Rossello, E., 1992. La cuenca intraserrana de Claromecó, Argentina: un ejemplo de cuenca de antepaís hercínica. 8 Congreso Latinoamericano de Geología, Salamanca, vol. 4, pp. 55–59.
  • López Gamundi, O.R., Conaghan, P., Rossello, E.A., Cobbold, P.R., 1995. The Tunas Formation (Permian) in the Sierras Australes Foldbelt, East-Central Argentina: evidence of syntectonic sedimentation in a Varisican foreland basin. Journal of South American Earth Sciences 8 (2), 129–142.
  • Loske, W., Márquez, M., Giacosa, R., Pezzuchi, H., Fernández, M.I., 1999. U/Pb geochronology of pre-Permian basement rocks in the Macizo del Deseado, Santa Cruz province, Argentine Patagonia. 15 Congreso Geológico Argentino (Salta), Actas 1, p. 102.
  • Manceñido, M., Damborenea, S., 1984. Megafaunas de invertebrados paleozoicos y mesozoicos. In: Ramos,V.A. (Ed.), Geología y recursos naturales de la provincia de Río Negro. 9 Congreso Geológico Argentino, Relatorio, Buenos Aires, vol. 413–463.
  • Martin, A.K., Hartnady, C.H.J.H., Goodland, S.W., 1981. A revised fit of South America and South Central Africa. Earth and Planetary and Science Letters 54, 293–305.
  • Martínez, C., 1980. Geologie des Andes Boliviennes. Estructure et evolution de la chaine hercynienne et de la chaine andine dans le nord de la Cordillere des Andes de Bolivie. Travaux et Documents de ORSTOM (Office Recherche Scientifique Technique Outre-Mer) 119, 1–352. París.
  • Massabie, A., Rossello, E., 1984. La discordancia pre-Sauce Grande y su entorno estratigráfico. 9 Congreso Geológico Argentino, Actas 1, pp. 337–352.
  • Max, M.D., Ghidella, M., Kovacs, L., Paterlini, M., Valladares, J.A., 1999. Geology of the Argentine continental shelf and margin from aeromagnetic survey. Marine and Petroleum Geology 16 (1), 41–64.
  • Milani, E.J., 2007. The Paraná Basin: a multi-cycle sedimentary and magmatic intracratonic province of W Gondwana. I Workshop: Problems in Western Gondwana Geology (Gramado), Extended Abstracts, Porto Alegre, pp. 99–107.
  • Milani, E.J., De Wit, M.J., 2008. Correlations between the classic Parana’ and Cape- Karoo sequences of South America and southern Africa and their basin infills flanking the Gondwanides: du Toit revisited. In: Pankhurst, R.J., Trouw, R.A.J., Brito Neves, B.B., De Wit, M.J. (Eds.), West Gondwana: Pre-Cenozoic Correlations Across the South Atlantic Region. Geological Society, Special Publications, London, vol. 294, pp. 319–342.
  • Monger, J.H.W., Price, R.A., Tempelman-Kluit, D.J., 1982. Tectonic accretion and the origin of two major metamorphic and plutonic welts in the Canadian Cordillera. Geology 10, 70–75.
  • Moreira, P., Fernández, R., Hervé, F., Fanning, C.M., 2007. U–Pb SHRIMP ages from detrital zircons of the La Modesta Formation, Deseado Massif, Argentina. Universidad de Chile, Geosur 207, An International Congress on the Geology and Geophysics of the Southern Hemisphere, Abstract, Santiago, vol. 104.
  • Moreno, F.P., 1882. Patagonia, resto de un antiguo continente hoy sumergido. Anales de la Sociedad Científica Argentina 14, 97–131. Buenos Aires.
  • Mosquera, A., 2008. Mecánica de deformación en la cuenca neuquina, Neuquén. Universidad de Buenos Aires, Ph.D. Thesis, Unpublished, Buenos Aires, 205 pp.
  • Mosquera, A., Ramos, V.A., 2006. Intraplate deformation in the Neuquén Basin. In: Kay, S.M., Ramos, V.A. (Eds.), Evolution of an Andean Margin: A Tectonic and Magmatic View from the Andes to the Neuquén Basin (35–39S latitude). Geological Society of America, Special Paper 407, pp. 97–124.
  • Page, S., 1984. Los gabros bandeados de la Sierra de Tepuel. Cuerpos del sector Suroeste, provincia de Chubut. 9 Congreso Geológico Argentino (Bariloche), Buenos Aires, Actas 2, pp. 584–599.
  • Page, R.F.N., Limarino, C.O., López Gamundi, O., Page, S., 1984. Estratigrafía del Grupo Tepuel en su perfil tipo y en la región El Molle, provincia de Chubut. 9 Congreso Geológico Argentino (Bariloche), Buenos Aires, Actas 1, pp. 619–632.
  • Pankhurst, R., Caminos, R., Rapela, C.W., 1993. Problemas geocronológicos de los granitoides Gondwánicos de Nahuel Niyeu, Macizo Norpatagónico. 12 Congreso Geológico Argentino y 2 Congreso de Exploración de Hidrocarburos (Mendoza), Buenos Aires, Actas 4, pp. 99–104.
  • Pankhurst, R.J., Rapela, C.W., Fanning, C.M., 2001. The Mina Gonzalito Gneiss: Early Ordovician Metamorphism in Northern Patagonia. 3 South American Symposium on Isotope Geology, Actas Electrónicas Sesion, , Pucón, vol. 6(4), pp. 1–4.
  • Pankhurst, R., Rapela, C., Loske, W.P., Márquez, C.M., Fanning, M.Y., 2003. Chronological study of pre-Permian basement rocks of southern Patagonia. Journal of South American Earth Sciences 16 (1), 27–44.
  • Pankhurst, R.J., Rapela, C.W., Fanning, C.M., Márquez, M., 2006. Gondwanide continental collision and the origin of Patagonia. Earth Science Reviews 76, 235–257.
  • Pezzuchi, H.D., 1978. Estudio geológico de la zona de la Estancia Dos Hermanos, Estancia 25 de Mayo y adyacencias, departamento Deseado, provincia de Santa Cruz. Universidad Nacional de La Plata, Ph.D. Thesis, unpublished, La Plata.
  • Ploszkiewicz, J.V., Orchuela, I.A., Vaillard, J.C., Viñes, R.F., 1984. Compresión y desplazamiento lateral en la zona de Falla Huincul: estructuras asociadas, provincia del Neuquén. 9E Congreso Geológico Argentino, Actas 2, 163–169. Buenos Aires.
  • Poma, S., 1986. Petrología de las rocas básicas pre-cretácicas de la sierra de Tepuel, provincia de Chubut. Ph.D. Dissertation, Universidad de Buenos Aires, Buenos Aires, unpublished, 258 pp.
  • Ramos, V.A., 1975. Geología del sector oriental del macizo Nordpatagónico entre Aguada Capitán y la Mina Gonzalito, provincia de Río Negro. Revista de la Asociación Geológica Argentina 30 (3), 274–285. Buenos Aires.
  • Ramos, V.A., 1983. Evolución tectónica y metalogénesis de la Cordillera Patagónica. 2 Congreso Nacional Geología Económica (San Juan), Actas 1, pp. 108–124.
  • Ramos, V.A., 1984. Patagonia: ¢ Un continente paleozoico a la deriva? 9 Congreso Geológico Argentino (Bariloche), Actas 2, pp. 311–325.
  • Ramos, V.A., 1986. Tectonostratigraphy, as applied to analysis of South African Phanerozoic Basins by H. de la R. Winter, discussion. Transactions Geological Society South Africa, vol. 87(2), pp. 169–179.
  • Ramos, V.A., 1988. Tectonics of the Late Proterozoic–Early Paleozoic: a collisional history of Southern South America. Episodes 11 (3), 168–174.
  • Ramos, V.A., 1996. Evolución tectónica de la Plataforma Continental. In: Ramos, V.A., Turic, M.A. (Eds.), Geología y Recursos Naturales de la Plataforma Continental Argentina. Asociación Geológica Argentina e Instituto Argentino del Petróleo, Buenos Aires, pp. 359–368.
  • Ramos, V.A., 1999. Plate tectonic setting of the Andean Cordillera. Episodes 22 (3), 183–190.
  • Ramos, V.A., 2004a. Tectonics of the southernmost Andes: a comparison between the Patagonian and the Fuegian Cordilleras. Bolletino di Geofisica Teorica ed Applicata 45, 1–10 (2, supplement, Trieste).
  • Ramos, V.A., 2004b. La Plataforma Patagónica y sus relaciones con la Plataforma Brasilera. In: Mantesso-Neto, V., Bartorelli, A., Ré Carneiro, C.D., Brito Neves, B.B. (Eds.) Geologia do Continente Sul-Americano, Sao Paulo, vol. 22, pp. 371–381.
  • Ramos, V.A., 2008. The basement of the Central Andes: the Arequipa and related terranes. Annual Review on Earth and Planetary Sciences 36, 289–324.
  • Ramos, V.A., Kostadinoff, J., 2005. La cuenca de Claromecó. In: de Barrio, R.E., Etcheverry, R.O., Caballé, M.F., Llambías, E. (Eds.), Geología y recursos minerales de la Provincia de Buenos Aires. 16 Congreso Geológico Argentino, Relatorio, pp. 473–480.
  • Ramos, V.A., Palma, M.A., 1996. Tectónica del Pérmico de Argentina. In Archangelsky, S. (Ed.), El sistema Pérmico en la República Argentina y en la República Oriental del Uruguay. Academia Nacional de Ciencias, Córdoba, pp.
    239–254.
  • Ramos, V.A., Riccardi, A.C., Rolleri, E.O., 2004. Límites Naturales de la Patagonia. Revista de la Asociación Geológica Argentina 59 (4), 785–786.
  • Rapalini, A.E., 2005. The accretionary history of southern South America from the latest Proterozoic to the Late Paleozoic: some paleomagnetic constraints. In: Vaughan, A.P.M., Leat, P.T., Pankhurst, R.J. (Eds.), Terrane Processes at the Margins of Gondwana. Geological Society, London, Special Publications, vol. 246, pp. 305–328.
  • Rapalini, A.E., López de Luchi, M., Croce, F., Lince Klinger, F., Tomezzoli, R., Gímenez, M., 2008. Estudio geofísico del complejo plutónico Navarrte: implicancias para la evolución tectónica de Patagonia en el Paleozoico Tardío. 5 Simposio Argentino del Paleozoico Superior, Resúmenes, Buenos Aires, vol. 34, p. 34.
  • Rapela, C., Caminos, R., 1987. Geochemical characteristics of the Upper Paleozoic magmatism in the eastern sector of Northpatagonian massif. Revista Brasileira de Geociencias 17 (4), 535–543. Sao Paulo.
  • Rapela, C.W., Kostadinoff, J., 2005. El basamento de Sierra de la Ventana: historia tectonomagmática. In: de Barrio, R.E., Etcheverry, R.O., Caballé, M.F., Llambías, E. (Eds.), Geología y recursos minerales de la Provincia de Buenos Aires. 16 Congreso Geológico Argentino, Relatorio, pp. 69–84.
  • Rapela, C.W., Llambías, E.J., 1985. Evolución magmática y relaciones regionales de los Complejos Eruptivos de La Esperanza. Provincia de Río Negro. Revista de la Asociación Geológica Argentina 40 (1–2), 4–25.
  • Rapela, C.W., Pankhurst, R.J., Harrison, S.M., 1989. Gondwana Plutonism of Northern Patagonia. 28th International Geological Congreso, Washington DC, Abstracts, vol. 2, p. 675.
  • Rapela, C.W., Pankhurst, R.J., Fanning, C.M., Grecco, L.E., 2003. Basement evolution of the Sierra de la Ventana Fold Belt: new evidence for Cambrian continental rifting along the southern margin of Gondwana. Journal of the Geological Society (London) 160, 613–628.
  • Sandwell, D.T., Smith, W.H.F., 1997. Marine gravity anomaly fromGeosat and 262 ERS- 1 satellite altimetry. Journal of Geophysical Research 102 (B5). 263, 10039–10054.
  • Santos, R.V., Souza, P.A., Souza de Alvarenga, C.J., Dantas, E.L., Pimentel, M.M., Gouveia de Oliveira, G., Medeiros de Araújo, L., 2006. Shrimp U–Pb zircon dating and palynology of bentonitic layers from the Permian Iratí Formation, Paraná Basin, Brazil. Gondwana Research 9 (4), 456–463.
  • Silvestro, J., Zubiri, M., 2008. Convergencia oblicua: Modelo estructural alternativo para la dorsal neuquina (39S) Neuquén. Revista de la Asociación Geológica Argentina 63 (1), 49–64.
  • Söllner, F., Miller, H., Hervé, F., 2000. An early Cambrian granodiorite age from the pre-Andean basement of Tierra del Fuego (Chile): the missing link between South America and Antactica? Journal of South American Earth Sciences 13, 163–177.
  • Stern, C.H.R., Frey, F.A., Zartman, R.E., Peng, Z., Kyser, T.K., 1990. Trace element and Sr, Nd, Pb and O isotopic composition of Pliocene and Quaternary alkalic basalts of the Patagonian Plateau lavas of southernmost South America. Contributions to Mineralogy and Petrology 104, 94–308.
  • Sylwan, C.A., 2001. Geology of the Golfo San Jorge Basin, Argentina. Journal of Iberian Geology 27, 123–157.
  • Tankard, A.J., Jackson, M.P.A., Eriksson, K.A., Hobday, D.K., Hunter, D.R., Minter, W.E.L., 1982. Crustal Evolution of Southern Africa: 3.8 Billion years of Earth History. Springer-Verlag, New York. pp. 480.
  • Tickyj, H., Dimieri, L.V., Llambías, E.J., Sato, A.M., 1997. Cerro Los Viejos (38 28’S – 64 26’O): Cizallamiento dúctil en el sudeste de La Pampa. Revista de la Asociación Geológica Argentina 52 (3), 311–321.
  • Tohver, E., Cawood, P.A., Rossello, E.A., Riccomini, C., 2007. Demise of the Gondwana ice sheet in SW Gondwana: time constraints from U–Pb zircon ages. Universidad de Chile, Geosur 207, An International Congress on the Geology and Geophysics of the Southern Hemisphere, Abstract, Santiago, vol. 163.
  • Tomezzoli, R.N., Cristallini, E.O., 1998. Nuevas evidencias sobre la importancia del fallamiento en la estructura de las Sierras Australes de la Provincia de Buenos Aires. Revista de la Asociación Geológica Argentina 53 (1), 117–129.
  • Tomezzoli, R.N., Vilas, J.F., 1999. Paleomagnetic constraints on the age of deformation of the Sierras Australes thrust and fold belt, Argentina. Geophysical Journal Interior 138, 857–870.
  • Toubes, R.O., Spikermann, P.J., 1974. Algunas edades K/Ar y Rb/Sr de plutonitas de la Cordillera Patagónica entre los paralelos 40 y 44 de latitud sur. Revista de la Asociación Geológica Argentina 2 (4), 382–396.
  • Uliana, M.A., Biddle, K.T., 1987. Permian to Late Cenozoic evolution of Northern Patagonia: main tectonic events, magmatic activity and depositional trends. In: McKenzie, G.D. (Ed.), Gondwana Six: Structure, Tectonics and Geophysics. American Geophysical Union, Geophysical Monograph, Washington, vol. 40, pp. 271–286.
  • Valencio, D.A., Vilas, J.F.A., 1985. Evidence of a microplate in the Southern Andes. Journal of Geodynamics 2 (2-3), 183–192.
  • Varela, R., Dalla Salda, L., Cingolani, C., 1986. Estructuras de la Sierra Colorada, Chasicó y Cortapié, Sierras Australes, provincia de Buenos Aires. Revista de la Asociación Geológica Argentina 40 (3–4), 254–261. 1985.
  • Varela, R., Cingolani, C., Sato, A., Dalla Salda, L., Brito Neves, B., Basei, M., Siga Jr., O., Texeira, W., 1997. Proterozoic and Paleozoic evolution of Atlantic area of North- Patagonian Massif, Argentina. 1 South American Symposium on Isotope Geology, Actas, San Pablo, pp. 326–329.
  • Varela, R., Basei, M., Sato, A., Siga Jr., O., Cingolani, C., Sato, K., 1998a. Edades isotópicas Rb/Sr y U/Pb en rocas de Mina Gonzalito y Arroyo Salado. Macizo Norpatagónico Atlántico, Río Negro, Argentina. 10 Congreso Latinoamericano de Geología y 6 Congreso Nacional de Geología Económica, Buenos Aires, Actas 1, pp. 71–76.
  • Varela, R., Basei, M., Sato, A.M., Llambías, E.J., Siga Jr., O., 1998b. Edades isotópicas del Complejo Yaminué, Macizo Norpatagónico. 2 Simposio Argentino Paleozoico Superior, Resúmenes, Trelew, vol. 31, p. 33.
  • Varela, R., Basei, M.A.S., Brito Neves, B.B., Sato, A.M., Teixeira, W., Cingolani, C.A., Siga Jr., O., 1999. Isotopic study of igneous and metamorphic rocks of Comallo- Paso Flores, Río Negro, Argentina. 2o Simposio Sudamericano de Geología Isotópica (Carlos Paz). Servicio Geológico Minero Argentino, Anales, Buenos Aires, vol. 34, pp. 148–151.
  • Varela, R., Basei, M., Cingolani, C.A., Siga Jr., O., Passarelli, C.R., 2005. El basamento cristalino de los Andes norpatagónicos en Argentina: geocronología e interpretación tectónica. Revista geológica de Chile 32 (2), 167–187.
  • Varela, R., Sato, K., González, P.D., Sato, A., Basei, M., 2007. Descifrando la edad y significado del plutonismo Paleozoico en Sierra Grande, Noreste Patagónico, Argentina. 5o Congreso Uruguayo de Geología, Resúmenes, vol. 132.
  • Von Gosen, W., 2002. Polyphase structural evolution in the northeastern segment of the North Patagonian Massif (southern Argentina). Journal of South America Earth Sciences 15 (5), 591–623.
  • Von Gosen, W., 2003. Thrust tectonics in the North Patagonian Massif (Argentina): implication for a Patagonian plate. Tectonics 22 (1), 1005. doi:10.1029/ 2001ITC901039.
  • Weber, E.I., 1983. Descripción geológica de la Hoja 40j, Cerro El Fuerte, Prov. de Río Negro. Boletín del Servicio Geológico Nacional 196, 1–69. Buenos Aires.
  • Windhausen, A., 1931. Geología Argentina. Geología Histórica y Regional del Territorio Argentino. J. Peuser 2, 1–645. Buenos Aires.
  • volver arriba