Next Article in Journal
Geochemical Modelling of the Evolution of Caprock Sealing Capacity at the Shenhua CCS Demonstration Project
Next Article in Special Issue
Compositional Variability of Monazite–Cheralite–Huttonite Solid Solutions, Xenotime, and Uraninite in Geochemically Distinct Granites with Special Emphasis to the Strongly Fractionated Peraluminous Li–F–P-Rich Podlesí Granite System (Erzgebirge/Krušné Hory Mts., Central Europe)
Previous Article in Journal
Chemical and Phase Reactions on the Contact between Refractory Materials and Slags, a Case from the 19th Century Zn-Pb Smelter in Ruda Śląska, Poland
Previous Article in Special Issue
Geochronology and Geochemistry of Archean TTG and Tremolite Schist Xenoliths in Yemadong Complex: Evidence for ≥3.0 Ga Archean Continental Crust in Kongling High-Grade Metamorphic Terrane, Yangtze Craton, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Tres Arroyos Granitic Aplite-Pegmatite Field (Central Iberian Zone, Spain): Petrogenetic Constraints from Evolution of Nb-Ta-Sn Oxides, Whole-Rock Geochemistry and U-Pb Geochronology

by
Idoia Garate-Olave
*,
Encarnación Roda-Robles
,
Pedro Pablo Gil-Crespo
,
Alfonso Pesquera
and
Jon Errandonea-Martin
Department of Geology, Faculty of Science and Technology, University of the Basque Country UPV/EHU, 48940 Leioa, Spain
*
Author to whom correspondence should be addressed.
Minerals 2020, 10(11), 1008; https://doi.org/10.3390/min10111008
Submission received: 27 October 2020 / Revised: 6 November 2020 / Accepted: 10 November 2020 / Published: 12 November 2020
(This article belongs to the Special Issue Geochemistry of Granites and Granitic Pegmatites)

Abstract

:
Abundant Li-Cs-Ta aplite-pegmatite dykes were emplaced in the western Central Iberian Zone of the Iberian Massif during the Variscan Orogeny. Their origin and petrogenetic relationships with the widespread granitoids have led to a currently rekindled discussion about anatectic vs. granitic origin for the pegmatitic melts. To deal with these issues, the aplite-pegmatite dykes from the Tres Arroyos area, which constitute a zoned pegmatitic field related to the Nisa-Alburquerque granitic batholith, have been studied. This work comprises a complete study of Nb-Ta-Sn oxides’ mineralogy, whole-rock geochemistry, and U-Pb geochronology of the aplite-pegmatites that have been grouped as barren, intermediate, and Li-rich. The most abundant Nb-Ta-Sn oxides from Tres Arroyos correspond to columbite-(Fe), columbite-(Mn) and cassiterite. Niobium-Ta oxides show a marked increase in the Mn/(Mn+Fe) ratio from the barren aplite-pegmatites up to the Li-rich bodies, whereas variations in the Ta/(Ta+Nb) ratio are not continuous. The probable factors controlling fractionation of Mn/Fe and Ta/Nb reflected in Nb-Ta oxides may be attributed to the crystallization of tourmaline, phosphates and micas. The lack of a progressive Ta/Nb increase with the fractionation may be also influenced by the high F and P availability in the parental pegmatitic melts. Most of the primary Nb-Ta oxides would have crystallized by punctual chemical variations in the boundary layer, whereas cassiterite formation would be related to an undercooling of the system. Whole-rock composition of the distinguished lithotypes reflects similar tendencies to those observed in mineral chemistry, supporting a single path of fractional crystallization from the parental Nisa-Alburquerque monzogranite up to the most evolved Li-rich aplite-pegmatites. The age of 305 ± 9 Ma, determined by LA-ICP-MS U-Pb dating of columbite-tantalite oxides, reinforces the linkage of the studied aplite-pegmatites and the cited parental monzogranite.

1. Introduction

Niobium-tantalum-tin (Nb-Ta-Sn) oxides are characteristic phases in evolved granitic systems, since Nb, Ta and Sn behave as incompatible elements during magmatic fractionation (e.g., [1,2,3]. In this sense, pegmatites, as one type of these granitic systems, are of great interest and represent an important target for the mining of columbite-tantalite (CT) group minerals and other strategic elements (e.g., [1,4,5,6]). The chemistry of Nb-Ta-Sn oxides, as well as quartz, feldspars, and micas, among others, is frequently used to reveal local trends within individual pegmatite bodies and regional magma fractionation trends across pegmatite fields (e.g., [7,8,9,10,11,12,13,14]). In Nb-Ta oxides, the increase in the Ta/(Ta+Nb) and/or Mn/(Mn+Fe) ratios (XTa and XMn, respectively), for example, indicates increasing fractionation degrees of the pegmatite melt (e.g., [15,16,17]). Similarly, the presence of cassiterite seems to be a marker of the degree of magmatic differentiation, since it usually appears to be restricted to the most evolved pegmatites according to the classification of Černý and Ercit [18].
This model of the genetic linkage between granite plutons and pegmatites, as residual melts derived from fractional crystallization in large-scale granitic plutons (e.g., [19]), is supported by field relationships and petrographic observations, combined with mineralogical, geochemical, and geochronological data in some pegmatite provinces (e.g., [20,21,22]). However, a number of recent geochronological studies revealed large age differences (>10 Ma) between pegmatites and spatially related granites, making their genetic relationship difficult to identify (e.g., [23,24,25]). In these cases, the pegmatites are interpreted to be of anatectic origin, meaning that they crystallized directly from melts formed by a low-degree of partial melting of high-grade metamorphic rocks (e.g., [23,24,25]). Likewise, in other regions, the observations and data are not conclusive, particularly in pegmatite provinces with overlapping magmatic events, which cannot be resolved by geochronology or where granite plutons are absent (e.g., [26,27,28]). It has been observed that although mineralogical and whole-rock geochemical inconsistencies seem to be enough to exclude the linkage between granites and pegmatites, in some cases a geochronological study is decisive (e.g., [29,30]).
Because of the general lack of suitable mineral phases for dating, such as zircon or monazite in Li-Cs-Ta (LCT) pegmatites [18], or the disturbance of the U-Pb system in those minerals, reliable geochronological data of this type of pegmatite is scarce (e.g., [1,20,31]). Nevertheless, the development of new techniques, such as LA-ICP-MS, has allowed precise U-Pb dating of other mineral phases instead of those of the CT group, which represent an interesting alternative on account of their high U but low common Pb contents (e.g., [6,32]).
In this study, we combine new mineralogical and petrographic data, together with whole-rock geochemistry and U-Pb geochronology on columbite-tantalite group minerals of aplite-pegmatites from the Tres Arroyos field (Central Iberian Zone, CIZ), in order to reveal (1) the relationships between chemical variations of Nb-Ta-Sn oxides and aplite-pegmatites evolution; (2) formation conditions; and (3) the linkage between the studied aplite-pegmatites and associated granites.

2. Geological Setting and Lithotypes of Tres Arroyos

The Tres Arroyos granite-aplite-pegmatite system is located in the southern part of the CIZ, south-west of the easternmost segment of the Nisa-Alburquerque (NA) batholith, close to the limit with the Ossa-Morena Zone (Figure 1a,b). The CIZ, characterized by abundant granitic intrusions, represents the innermost part of the European Variscan Belt (e.g., [33,34]). A Neoproterozoic-Lower Cambrian metasedimentary sequence, known regionally as the Schist-Greywacke Complex (SGC), constitutes the basement of the southern CIZ (Figure 1b; e.g., [35]). Overlying the SGC, a relatively thick Ordovician-Devonian siliciclastic succession and a thick Lower Carboniferous synorogenic sedimentary sequence are deposited, which also include Cambro-Ordovician felsic metaigneous rocks (e.g., [36,37,38]). Variscan magmatism in the southern CIZ took place during the late stages of the orogeny, and is almost completely confined to the so-called Nisa-Alburquerque-Los Pedroches Magmatic Alignment and the Central Extremadura batholith (Figure 1b; e.g., [39,40]). The NA batholith is the most extensive cordierite-bearing granitic intrusion of the cited alignment, and is composed mainly by S-type peraluminous monzogranite and minor two-mica leucogranite (Figure 1b; [41,42]). The emplacement level of the batholith was estimated at ~2–3 kbar [43], and the age of emplacement was stablished at ca. 305–309 Ma (U-Pb SHRIMP and LA-ICP-MS in zircon; [44,45]).
Two types of granites (monzogranite and marginal leucogranite) and three types of aplite-pegmatite bodies (barren, intermediate, and Li-rich) can be found in the Tres Arroyos area (e.g., [11,46,47,48,49]). These five units display a spatial zonation from the northeastern monzogranite up to the southwestern Li-rich aplite-pegmatites (Figure 1c). The monzogranite exhibits a coarse-grained seriate to porphyritic texture, with K-feldspar megacrysts, and represents the main granitic unit of the NA batholith. Quartz, K-feldspar, plagioclase, biotite, and muscovite are the major components with minor but variable contents of cordierite, tourmaline, and andalusite. Apatite group minerals, zircon, monazite, and ilmenite are common accessory phases. The leucogranitic unit represents a marginal facies of variable extension (less than 500 m across) (Figure 1c), and is characterized by a finer (up to 5 mm) grain size and noticeably higher tourmaline contents than the monzogranitic unit. Regarding the three aplite-pegmatite types, barren, intermediate and Li-rich, the most important petrographic and mineralogical features are described below (Table 1; see Garate-Olave et al. [11,47,49] for more details).
The barren aplite-pegmatites crop out near to the NA batholith, up to ~300 m away from the contact (Figure 1c). Up to 15 irregular to tabular bodies with variable thicknesses (~1–5 m) have been recognized in the field. Quartz, feldspars, Al-micas and schorl are the main constituents, whereas topaz, Fe-(Li)-micas, Fe-Mn-(±Mg±Al) phosphates, fluorapatite, and Nb-Ta-Sn oxides are accessory ones. The lack of dark micas and occurrence of topaz plus Fe-Mn phosphates distinguish them from the leucogranitic mineral association. Although these bodies do not show an internal concentric zoning, many of the barren aplite-pegmatite bodies display a layering across the whole dyke, with alternating aplitic levels of variable thickness and pegmatitic bands (Figure 2a).
The intermediate bodies occur further away from the NA batholith than the barren aplite-pegmatites (up to ~500 m-far), constituting subhorizontal leucocratic dykes that intrude discordantly into the host Neoproterozoic-Lower Cambrian metasediments (Figure 1c and Figure 2b). These aplite-pegmatites seem to be scarcer than the barren bodies, since only seven bodies have been identified. They show variable thickness (~1–8 m) and no internal concentric zoning. Nevertheless, some bodies display a rhythmic layering, where coarse crystals show comb-textures (Figure 2c). Intermediate aplite-pegmatites are constituted mainly by albite and quartz, with minor K-feldspar and Al-micas. Topaz, Li-Al-phosphates, Nb-Ta-Sn-oxides, and fluorapatite appear as accessory phases (Table 1).
The lithium-rich dykes correspond to highly evolved aplite-pegmatites, comprising a dozen bodies that appear up to 1 km away from the monzogranite (Figure 1c and Figure 2d). These dykes display high contents of Li-rich micas, feldspars, topaz, and amblygonite-montebrasite series. These bodies are characterized by the lack of internal concentric zoning, showing instead a more or less complex layering (Figure 2e,f).

3. Materials and Methods

The studied oxide minerals were selected from the three aplite-pegmatite types distinguished in the Tres Arroyos field, and to complete the sampling, a few of the studied specimens were collected from the dumps. Over 70 microprobe analyses were obtained from polished thin sections using a JEOL Superprobe JXA-8900 M electron microprobe (EM) at the “Centro de Microscopía Electrónica Luis Bru” from the Complutense University of Madrid (Madrid, Spain), and a CAMECA SX100 at the Scientific-Technical Services of the University of Oviedo (Oviedo, Spain). The operating conditions for the first laboratory were a voltage of 15 kV and beam current of 20 nA, and the calibrations standards used were lithium niobate and tantalite for Nb and Ta, kaersutite for Mg, almandine for Mn and pure metals for W, Ti, Fe and Sn. The operating conditions for the second were a voltage of 20 kV and beam current of 20 nA, and the calibrations standards used were apatite for Ca, modified titanite for Ti, MnS for Mn, magnetite for Fe, pure metals for Nb, Sn and Ta, modified MnWO4 for W and LiF for F. Data were reduced using the procedure of Pouchou and Pichoir [54], and analytical errors were estimated to be of the order of ±1–2% for major elements and ±10% for minor elements. The structural formulae of CT oxides and cassiterite were calculated on the basis of 6 and 4 oxygens, respectively.
Backscattered electron (BSE) images of oxides were obtained in the SGIker facility of the Basque Country University (UPV/EHU). The samples were previously polished and then covered with a carbon layer (20 nm) in order to obtain a conductive surface. The measurements were carried out by a JEOL JSM-6400 scanning electron microscope with a detector EDX Oxford Instruments Inca Energy 350. The operating conditions were a voltage of 20 kV and beam current of 10 nA. All images were taken with BSE to obtain compositional variations.
Fourteen bulk-rock samples of approximately 2–3 kg were collected and were jaw crushed to a nominal grain size of 5 mm. The crushed sample material was split using a riffle splitter to obtain a representative sub-sample, then milled to a grain size of approximately 100 microns (µm) using a vibratory micro mill (Pulverisette 0, Fritsch, Idar-Oberstein, Germany)). Ten representative bulk samples of aplite-pegmatites were analyzed for major-and trace-elements at the Activation Laboratories Ltd. (Actlabs, ON, Canada). Major elements and Zr values were provided by fusion ICP, whereas trace elements data were obtained by ICP-MS. Fluorine and Li values were obtained by fusion specific ion electrode (ISE) and by sodium peroxide fusion, respectively.
Measurements for the geochronological study were realized by laser-ablation quadrupole inductively-coupled plasma mass-spectrometry (LA-Q-ICP-MS) in the SGIker facility of the Basque Country University (UPV/EHU), using the UP213 laser system and the XSeries 2 ICP-MS. The laser beam was fixed to a 55 μm-wide square section. The spot was pre-ablated for 25 s using a laser repetition rate of 10 Hz and 40% output energy, followed by an ablation of 45 s at 10 Hz with a laser output energy of 75%. Zircon GJ_1 [55] was used for the calibration and correction every five measurements of the analyzed samples. Data reduction was carried out with Iolite 3.32 software (version, manufacturer, city, state abbreviation, country) [56] and VizualAge_UcomPbine (version, manufacturer, city, state abbreviation, country) [57]. See García de Madinabeitia et al. [58] for more details.

4. Results

4.1. Petrography of Nb-Ta-Sn Oxides

Columbite-tantalite group minerals from barren and intermediate aplite-pegmatites of Tres Arroyos occur as small-sized (<2 mm), elongated-prismatic to tabular, anhedral-subhedral crystals, usually with corrosion gulfs (Figure 3a–d; Table 1). Backscattered electron images reveal a magmatic concentric zoning, as well as irregular “patches” crosscutting the limits of these magmatic zones (Figure 3a,b). Although scarcer, the biggest-sized CT crystals (<3 mm) are found in the Li-rich aplite-pegmatites (Figure 2e). Some of them show a prismatic to acicular habit with well-developed primary magmatic concentric zoning (Figure 3e,f; Table 1).
In the Tres Arroyos field, cassiterite occurs also as an accessory phase and shows different textural features depending on the aplite-pegmatite type (Table 1). It appears as fine to medium-grained, subhedral, dark crystals (<2 mm) in barren and intermediate aplite-pegmatites, whereas in the most evolved aplite-pegmatites, cassiterite is anhedral to subhedral with a coarser grain size (<5 mm) (Figure 4a). Cassiterite usually displays a typical strong pleochroism and chromatic zoning, with alternating dark (dark-brown to reddish-brown) and pale bands (light-brown to yellow; Figure 4b).

4.2. Mineral Chemistry

Overall, the CT oxides from the Tres Arroyos aplite-pegmatites correspond to columbite-(Fe) and columbite-(Mn) terms (Figure 5), with significant chemical variations depending on the associated aplite-pegmatite type (Table 2; Table S2). The XMn ratio increases from barren, through intermediate, up to Li-rich aplite-pegmatites, with a remarkable Mn-enrichment (avg. XMn = 0.011 ± 0.004 (n = 11), 0.523 ± 0.166 (n = 19) and 0.986 ± 0.011 (n = 8), respectively). The XTa ratio shows a more complex trend, with an initial increase from barren (avg. XTa = 0.180 ± 0.161) to intermediate aplite-pegmatites (avg. XTa = 0.395 ± 0.130), and then decreasing to the most evolved bodies (avg. XTa = 0.211 ± 0.020).
Columbite-tantalite crystals from the barren aplite-pegmatites mainly correspond to the columbite-(Fe) term, except one analysis which corresponds to tantalite-(Fe), with XMn values varying in a very narrow range (0.005–0.017), whereas the XTa ratio shows a much wider one (0.053–0.581) (Figure 5; Table 2; Table S2). In the intermediate bodies, CT oxides display more variable values for XMn and XTa (0.134–0.931 and 0.118–0.622, respectively; Table 2; Table S2). Accordingly, most of the CT crystals from the intermediate bodies may be classified as columbite-(Fe) or columbite-(Mn). Only a few analyses plot in the field of tantalite-(Mn) and tantalite-(Fe) (Figure 5). In contrast, XMn and XTa values for the Li-rich aplite-pegmatites show very narrow ranges (0.960–0.993 and 0.194–0.257, respectively), with a considerable enrichment in Mn but a slight decrease in Ta compared to CT oxides from the intermediate dykes (Figure 5; Table 2; Table S2). In short, in the most fractionated aplite-pegmatites, CT crystals belong to the columbite-(Mn) term (Figure 5).
The sum of positive charges of the analyzed samples from Tres Arroyos is in the range 11.994–12.000 (Table 2; Table S2). Titanium, W, and Sn are the main trace elements in the CT oxides from Tres Arroyos. The Ti contents are higher in CT from barren aplite-pegmatites (up to 5.83 wt% TiO2) than in the other types (<1.49 wt% TiO2) (Table 2; Table S2). The highest W values are also mainly associated with CT from barren aplite-pegmatites (2.67–4.71 wt% WO3) (with the exception of one analysis related to irregular patches with WO3 = 1.34 wt%). The lowest W values are observed in CT from Li-richest aplite-pegmatites (0.34–0.60 wt% WO3), and mid-range values correspond to Nb-Ta oxides from the intermediate bodies (0.36–2.55 wt% WO3) (with the exception of one analyses, that shows a WO3 value of 5.45 wt%) (Table 2; Table S2).
Like the Nb-Ta oxides, cassiterite also presents compositional variations depending on the different types of aplite-pegmatites. Some analyzed cassiterite crystals are close to the ideal SnO2 composition, whereas others deviate significantly with SnO2 contents as low as 90.01 wt% (Table 3; Table S3). Cassiterite associated with the Li-rich bodies is overall the purest one, with average wt% SnO2 values close to 98%. Niobium, Ta, Fe, Mn, Ti, and W are the main elements substituting for Sn (Table 3; Table S3). The highest contents belong to Ta and Nb, which range from 0.00 to 5.80 wt% Ta2O5 and from 0.00 to 3.45 wt% Nb2O5. The rest of the elements occur in general in lower proportions. Iron contents are higher in the barren and intermediate aplite-pegmatites, where it may reach up to 0.86 and 1.55 wt% FeO, respectively, whereas in the Li-rich aplite-pegmatites Fe contents are much lower, usually <0.1 wt% FeO (Table 3; Table S3). Only in this case does Mn predominate over Fe, with values up to 0.6 wt% MnO (Table 3; Table S3). In general, Mn contents are very low, only exceptionally >0.1 wt% MnO (Table 3; Table S3). The sum of cationic charges for all the analyzed cassiterite crystals is in the range of 7.95–8.00 (Table 3; Table S3). Titanium values for cassiterite are higher in the barren aplite-pegmatites (0.61–0.85 wt% TiO2) than in the cassiterite from the intermediate and Li-rich aplite-pegmatites, where TiO2 contents are in general <0.1 wt% (Figure 6a). The opposite behavior is observed for W, which presents the highest values in the cassiterite crystals from the Li-rich aplite-pegmatites, with values up to 0.98 wt% WO3 (Table 3; Table S3).

4.3. Geochronology

Columbite-tantalite crystals of two intermediate bodies were analyzed by the LA-Q-ICP-MS method and U-Th-Pb systematics for the age determination. Seventeen discordant analyses from two samples were selected, which display variable common Pb. After discarding a possible 207Pb correction (cf. [57]), the regression of the data allowed us to provide a lower intercept age of 305 ± 9 Ma for the studied aplite-pegmatite bodies (Figure 7; Table 4).

4.4. Whole-Rock Geochemistry

A collection of 14 rock samples were selected for the whole-rock geochemical characterization of the aplite-pegmatites from Tres Arroyos (Table 5). Overall, the aplite-pegmatites show lower SiO2 (67.17–73.74 wt%) and K2O (0.75–4.00 wt%) contents than the NA monzogranite, with a significantly wider range of K2O (Table 5). Compared with the granites, they also show lower MgO, Fe2O3t, TiO2, and CaO contents, but they display markedly higher values in Al2O3 (16.13–18.50 wt%) and Na2O (4.04–7.54 wt%). The high Na2O contents of the three types of aplite-pegmatites result in lower K2O/Na2O ratios (<1) than those of the granites (>1). Due to their low CaO and high P contents (P2O5 = 0.46–2.50 wt%), all aplite-pegmatites correspond to perphosphorous rocks. They exhibit highly peraluminous compositions, with remarkably higher A/CNK values ([62]) for the Li-rich dykes (1.54–1.60) compared to those of the barren and intermediate bodies (1.18–1.32). The Li-rich bodies show also the highest F contents (0.04–1.80 wt%). The B-A diagram of Debon and Lefort [63], modified by Villaseca et al. [64], illustrates very well the peraluminous trend and low Fe + Mg + Ti values of the aplite-pegmatites relative to monzogranites (Figure 8a). Positive Ta and Nb and negative K, Ti, Y, Nd and REE anomalies are distinctive features of the aplite-pegmatites compared to the monzogranite in the upper continental crust-normalized ([65]) spider diagrams (Figure 8b). The patterns of the least incompatible elements in these multielemental diagrams show clear differences between the NA monzogranite, the marginal granitic facies and the aplite-pegmatites (Figure 8b). The K/Rb ratio is a good parameter to discriminate the distinguished lithotypes, with average values decreasing from the NA monzogranite through the marginal granite, barren and intermediate aplite-pegmatites, up to the Li-rich dykes (125, 53.7, 24.96, 25.8 and 7.24, respectively) (Figure 8c–h). Values of Fe + Mg + Mn and ΣREE contents decrease (as Ba, Y, Zr and Th), and trace elements, such as Nb, Ta, Sn, Rb, Ge, and Li, increase, with decreasing K/Rb, from the NA monzogranite up to the most evolved aplite-pegmatites (Figure 8c–h; Table 5). The Sr shows a more complex behavior because it decreases from the monzogranite to the marginal granite and barren aplite-pegmatites, and then increases from barren, through intermediate, up to the Li-rich dykes (Table 5).

5. Discussion

5.1. Contraints on Mineral Chemistry of Nb-Ta-Sn Oxides

5.1.1. Substitution Mechanisms in Nb-Ta-Sn Oxides

Chemical compositions of the members of the columbite-tantalite series can be expressed with the general formula AB2O6, where A and B positions are mainly occupied by Fe2+ and/or Mn2+ and Nb5+ and/or Ta5+, respectively, although Ca2+, Mg2+, Fe3+, Sc3+, Ti4+, Sn4+, and W6+ elements may also enter (e.g., [66]). Taking into account the sum of positive charges of the analyzed samples from Tres Arroyos (Table 2; Table S2), the amount of Fe3+ in the studied CT can be considered negligible. Plotting Mn2+ vs. Fe2+ (Figure 9a), the clear deviation of the data from the main trend for the CT oxides of barren aplite-pegmatites reflects the presence of other elements in the A site, in addition to Fe and Mn. Contrarily, Nb and Ta seem to exclusively fill the B site in most of the analyzed samples (Figure 9b). The plot 2(Sn+Ti) + W + vacancy vs. 2(Nb+Ta) + 2(Fe+Mn) (Figure 9c) shows that the substitutional scheme that appears to exert the dominant control over the incorporation of Ti and W into these oxides is the (Ti4+,Sn4+)2W6+1vacancy1(Nb5+,Ta5+) exchange vector, particularly for CT associated with barren aplite-pegmatites, with a minor influence in the intermediate and Li-rich dykes.
Similarly to Nb-Ta oxides, the sum of cationic charges for all the analyzed cassiterite crystals indicates that the amount of Fe3+ is likely to be negligible. This is characteristic of pegmatitic cassiterite, since in hydrothermal cassiterites the presence of Fe3+ becomes important (e.g., [60,67]). Moreover, the predominance of Mn over Fe seems to be a feature of extremely fractionated pegmatites [60]. The decrease in Ti in cassiterite from the less evolved bodies to the highly evolved ones has been previously observed in other aplite-pegmatite fields from the CIZ, and related to an increase in the fractionation degree (e.g., [12]). The incorporation of Nb, Ta, Fe, and Mn into cassiterite seems to be controlled by the substitution (Nb5+,Ta5+)2(Fe2+,Mn2+)1Sn4+−3 (Figure 6b). However, some cassiterite crystals from the Li-rich bodies deviate from this vector, indicating that other substitutions may have also been operative. For example, taking into account the small difference in cationic radii between Li+ (0.82Å) and Sn4+ (0.77Å), a possible incorporation of Li, via the substitution (Nb5+, Ta5+)3Li+Sn4+−4 could balance the excess of charge in cassiterite generated by the presence of Nb and/or Ta (Figure 6c) [68]. Another substitution mechanism proposed for cassiterite is Sn4+−2Fe3+Ta5+ [67], but the negligible amount of Fe3+ inferred from the chemical formula calculation excludes this mechanism.

5.1.2. Fractionation of Nb-Ta and Fe-Mn in Columbite-Tantalite Group Minerals

Fractionation of Nb-Ta and Fe-Mn in columbite-tantalite group minerals can be controlled by several factors, including the crystallization of different mineral phases during magma differentiation (e.g., [2,3,17]). The Ta/Nb ratio of CT oxides is expected to increase with fractionation (e.g., [15,16,17]). The solubility of HFSE (Sc, Y, Th, U, Pb, Zr, Hf, Nb, Ta, Ti) seems to be controlled by the temperature and abundance of fluxing elements, such as H2O, Li, B, P, and F [17]. It is experimentally proved that the solubility degree of MnTa2O6 is higher than that of MnNb2O6 in H2O-saturated peraluminous and metaluminous granitic melts [69], and hence columbite-(Mn) should crystallize before tantalite-(Mn). Fluorine, together with Li and P to a lesser extent, may play an important role in the solubility of Ta relative to Nb, with an important increase in the Nb and Ta solubilities when the content in F±Li±P increases (e.g., [3,70,71,72,73,74]). In a “normal” fractionation sequence, the gradual increase in the F±Li±P content would lead to a higher solubility of both elements, but mostly in Ta [73], increasing the Ta/Nb ratio in the melt with fractionation. However, in the case of the CT oxides from Tres Arroyos, the Ta/Nb ratio firstly increases from the barren to the intermediate aplite-pegmatites, but later decreases from the intermediate to the Li-rich dykes. This lack of a progressive Ta/Nb increase could be related to the high F and P availability in the melts from the beginning of the pegmatite crystallization. Evidence for high F and P contents is provided by the presence of topaz and phosphates in the three aplite-pegmatite types, with a minimum of 2.5–3.0 wt% F in the melts (e.g., [49,75,76]). This would forestall the differential solubility of Nb and Ta during fractionation. The Ta/Nb decrease observed in the Li-rich aplite-pegmatites could be caused by the depletion of F, Li, and P in the melt by the crystallization of Li-F-bearing minerals, and subsequent reduction in the solubility of both Nb and Ta. Nevertheless, this decrease could be enough for Nb saturation, but not for the saturation of the (more soluble) Ta, which would remain mainly in the melt. Late metasomatic processes could be responsible for the development of Ta-rich irregular “patches” cross-cutting and masking the primary zoning of many CT crystals from barren and intermediate aplite-pegmatites, as it is found in other pegmatitic belts (e.g., [77,78,79]). Recent experimental works, nevertheless, question the role of fluorine in the partition of Ta/Nb ([80,81]). Some researchers, such as Stepanov et al. [2] or López-Moro et al. [13], suggested that the peraluminosity and F-content have limited consequences in the partition of Ta/Nb, with the crystallization of muscovite being determinant for the fractionation of these elements. Thus, CT oxides could reflect certain control of micas on Ta and Nb fractionation in the parental melt, as suggested by Raimbault and Burnol [82] and Stepanov et al. [2] for some granitic melts and Llorens González et al. [83] for some Nb-Ta mineralizations from the CIZ. Moreover, average values of the Ta/Nb ratios in Al-micas of barren (0.19), intermediate (0.69) and Li-rich aplite-pegmatites (0.41) from Tres Arroyos [47] would correspond to the trend observed for the Ta/Nb ratios in CT oxides.
The controlling factors that lead to the fractionation of Fe-Mn are not consensual (e.g., [9] and references therein). The solubilities of the Fe-bearing CT members are more complex due to their dependence on redox conditions [17]. Iron-rich end-members are an order of magnitude more soluble than Mn-rich ones and, therefore, an enrichment in Fe with magma evolution should be expected (e.g., [71]). However, usually Mn values increase in pegmatitic bodies as fractionation proceeds, parallel to an increase in the XTa ratio (e.g., [14,15,16]). Some authors related the enrichment of Mn in the melt to the high activity of fluorine (e.g., [15,79,84]). In contrast, other authors suggested that the Fe-Mn contents in the CT oxides are controlled by other phases (e.g., [12,16,17,85,86,87]). Tourmaline, for example, is reported to show Mn-enrichment with fractionation (e.g., [88,89]). However, in Tres Arroyos, tourmaline occurs only in the granitic facies (with no CT oxides) and in the barren aplite-pegmatites. Consequently, early crystallization of tourmaline and phosphates would deplete the melt in Fe, contributing to the observed increase in the Mn/Fe ratio from those lithotypes to the more evolved ones. Nevertheless, tourmaline and phosphates crystallization could not explain the Mn-enrichment in the CT oxides from the most evolved aplite-pegmatites from Tres Arroyos. In this case, Fe enrichment should be expected rather than a Mn increase [87]. Micas were also proposed to control the Mn/Fe in the melt. Based on the muscovite-matrix partition coefficients reported by Raimbault and Burnol [82], a melt will become strongly depleted in Fe relative to Mn with muscovite crystallization. Additionally, the progressive crystallization of zinnwaldite may contribute to an increase in the Mn/Fe ratio, as described for the Beauvoir granite [90]. Similarly to the Ta/Nb ratios, the Mn/Fe ratio in white micas of Tres Arroyos increases progressively from the barren (0.00–0.06) through intermediate (0.00–5.2) to most evolved aplite-pegmatites (0.04–103.44) [47], reflecting the same pattern of CT oxides. It is conceivable, therefore, that the progressive depletion in Fe content could be caused by tourmaline, phosphates and mica crystallization.
As observed in other pegmatitic fields, such as Fregeneda-Almendra [26,29], the predominant chemical variation of CT oxides in Tres Arroyos may be capitalized on by the XMn ratio, which clearly increases with fractionation, whereas the variation in XTa is subordinated and not gradual. The progressive increase in the XMn ratio from barren, through intermediate to the Li-rich aplite-pegmatites is interpreted to reflect a continuum of magmatic evolution. The crystallization sequence will depend on several factors (e.g., [29]), such as: (1) abundance of Fe, Mn, Nb, Ta, and Sn in the system; (2) temperature and undercooling degree; (3) degree of solubility of Nb-Ta oxides, affected by the presence of H2O, F, Li, B, and/or P; and (4) the coexisting mineral assemblage.
Assuming a crystallization temperature for the barren bodies >500 °C [11], the saturation of tantalite could be estimated over 1300 ppm of Ta for a system with a high content of fluxing elements (2 wt% F, 2 wt% Li) and over 120 ppm of Nb [17]. These values are higher than the data of bulk-rock for the Tres Arroyos aplite-pegmatites (54 ppm of Nb and 23 ppm of Ta; 40–118 ppm of Nb and 51–152 ppm of Ta; and 72–93 ppm of Nb and 64–91 ppm of Ta for barren, intermediate and Li-rich aplite-pegmatites, respectively). Contents needed for the saturation in Li-F-free environments are noticeably lower: 70 ppm for Nb and 510 ppm for Ta [17]. This may explain the late crystallization of CT oxides, mainly those richer in Ta, once the F-Li-bearing minerals have crystallized, hence reducing the Nb and Ta values needed for CT oxides saturation. The formation of some primary columbite-(Fe) and columbite-(Mn) could be related to punctual fluctuations in the composition of mineral-forming environment (local melt pools relatively rich in Nb-Ta) during centripetal aplite-pegmatites crystallization, probably with the local development of boundary layers where some incompatible elements are concentrated.

5.1.3. Formation of Primary Cassiterite

The stability of cassiterite depends on different chemical and physical conditions of the melts (e.g., [91] and references therein). According to some authors (e.g., [72,73,82]), the solubility of SnO2 increases considerably if the melt is enriched in elements, such as F or Li. In contrast, Bhalla et al. [91] suggested that the F’s effect in the solubility of SnO2 in granitic melts is almost negligible, especially in reducing environments where Sn is divalent and acts as a HFSE, with the solubility being highly dependent on the fO2, Cl, Al and temperature. In the case of Tres Arroyos, the F content may have been a capital factor in controlling the crystallization of Sn-oxides. Tin would be transported as a fluoride (SnF4) during the magmatic crystallization, and with increasing f(H2O) may precipitate via the reaction SnF4 + 2H2O ↔ SnO2 + 4HF (e.g., [92]). Furthermore, the release of volatiles, including HCl and HF, during the boiling of the system could also lead to the cassiterite saturation (e.g., [93]). This boiling can be caused by decompression of the system (first boiling) or by crystallization of non-hydrous minerals under isobaric conditions (resurgent or second boiling; [94]). In the case of Tres Arroyos, a significant pressure change could be related to the injection of the pegmatitic melts in open fractures, at low-pressure conditions (≈2–3 kbar; [43,46]), once the magmatic system opened and the most fractionated melts, accumulated in the apical part of the magma chamber, were expelled [20,31]. In this sense, Linnen et al. [95] and London [96] suggested that the crystallization of cassiterite in pegmatites may be associated with an undercooling of the system, exceeding the magmatic saturation temperature. Undercooling may be prompted by a sudden P or T change, or by a chemical variation implying an important decrease in fluxing components in the melt (e.g., [19,97]). In the case of Tres Arroyos, the undercooling of the system could be produced by a combination of these three factors. In addition to the just described sudden drop in pressure with the opening of the system (pressure quenching), the heat transfers from the pegmatitic melt to the colder host rocks would imply a thermal undercooling of the melt [48]. Moreover, the possible exsolution of a H2O-rich fluid from the pegmatitic melt associated with the boiling of this fluid could suppose a dramatic T decrease in the remaining melt, as this is an endothermic reaction. A loss of fluxing components from the pegmatitic melt, such as H2O and B (no tourmaline is found inside the intermediate and Li-rich aplite-pegmatite dykes, but it is common in their hosting metasediments as a result of a metasomatic reaction) would provoke a chemical quenching of the melt and the subsequent rise in the solidus [47].

5.2. Petrogenesis of the Aplite-Pegmatites from Tres Arroyos

As Simmons and Webber [97] postulated, there is no a universally accepted model for the granitic pegmatite genesis. However, pegmatitic melt-forming mechanisms can be grouped into two major processes: (1) fractional crystallization of a granitic magma and (2) direct anatexis. In the case of the Tres Arroyos aplite-pegmatite field, the low metamorphic grade of the host metasedimentary rocks [43,46], together with the absence of migmatitic terrains near the studied aplite-pegmatites, and compositional differences between whole-rock analyses of the studied aplite-pegmatites and those of minimum melt fractions from similar country rocks [20] preclude their origin by direct anatexis. The spatial distribution of the distinguished five lithotypes, from the monzogranite of the NA batholith, through the marginal granitic facies, barren aplite-pegmatites, and intermediate aplite-pegmatites, up to the Li-rich aplite-pegmatites, is consistent with a coherent genetic relation between the granitic massif and the aplite-pegmatites. The occurrence of mafic minerals (Fe-tourmaline and/or Fe-Mg-Mn-phosphates) in the barren aplite-pegmatites, as well as that of amblygonite-montebrasite in the intermediate and Li-rich dykes, indicate lower fractionation degrees for the barren and higher ones for the intermediate and Li-rich dykes (e.g., [49,98]). Compositional variations of the main minerals are also coherent with the observed regional zonation of the lithotypes (e.g., [11,47]). The decrease in Ti contents in quartz, as well as the increase and strong correlation of Al/Ti and Ge/Ti ratios, the trends followed by the Fe/(Fe+Mg) in tourmaline and K/Rb ratios in micas and K-feldspar, are consistent with the increase in the fractionation degree, from the NA monzogranite, through the marginal granitic facies, barren and intermediate aplite-pegmatites to Li-rich aplite-pegmatites [11,47]. Likewise, the progressive increase in the Mn/(Mn+Fe) ratio in columbite-tantalite group minerals from barren aplite-pegmatites up to the Li-rich bodies (Figure 5), suggests a single evolutionary sequence for the three types of studied aplite-pegmatites (e.g., [15,26]). Similarly, the higher Ti contents in cassiterites from the barren aplite-pegmatites, compared to those from the intermediate and Li-rich bodies (Figure 6), suggests that the parental melt of the former was less evolved (e.g., [12]). Whole-rock major- and trace-element compositions of the studied samples show a progressive evolution in the same sense of NA monzogranite → marginal granitic facies → barren aplite-pegmatites → intermediate aplite-pegmatites → Li-rich aplite-pegmatites (e.g., K/Rb ratio, Nb, Ta, Sn, Li, REE contents; Figure 8). In addition, geochemical modelling carried out for these aplite-pegmatites indicates that the evolution of the system could be explained by different fractionation degrees (up to ~99%) of quartz + K-feldspar + plagioclase (±biotite, ±muscovite) [11]. Thus, all the above mentioned evidence supports an origin by fractional crystallization for the aplite-pegmatites from Tres Arroyos.
In the case of an origin by fractional crystallization of a granitic magma, the relationship between the pegmatitic melts and their parental granitic magma could be evaluated by means of a geochronological study, since this relationship is frequently obscured by the lack of a spatial and/or chemical continuity between them (e.g., [20,27,28]). The use of CT oxides with this purpose is supported by different facts: (1) columbite-tantalite group minerals cannot be inherited from the source of a granitic magma, (2) they show high U and low common Pb contents, and (3) they occur in pegmatites where zircon and monazite could be absent or inappropriate for dating (e.g., [1,6,32,99,100]). Moreover, the development of techniques, such as LA-ICP-MS, has allowed the measurement of unaltered and inclusion-free crystal zones in these oxides, becoming suitable mineral phases for the dating of pegmatitic bodies (e.g., [6,32,99]). Existing geochronological data on CIZ aplite-pegmatites manifest a diachronic character of these pegmatites based on U-Pb dating of minerals from the CT group, showing a southern propagation of ages in the range of 310–301 Ma ([101]). Based on U-Pb zircon dating, granitic magmatism in the Nisa-Alburquerque-Los Pedroches Magmatic Alignment of the southern CIZ occurred from 314 to 304 Ma, (e.g., [40] and references therein), and the emplacement of the NA batholith was considered to be in the range of 305–309 Ma [44,45]. Thus, the obtained U-Pb age (305 ± 9 Ma) for the Tres Arroyos aplite-pegmatites would be in agreement with that of the granitic magmatism in the southern CIZ and would coincide with that of the latest pulses of the NA batholith and the CIZ pegmatites.
The NA batholith belongs to the P-rich, Ca-poor, highly peraluminous S2 granitic suite defined by Villaseca [102] for the southern realm of the CIZ. Granites belonging to this suite are considered to be derived from Neoproterozoic to Early Cambrian psammopelitic metasediments of the SGC that were Ca-poor and P-(F, Li) rich [20]. The low Ca proportions in the starting melts derived from these metasediments would prevent the crystallization of apatite, allowing P and F to remain in a residual melt together with other fluxing elements. A high concentration of fluxing components such as P, F, B, Li and H2O in the granitic melts is believed to reduce their viscosity and polymerization degree [103] as well as their liquidus temperature. Thereby, internal diffusion, mobility of melts, and their ability to segregate from the source area are increased [104]. The lack of internal zonation in the aplite-pegmatite dykes without appreciable geochemical differences between the distinguished layers constituting the dykes, together with the crystallization of lepidolite across the whole width of the Li-rich dykes, suggest that the fractionation of the pegmatitic melt was insignificant during the crystallization of the aplite-pegmatites (e.g., [11]). Consequently, the fractionation process should be developed before the crystallization of the dykes, either (1) during lateral propagation of the pegmatitic melts or “en route” to the surface as proposed by Tartèse and Boulvais [105], or (2) by fractional crystallization + vertical movement of fluxing elements in the magmatic chamber (NA batholith). Although the lateral fractionation cannot be excluded, the strongly evolved compositions of the Li-rich dykes, together with the short distances from the granitic batholith and fast crystallization rates [11], point to the second hypothesis as the most likely scenario. Assuming this, a vertical zonation of the magmatic chamber should have been developed in Tres Arroyos, as it has been described for other pegmatitic fields in the CIZ [20,106], comprising different zones constituted by the parental magmas of each type of the distinguished aplite-pegmatites. Therefore, at the low emplacement levels of the NA batholith, once the vertical zoning of the magma chamber was developed, fluid pressure could be high enough for the system to open it at the apical part. In this way, the different melt fractions (± some minor mineral phases) would scape and reach up to distances that would vary depending on their location in the chamber, degree of fractionation and mobility. This would result in the zonation observed in the granite‒aplite-pegmatite system of Tres Arroyos, with the Li-F-richest melt portions moving the furthest from the magma source.

6. Conclusions

The main conclusions of this contribution can be summarized as follows:
(1)
The main substitution mechanism for the CT group minerals seems to be (Ti4+,Sn4+)2W6+1vacancy1(Nb5+,Ta5+), whereas that for cassiterite would be (Nb5+,Ta5+)2(Fe2+,Mn2+)1Sn4+-3.
(2)
The Mn/(Mn+Fe) ratio in CT oxides can be used as a petrogenetic indicator in the Tres Arroyos system, suggesting a single sequence of magmatic evolution from the less evolved to the most evolved bodies. The high F and P availability in the parental pegmatitic melts may have caused the lack of an expected progressive Ta/Nb increase with the fractionation.
(3)
The fractionation of Nb-Ta and Fe-Mn in the pegmatitic melts reflected in CT oxides seems to be controlled by tourmaline, phosphates and mica crystallization.
(4)
Some of the columbite-(Fe) and columbite-(Mn) crystals would have been crystallized by punctual chemical variations in the boundary layer, whereas the crystallization of other CT oxides, mainly those richer in Ta, would be controlled by Ta saturation in the pegmatitic melt. On the other hand, the formation of primary cassiterite may have been related to a high undercooling of the system.
(5)
Whole-rock geochemical data of the studied aplite-pegmatites and granitic units mirror the trends defined by mineral chemistry and strengthen the observed geochemical evolution from the NA monzogranite up to the Li-rich bodies.
(6)
The newly obtained U-Pb age of 305 ± 9 Ma on CT minerals from the Tres Arroyos aplite-pegmatites reinforces the genetic relationship between the studied aplite-pegmatites and the NA monzogranite, which would represent the parental granitic magma.

Supplementary Materials

The following are available online at https://www.mdpi.com/2075-163X/10/11/1008/s1. Table S1: Universal Transverse Mercator (UTM) coordinates of the studied samples (Zone 29S); Table S2: Microprobe analyses of columbite-tantalite group minerals from the Tres Arroyos aplite-pegmatites; Table S3: Microprobe analyses of cassiterite from the Tres Arroyos aplite-pegmatites.

Author Contributions

Conceptualization and investigation, I.G.-O. and E.R.-R.; writing—original draft preparation, I.G.-O., E.R.-R. and J.E.-M.; writing—review and editing, P.P.G.-C. and A.P.; visualization, I.G.-O and J.E.-M; project administration, E.R.-R.; funding acquisition, E.R.-R., A.P. and I.G.-O. All authors took part of the field work. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Spanish Ministry of Economy, Industry and Competitiveness (Project RTI2018-094097-B-100, with ERDF funds), the University of the Basque Country UPV/ EHU (grant GIU18/084) and the European Union’s Horizon 2020 Innovation Programme (grant agreement No 869274, project GREENPEG: New Exploration Tools for European Pegmatite Green-Tech Resources). Work of Idoia Garate-Olave has been supported also by the UPV/EHU by means of the “Convocatoria de contratación para la especialización de personal investigador doctor en la UPV/EHU 2019”.

Acknowledgments

S. García de Madinabeitia is gratefully acknowledged for the help on geochronological methods, as well as the IMERYS mining company for the access to the quarries. The authors would like to thank Assistant Editor for the editorial handling, the Academic Editor and two anonymous reviewers for their constructive comments, as well as the SGIker of UPV/EHU and the Microprobe Facility of the University of Oviedo for technical and human support. Helpful comments and suggestions made by David Lentz and Axel Müller in a previous version of the manuscript are also acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kendall-Langley, L.A.; Kemp, A.I.S.; Grigson, J.L.; Hammerli, J. U-Pb and reconnaissance Lu-Hf isotope analysis of cassiterite and columbite group minerals from Archean Li-Cs-Ta type pegmatites of Western Australia. Lithos 2020, 352. [Google Scholar] [CrossRef]
  2. Stepanov, A.; Mavrogenes, J.; Meffre, S.; Davidson, P. The key role of mica during igneous concentration of tantalum. Contrib. Mineral. Petrol. 2014, 167, 1009. [Google Scholar] [CrossRef]
  3. Van Lichtervelde, M.; Holtz, F.; Hanchar, J.M. Solubility of manganotantalite, zircon and hafnon in highly fluxed peralkaline to peraluminous pegmatitic melts. Contrib. Mineral. Petrol. 2010, 160, 17–32. [Google Scholar] [CrossRef]
  4. Küster, D.; Romer, R.L.; Tolessa, D.; Zerihun, D.; Bheemalingeswara, K.; Melcher, F.; Oberthür, T. The Kenticha rare-element pegmatite, Ethiopia: Internal differentiation, U-Pb age and Ta mineralization. Mineral. Depos. 2009, 44, 723–750. [Google Scholar] [CrossRef]
  5. Mackay, D.A.R.; Simandl, G.J. Geology, market and supply chain of niobium and tantalum-a review. Mineral. Depos. 2014, 49, 1025–1047. [Google Scholar] [CrossRef]
  6. Che, X.-D.; Wu, F.-Y.; Wang, R.-C.; Gerdes, A.; Ji, W.-Q.; Zhao, Z.-H.; Yang, J.-H.; Zhu, Z.-Y. In situ U-Pb isotopic dating of columbite-tantalite by LA-ICP-MS. Ore Geol. Rev. 2015, 65, 979–989. [Google Scholar] [CrossRef]
  7. Ballouard, C.; Elburg, M.A.; Tappe, S.; Reinke, C.; Ueckermann, H.; Doggart, S. Magmatic-hydrothermal evolution of rare metal pegmatites from the Mesoproterozoic Orange River pegmatite belt (Namaqualand, South Africa). Ore Geol. Rev. 2020, 116, 103252. [Google Scholar] [CrossRef]
  8. Breiter, K.; Ackerman, L.; Ďurišová, J.; Svojtka, M.; Novák, M. Trace element composition of quartz from different types of pegmatites: A case study from the Moldanubian Zone of the Bohemian Massif (Czech Republic). Mineral. Mag. 2014, 78, 703–722. [Google Scholar] [CrossRef]
  9. Fuchsloch, W.C.; Nex, P.A.M.; Kinnaird, J.A. The geochemical evolution of Nb–Ta–Sn oxides from pegmatites of the Cape Cross–Uis pegmatite belt, Namibia. Mineral. Mag. 2019, 83, 161–179. [Google Scholar] [CrossRef]
  10. Galliski, M.A.; Marquez-Zavalia, M.F.; Černý, P.; Martínez, V.A.; Chapman, R. The Ta-Nb-Sn-Ti oxide-mineral paragenesis from La Viquita, a spodumene-bearing rare-element granitic pegmatite, San Luis, Argentina. Can. Mineral. 2008, 46, 379–393. [Google Scholar] [CrossRef]
  11. Garate-Olave, I.; Müller, A.; Roda-Robles, E.; Gil-Crespo, P.P.; Pesquera, A. Extreme fractionation in a granite–pegmatite system documented by quartz chemistry: The case study of Tres Arroyos (Central Iberian Zone, Spain). Lithos 2017, 286, 162–164. [Google Scholar] [CrossRef] [Green Version]
  12. Llorens, T.; Moro, M.C. Oxide minerals in the granitic cupola of the Jálama Batholith, Salamanca, Spain. Part II: Sn, W and Ti minerals in intra-granitic quartz veins. J. Geosci. 2012, 155–171. [Google Scholar] [CrossRef] [Green Version]
  13. López-Moro, F.J.; Garcia Polonio, F.; Llorens González, T.; Sanz Contreras, J.L.; Fernández Fernández, A.; Moro Benito, M.C. Ta and Sn concentration by muscovite fractionation and degassing in a lens-like granite body: The case study of the Penouta rare-metal albite granite (NW Spain). Ore Geol. Rev. 2017, 82, 10–30. [Google Scholar] [CrossRef]
  14. Novák, M.; Černý, P.; Uher, P. Extreme variation and apparent reversal of Nb-Ta fractionation in columbite-group minerals from the Scheibengraben beryl-columbite granitic pegmatite, Maršíkov, Czech Republic. Eur. J. Mineral. 2003, 15, 565–574. [Google Scholar] [CrossRef]
  15. Černý, P.; Goad, B.E.; Hawthorne, F.C.; Chapman, R. Fractionation trends of the Nb- and Ta-bearing oxide minerals in the Greer Lake pegmatitic granite and its pegmatite aureole, southeastern Manitoba. Am. Mineral. 1986, 71, 501–517. [Google Scholar]
  16. Mulja, T.; Williams-Jones, A.E.; Martin, R.F.; Wood, S.A. Compositional variation and structural state of columbite-tantalite in rare-element granitic pegmatites of the Preissac-Lacorne Batholith, Quebec, Canada. Am. Mineral. 1996, 81, 146–157. [Google Scholar] [CrossRef]
  17. Linnen, R.L.; Cuney, M. Granite-related rare-element deposits and experimental constraints on Ta-Nb-W-Sn-Zr-Hf mineralization. In Rare-Element Geochemistry and Mineral Deposits; Linnen, R.L., Sammson, I.M., Eds.; Geological Society: London, UK, 2005; pp. 45–68. [Google Scholar]
  18. Černý, P.; Ercit, T.S. The classification of granitic pegmatites revisited. Can. Mineral. 2005, 43, 2005–2026. [Google Scholar] [CrossRef] [Green Version]
  19. London, D. Pegmatites; The Canadian Mineralogist, Special Publication nº 10: Ottawa, ON, Canada, 2008; p. 347. [Google Scholar]
  20. Roda-Robles, E.; Villaseca, C.; Pesquera, A.; Gil-Crespo, P.P.; Vieira, R.; Lima, A.; Garate-Olave, I. Petrogenetic relationships between Variscan granitoids and Li-(F-P)-rich aplite-pegmatites in the Central Iberian Zone: Geological and geochemical constraints and implications for other regions from the European Variscides. Ore Geol. Rev. 2018, 95, 408–430. [Google Scholar] [CrossRef]
  21. Breiter, K.; Durisova, J.; Hrstka, T.; Korbelova, Z.; Galiova, M.V.; Muller, A.; Simons, B.; Shail, R.K.; Williamson, B.J.; Davies, J.A. The transition from granite to banded aplite-pegmatite sheet complexes: An example from Megiliggar Rocks, Tregonning topaz granite, Cornwall. Lithos 2018, 302, 370–388. [Google Scholar] [CrossRef]
  22. Galliski, M.Á.; Márquez-Zavalía, M.F.; Pagano, D.S. Metallogenesis of the Totoral LCT rare-element pegmatite district, San Luis, Argentina: A review. J. S. Am. Earth. Sci. 2019, 90, 423–439. [Google Scholar] [CrossRef]
  23. Müller, A.; Romer, R.L.; Pedersen, R.-B. The Sveconorwegian Pegmatite Province-Thousands of Pegmatites Without Parental Granites. Can. Mineral. 2017, 55, 283–315. [Google Scholar] [CrossRef] [Green Version]
  24. Konzett, J.; Schneider, T.; Nedyalkova, L.; Hauzenberger, C.; Melcher, F.; Gerdes, A.; Whitehouse, M. Anatectic Granitic Pegmatites from the Eastern Alps: A Case of Variable Rare-Metal Enrichment During High-Grade Regional Metamorphism–I: Mineral Assemblages, Geochemical Characteristics, and Emplacement Ages. Can. Mineral. 2018, 56, 555–602. [Google Scholar] [CrossRef]
  25. Feng, Y.; Liang, T.; Yang, X.; Zhang, Z.; Wang, Y. Chemical Evolution of Nb-Ta Oxides and Cassiterite in Phosphorus-Rich Albite-Spodumene Pegmatites in the Kangxiwa–Dahongliutan Pegmatite Field, Western Kunlun Orogen, China. Minerals 2019, 9, 166. [Google Scholar] [CrossRef] [Green Version]
  26. Roda-Robles, E.; Pesquera Pérez, A.; Velasco Roldan, F.; Fontan, F. The granitic pegmatites of the Fregeneda area (Salamanca, Spain): Characteristics and petrogenesis. Mineral. Mag. 1999, 63, 535–558. [Google Scholar] [CrossRef]
  27. Vieira, R.; Roda-Robles, E.; Pesquera, A.; Lima, A. Chemical variation and significance of micas from the Fregeneda-Almendra pegmatitic field (Central-Iberian Zone, Spain and Portugal). Am. Mineral. 2011, 96, 637–645. [Google Scholar] [CrossRef]
  28. Tang, Y.; Zhao, J.-Y.; Zhang, H.; Cai, D.-W.; Lv, Z.-H.; Liu, Y.-L.; Zhang, X. Precise columbite-(Fe) and zircon U-Pb dating of the Nanping No. 31 pegmatite vein in northeastern Cathaysia Block, SE China. Ore Geol. Rev. 2017, 83, 300–311. [Google Scholar] [CrossRef]
  29. Vieira, R. Aplitopegmatitos com Elementos Raros da Região Entre Almendra (V. N. de Foz-Côa) e Barca d’Alba (Figueira de Castelo Rodrigo). Campo Aplitopegmatítico da Fregeneda-Almendra. Ph.D. Thesis, Universidade do Porto, Porto, Portugal, 2010. [Google Scholar]
  30. Antunes, I.M.H.R.; Neiva, A.M.R.; Farinha Ramos, J.M.; Silva, P.B.; Silva, M.M.V.G.; Corfu, F. Petrogenetic links between lepidolite-subtype aplite-pegmatite, aplite veins and associated granites at Segura (central Portugal). Chem. Erde Geochem. 2013, 73, 323–341. [Google Scholar] [CrossRef] [Green Version]
  31. Roda-Robles, E.; Pesquera, A.; Gil-Crespo, P.P.; Vieira, R.; Lima, A.; Garate-Olave, I.; Martins, T.; Torres-Ruiz, J. Geology and mineralogy of Li mineralization in the Central Iberian Zone (Spain and Portugal). Mineral. Mag. 2016, 80, 103–126. [Google Scholar] [CrossRef] [Green Version]
  32. Legros, H.; Mercadier, J.; Villeneuve, J.; Romer, R.L.; Deloule, E.; Van Lichtervelde, M.; Dewaele, S.; Lach, P.; Che, X.-D.; Wang, R.-C.; et al. U-Pb isotopic dating of columbite-tantalite minerals: Development of reference materials and in situ applications by ion microprobe. Chem. Geol. 2019, 512, 69–84. [Google Scholar] [CrossRef] [Green Version]
  33. Martínez Catalán, J.R.; Fernández-Suárez, J.; Jenner, G.A.; Belousova, E.; Díez Montes, A. Provenance constraints from detrital zircon U-Pb ages in the NW Iberian Massif: Implications for Palaeozoic plate configuration and Variscan evolution. J. Geol. Soc. 2004, 161, 463–476. [Google Scholar] [CrossRef]
  34. Martínez Catalán, J.R.; Rubio Pascual, F.J.; Díez Montes, A.; Díez Fernández, R.; Gómez Barreiro, J.; Dias Da Silva, I.; González Clavijo, E.; Ayarza, P.; Alcock, J.E. The late Variscan HT/LP metamorphic event in NW and Central Iberia: Relationships to crustal thickening, extension, orocline development and crustal evolution. In Variscan Orogeny: Extent, Timescale and the Formation of the European Crust; Schulmann, K., Martínez Catalán, J.R., Lardeaux, J.M., Janoušek, V., Oggiano, G., Eds.; Geological Society: London, UK, 2014; Volume 405, pp. 225–247. [Google Scholar]
  35. Talavera, C.; Martínez Poyatos, D.; Francisco, G. SHRIMP U–Pb geochronological constraints on the timing of the intra-Alcudian (Cadomian) angular unconformity in the Central Iberian Zone (Iberian Massif, Spain). Int. J. Earth Sci. 2015, 104, 1739–1757. [Google Scholar] [CrossRef]
  36. Bea, F.; Montero, P.; González-Lodeiro, F.; Talavera, C. Zircon inheritance reveals exceptionally fast crustal magma generation processes in central iberia during the cambro-ordovician. J. Petrol. 2007, 48, 2327. [Google Scholar] [CrossRef] [Green Version]
  37. Martínez-Poyatos, D.; Carbonell, R.; Palomeras, I.; Simancas, J.F.; Ayarza, P.; Martí, D.; Azor, A.; Jabaloy, A.; González-Cuadra, P.; Tejero, R.; et al. Imaging the crustal structure of the Central Iberian Zone (Variscan Belt): The ALCUDIA deep seismic reflection transect. Tectonics 2012, 31, TC3017. [Google Scholar] [CrossRef] [Green Version]
  38. Villaseca, C.; Merino Martínez, E.; Orejana, D.; Andersen, T.; Belousova, E. Zircon Hf signatures from granitic orthogneisses of the Spanish Central System: Significance and sources of the Cambro-Ordovician magmatism in the Iberian Variscan Belt. Gondwana Res. 2016, 34, 60–83. [Google Scholar] [CrossRef]
  39. Castro, A.; Patiño Douce, A.E.; Corretgé, L.G.; De La Rosa, J.D.; El-Biad, M.; El-Hmidi, H. Origin of peraluminous granites and granodiorites, Iberian massif, Spain: An experimental test of granite petrogenesis. Contrib. Mineral. Petrol. 1999, 135, 255–276. [Google Scholar] [CrossRef]
  40. Errandonea-Martin, J.; Sarrionandia, F.; Janoušek, V.; Carracedo-Sánchez, M.; Gil-Ibarguchi, J.I. Origin of cordierite-bearing monzogranites from the southern Central Iberian Zone-Inferences from the zoned Sierra Bermeja Pluton (Extremadura, Spain). Lithos 2019, 342, 440–462. [Google Scholar] [CrossRef]
  41. González-Menéndez, L. Petrología del batolitio granítico de Nisa-Alburquerque. Rev. Soc. Geol. España 1998, 15, 233–246. [Google Scholar]
  42. Solá, A.R.; Ribeiro, M.L.; Moreira, M.E.; Moreira, M. Complexo eruptivo de Nisa cartografia geoquímica e mecanismo de implantação. In Actas do V Congresso Nacional de Geologia; Instituto Geológico e Mineiro: Lisboa, Portugal, November 1998; pp. B39–B42. [Google Scholar]
  43. González-Menéndez, L.; Azor, A.; Rubio-Ordóñez, Á.; Sánchez-Almazo, I. The metamorphic aureole of the Nisa-Alburquerque batholith (SW Iberia): Implications for deep structure and emplacement mode. Int. J. Earth Sci. 2011, 100, 1533–1550. [Google Scholar] [CrossRef]
  44. Gutiérrez-Alonso, G.; Fernández-Suárez, J.; Jeffries, T.E.; Johnston, S.T.; Pastor-Galán, D.; Murphy, J.B.; Franco, M.P.; Gonzalo, J.C. Diachronous post-orogenic magmatism within a developing orocline in Iberia, European Variscides. Tectonics 2011, 30, TC5008. [Google Scholar] [CrossRef] [Green Version]
  45. Solá, A.R.; Williams, I.S.; Neiva, A.M.R.; Ribeiro, M.L. U-Th-Pb SHRIMP ages and oxygen isotope composition of zircon from two contrasting late Variscan granitoids, Nisa-Albuquerque batholith, SW Iberian Massif: Petrologic and regional implications. Lithos 2009, 111, 156–167. [Google Scholar] [CrossRef]
  46. Gallego Garrido, M. Las mineralizaciones de Li asociadas a magmatismo ácido en Extremadura y su encuadre en la Zona Centro-Ibérica. Ph.D. Thesis, Complutense University of Madrid, Madrid, Spain, 1992. [Google Scholar]
  47. Garate-Olave, I.; Roda-Robles, E.; Gil-Crespo, P.P.; Pesquera, A. Mica and feldspar as indicators of the evolution of a highly evolved granite-pegmatite system in the Tres Arroyos area (Central Iberian Zone, Spain). J. Iberian Geol. 2018, 44, 375–403. [Google Scholar] [CrossRef]
  48. London, D.; Wolf, M.B.; Morgan, G.B.; Gallego-Garrido, M. Experimental Silicate–Phosphate Equilibria in Peraluminous Granitic Magmas, with a Case Study of the Alburquerque Batholith at Tres Arroyos, Badajoz, Spain. J. Petrol. 1999, 40, 215–240. [Google Scholar] [CrossRef]
  49. Garate-Olave, I.; Roda-Robles, E.; Gil-Crespo, P.P.; Pesquera, A. The phosphate mineral associations from the Tres Arroyos aplite-pegmatites (Badajoz, Spain): Petrography, mineral chemistry and petrogenetic implications. Can. Mineral. 2020, in press. [Google Scholar]
  50. Farias, P.; Gallastegui, G.; González-Lodeiro, F.; Marquínez, J.; Martín-Parra, L.M.; Martínez Catalán, J.R.; de Pablo Maciá, J.G.; Rodríguez-Fernández, L.R. Aportaciones al conocimiento de la litoestratigrafia y estructura de Galicia central. Mem. Museo e Lab. Miner. Geol. Fac. Ciencias, Univ. Oporto 1987, 1, 411–431. [Google Scholar]
  51. Julivert, M.; Marcos, A.; Truyols, J. L’évolution paléogéographique du NW de l’Espagne pendant l’Ordovicien–Silurien. Bull. Soc. Geol. Min. Bret. 1972, 4, 1–7. [Google Scholar]
  52. Lotze, F. Zur Gliederung der Varisziden in der Iberischen Meseta. Geoteckt Forsch 1945, 6, 78–92. [Google Scholar]
  53. Rodríguez Fernández, L.R.; Oliveira, J.T. Mapa Geológico de España y Portugal a escala 1:1.000.000; Instituto Geológico y Minero de España, Laboratório Nacional de Energía e Geología de Portugal: Madrid, Spain, 2015. [Google Scholar]
  54. Pouchou, J.L.; Pichoir, F. “PAP” φ(ρZ) procedure for improved quantitative microanalysis. In Microbean Analysis; Armstrong, J.T., Ed.; San Francisco Press: San Francisco, CA, USA, 1985; pp. 104–106. [Google Scholar]
  55. Jackson, S.E.; Pearson, N.J.; Griffin, W.L.; Belousova, E.A. The application of laser ablation-inductively coupled plasma-mass spectrometry to in situ U-Pb zircon geochronology. Chem. Geol. 2004, 211, 47–69. [Google Scholar] [CrossRef]
  56. Paton, C.; Hellstrom, J.; Paul, B.; Woodhead, J.; Hergt, J. Iolite: Freeware for the visualisation and processing of mass spectrometric data. J. Anal. At. Spectrom. 2011, 26. [Google Scholar] [CrossRef]
  57. Chew, D.M.; Petrus, J.A.; Kamber, B.S. U-Pb LA-ICPMS dating using accessory mineral standards with variable common Pb. Chem. Geol. 2014, 363, 185–199. [Google Scholar] [CrossRef]
  58. García de Madinabeitia, S.; Garate-Olave, I.; Roda-Robles, E.; Gil-Crespo, P.P.; Gil-Ibarguchi, J.I. Datación U-Pb de columbita-tantalita mediante LA-Q-ICP-MS. Ejemplo: Granito/aplopegmatitas de Tres Arroyos (Zona Centro-Ibérica). Macla 2019, 24, 23. [Google Scholar]
  59. Whitney, D.L.; Evans, B.W. Abbreviations for names of rock-forming minerals. Am. Mineral. 2010, 95, 185–187. [Google Scholar] [CrossRef]
  60. Černý, P.; Ercit, T.S. Mineralogy of niobium and tantalum: Crystal chemistry relationship, paragenetic aspects and their economic implications. In Lanthanides, Tantalum and Niobium; Moller, P., Černý, P., Saupé, F., Eds.; Spring: Berlin, Germany, 1989; pp. 27–29. [Google Scholar]
  61. Tindle, A.G.; Breaks, F.W. Oxide minerals of the Separation Rapids rare-element granitic pegmatite group, northwestern Ontario. Can. Mineral. 1998, 36, 609–635. [Google Scholar]
  62. Shand, S.J. The Eruptive Rocks, 2nd ed.; John Wiley: New York, NY, USA, 1943. [Google Scholar]
  63. Debon, F.; Lefort, P. A chemical-mineralogical classification of common plutonic rocks and associations. Trans. R. Soc. Edinb. Earth Sci. 1983, 73, 135–149. [Google Scholar] [CrossRef]
  64. Villaseca, C.; Barbero, L.; Herreros, V. A re-examination of the typology of peraluminous granite types in intracontinental orogenic belts. Trans. R. Soc. Edinb. Earth Sci. 1998, 89, 113–119. [Google Scholar] [CrossRef]
  65. Rudnick, R.L.; Gao, S. 4.1-Composition of the Continental Crust. In Treatise on Geochemistry, 2nd ed.; Holland, H.D., Turekian, K.K., Eds.; Elsevier: Oxford, UK, 2014; Volume 4, pp. 1–51. [Google Scholar]
  66. Pistorino, M.; Nestola, F.; Boffa Ballaran, T.; Domeneghetti, M.C. The effect of composition and cation ordering on the compressibility of columbites up to 7 GPa. Phys. Chem. Minerals 2006, 33, 593–600. [Google Scholar] [CrossRef]
  67. Izoret, L.; Marnier, G.; Dusausoy, Y. Crystallochemical classification of cassiterite from tin and tungsten deposits in Galicia, Spain. Can. Mineral. 1985, 23, 221–231. [Google Scholar]
  68. Raimbault, L. Composition of complex lepidolite-type granitic pegmatites and of constituent columbite-tantalite, Chèdeville, Massif Central, France. Can. Mineral. 1998, 36, 563–583. [Google Scholar]
  69. Linnen, R.; Keppler, H. Columbite solubility in granitic melts: Consequences for the enrichment and fractionation of Nb and Ta in the Earth’s crust. Contrib. Mineral. Petrol. 1997, 128, 213–227. [Google Scholar] [CrossRef]
  70. Bartels, A.; Vetere, F.; Holtz, F.; Behrens, H.; Linnen, R.L. Viscosity of flux-rich pegmatitic melts. Contrib. Mineral. Petrol. 2011, 162, 51–60. [Google Scholar] [CrossRef]
  71. Breiter, K.; Škoda, R.; Uher, P. Nb-Ta-Ti-W-Sn-oxide minerals as indicators of a peraluminous P- and F-rich granitic system evolution: Podlesi, Czech Republic. Mineral. Petrol. 2007, 91, 225–248. [Google Scholar] [CrossRef]
  72. Keppler, H. Influence of fluorine on the enrichment of high field strength trace elements in granitic rocks. Contrib. Mineral. Petrol. 1993, 114, 479–488. [Google Scholar] [CrossRef]
  73. Linnen, R.L. Depth of emplacement, fluid provenance and metallogeny in granitic terranes: A comparison of western Thailand with other tin belts. Mineral. Depos. 1998, 33, 461–476. [Google Scholar] [CrossRef]
  74. Wolf, M.B.; London, D.; Morgan, G.B.V. Effects of boron on the solubility of cassiterite and tantalite in granitic liquids. Soc. Am. Prog. Abstr. 1994, 26, A450. [Google Scholar]
  75. London, D.; Morgan Vi, G.B.; Wolf, M.B. Amblygonite-montebrasite solid solutions as monitors of fluorine in evolved granitic and pegmatitic melts. Am. Mineral. 2001, 86, 225–233. [Google Scholar] [CrossRef]
  76. Lukkari, S. Magmatic Evolution of Topaz-Bearing Granite Stocks within the Wiborg Rapakivi Granite Batholith. Ph.D. Thesis, University of Helsinki, Helsinki, Finland, 2007. [Google Scholar]
  77. Černý, P.; Novák, M.; Chapman, R. Effects of sillimanite-grade metamorphism and shearing on Nb-Ta oxide minerals in granitic pegmatites, Marsikov, northern Moravia, Czechoslovakia. Can. Mineral. 1992, 30, 699–718. [Google Scholar]
  78. Lahti, S.I. Zoning in columbite-tantalite crystals from the granitic pegmatites of the Eräjärvi area, southern Finland. Geochim. Cosmochim. Acta 1987, 51, 509–517. [Google Scholar] [CrossRef]
  79. Tindle, A.G.; Breaks, F.W. Columbite-tantalite mineral chemistry from rare-element granitic pegmatites: Separation Lake area, N.W. Ontario, Canada. Mineral. Petrol. 2000, 70, 165–198. [Google Scholar] [CrossRef]
  80. Aseri, A.A.; Linnen, R.L.; Che, X.D.; Thibault, Y.; Holtz, F. Effects of fluorine on the solubilities of Nb, Ta, Zr and Hf minerals in highly fluxed water-saturated haplogranitic melts. Ore Geol. Rev. 2015, 64, 736–746. [Google Scholar] [CrossRef] [Green Version]
  81. Fiege, A.; Kirchner, C.; Holtz, F.; Linnen, R.L.; Dziony, W. Influence of fluorine on the solubility of manganotantalite (MnTa2O6) and manganocolumbite (MnNb2O6) in granitic melts-An experimental study. Lithos 2011, 122, 165–174. [Google Scholar] [CrossRef]
  82. Raimbault, L.; Burnol, L. The Richemont rhyolite dyke, Massif Central, France: A subvolcanic equivalent of rare-metal granites. Can. Mineral. 1998, 36, 265–282. [Google Scholar]
  83. Llorens González, T.; García Polonio, F.; López Moro, F.J.; Fernández Fernández, A.; Sanz Contreras, J.L.; Moro Benito, M.C. Tin-tantalum-niobium mineralization in the Penouta deposit (NW Spain): Textural features and mineral chemistry to unravel the genesis and evolution of cassiterite and columbite group minerals in a peraluminous system. Ore Geol. Rev. 2017, 81, 79–95. [Google Scholar] [CrossRef]
  84. Wise, M.A.; Francis, C.A.; Černý, P. Compositional and structural variations in columbite-group minerals from granitic pegmatites of the Brunswick and Oxford fields, Maine: Differential trends in F-poor and F-rich environments. Can. Mineral. 2012, 50, 1515–1530. [Google Scholar] [CrossRef]
  85. Beurlen, H.; Da Silva, M.R.R.; Thomas, R.; Soares, D.R.; Olivier, P. Nb-Ta-(Ti-Sn) oxide mineral chemistry as tracer of rare-element granitic pegmatite fractionation in the Borborema Province, Northeastern Brazil. Mineral. Depos. 2008, 43, 207–228. [Google Scholar] [CrossRef] [Green Version]
  86. Pieczka, A. Primary Nb-Ta minerals in the Szklary pegmatite, Poland: New insights into controls of crystal chemistry and crystallization sequences. Am. Mineral. 2010, 95, 1478–1492. [Google Scholar] [CrossRef]
  87. Van Lichtervelde, M.; Salvi, S.; Beziat, D.; Linnen, R.L. Textural features and chemical evolution in tantalum oxides: Magmatic versus hydrothermal origins for Ta mineralization in the Tanco Lower Pegmatite, Manitoba, Canada. Econ. Geol. 2007, 102, 257–276. [Google Scholar] [CrossRef]
  88. Jolliff, B.L.; Papike, J.J.; Laul, J.C. Mineral recorders of pegmatite internal evolution; REE contents of tourmaline from the Bob Ingersoll Pegmatite, South Dakota. Geochim. Cosmochim. Acta 1987, 51, 2225–2232. [Google Scholar] [CrossRef]
  89. Selway, J.B.; Novák, M.; Černý, P.; Hawthorne, F.C. The Tanco pegmatite at Bernic Lake, Manitoba. XIII. Exocontact tourmaline. Can. Mineral. 2000, 38, 869–876. [Google Scholar] [CrossRef]
  90. Monier, G.; Charoy, B.; Cuney, M.; Ohnenstetter, D.; Robert, J.L. Évolution spatiale et temporelle de la composition des micas du granite albitique á topaze-lépidolite de Beauvoir. Géol. France 1987, 2, 179–188. [Google Scholar]
  91. Bhalla, P.; Holtz, F.; Linnen, R.L.; Behrens, H. Solubility of cassiterite in evolved granitic melts: Effect of T, fO2, and additional volatiles. Lithos 2005, 80, 387–400. [Google Scholar] [CrossRef]
  92. Xie, L.; Wang, R.-C.; Groat, L.A.; Zhu, J.-C.; Huang, F.-F.; Cempírek, J. A combined EMPA and LA-ICP-MS study of Li-bearing mica and Sn-Ti oxide minerals from the Qiguling topaz rhyolite (Qitianling District, China): The role of fluorine in origin of tin mineralization. Ore Geol. Rev. 2015, 65, 779–792. [Google Scholar] [CrossRef]
  93. Heinrich, C.A. Geochemical evolution and hydrothermal mineral deposition in Sn(–W-base metal) and other granite-related ore systems: Some conclusions from Australian examples. In Magmas, Fluids, and Ore Deposits; Thompson, J.F.H., Ed.; Mineralogical Association of Canada Short Course: Québec, QC, Canada, 1995. [Google Scholar]
  94. Candela, P.A. A review of shallow, ore-related granites: Textures, volatiles, and ore metals. J. Petrol. 1997, 38, 1619–1633. [Google Scholar] [CrossRef]
  95. Linnen, R.; Williams-Jones, A.E.; Martin, R.F. Evidence of magmatic cassiterite mineralization at the Nong Sua aplite-pegmatite complex, Thailand. Can. Mineral. 1992, 30, 739–761. [Google Scholar]
  96. London, D. Ore-forming processes within granitic pegmatites. Ore Geol. Rev. 2018, 101, 349–383. [Google Scholar] [CrossRef]
  97. Simmons, W.B.; Webber, K.L. Pegmatite genesis: State of the art. Eur. J. Mineral. 2008, 20, 421–438. [Google Scholar] [CrossRef]
  98. Roda-Robles, E.; Vieira, R.; Pesquera, A.; Lima, A. Chemical variations and significance of phosphates from the Fregeneda-Almendra pegmatite field, Central Iberian Zone (Spain and Portugal). Mineral. Petrol. 2010, 100, 23–34. [Google Scholar] [CrossRef]
  99. Deng, X.D.; Li, J.W.; Zhao, X.F.; Hu, Z.C.; Hu, H.; Selby, D.; de Souza, Z.S. U-Pb isotope and trace element analysis of columbite-(Mn) and zircon by laser ablation ICP-MS: Implications for geochronology of pegmatite and associated ore deposits. Chem. Geol. 2013, 344, 1–11. [Google Scholar] [CrossRef] [Green Version]
  100. Romer, R.L.; Wright, J.A. U-Pb dating of columbites: A geochronologic tool to date magmatism and ore deposits. Geochim. Cosmochim. Acta 1992, 56, 2137–2142. [Google Scholar] [CrossRef]
  101. Melleton, J.; Gloaguen, E. Timing of rare-elements (Li-Be-Ta-Sn-Nb) magmatism in the European Variscan belt The Variscan Belt: Correlations and plate dynamics. Geol. France 2015, 1, 100. [Google Scholar]
  102. Villaseca, C. On the origin of granite types in the Central Iberian Zone: Contribution from integrated U-Pb and Hf isotope studies of zircon. In Proceedings of the VIII Congresso Ibérico de Geoquímica, Castelo Branco, Portugal, 24 September 2011; pp. 29–34. [Google Scholar]
  103. Dingwell, D.B.; Hess, K.U.; Knoche, R. Granite and granitic pegmatite melts: Volumes and viscosities. Trans. Roy. Soc. Edinb. Earth Sci. 1996, 87, 65–72. [Google Scholar] [CrossRef]
  104. Holtz, F.; Johannes, W.; Tamic, N.; Behrens, H. Maximum and minimum water contents of granitic melts generated in the crust: A reevaluation and implications. Lithos 2001, 56, 1–14. [Google Scholar] [CrossRef]
  105. Tartèse, R.; Boulvais, P. Differentiation of peraluminous leucogranites “en route” to the surface. Lithos 2010, 114, 353–368. [Google Scholar] [CrossRef]
  106. Roda-Robles, E.; Pesquera, A.; Gil-Crespo, P.P.; Torres-Ruiz, J. From granite to highly evolved pegmatite: A case study of the Pinilla de Fermoselle granite–pegmatite system (Zamora, Spain). Lithos 2012, 153, 192–207. [Google Scholar] [CrossRef]
Figure 1. (a) Geological map of the Iberian Massif (subdivision after [50,51], and [52], and variscan granitoids adapted from [53]). (I) Cantabrian Zone, (II) West Asturian-Leonese Zone, (III) Galicia-Trás-os-Montes Zone, (IV) Central Iberian Zone, (V) Ossa-Morena Zone, (VI) South Portuguese Zone, and (VII) variscan granitoids. (b) Regional geological map of the Nisa-Alburquerque batholith, based on [53] (c) Detailed map of the Tres Arroyos pegmatitic field (modified from [46]). Locations of the selected samples for the U-Pb dating (3AR-15 and 3AR-16) are shown (see Table S1 for locations of all studied samples).
Figure 1. (a) Geological map of the Iberian Massif (subdivision after [50,51], and [52], and variscan granitoids adapted from [53]). (I) Cantabrian Zone, (II) West Asturian-Leonese Zone, (III) Galicia-Trás-os-Montes Zone, (IV) Central Iberian Zone, (V) Ossa-Morena Zone, (VI) South Portuguese Zone, and (VII) variscan granitoids. (b) Regional geological map of the Nisa-Alburquerque batholith, based on [53] (c) Detailed map of the Tres Arroyos pegmatitic field (modified from [46]). Locations of the selected samples for the U-Pb dating (3AR-15 and 3AR-16) are shown (see Table S1 for locations of all studied samples).
Minerals 10 01008 g001
Figure 2. (a) Outcrop view of a layered barren aplite-pegmatite with alternating aplitic and pegmatitic bands. (b) Subhorizontal intermediate aplite-pegmatite dykes cutting the host metasedimentary sequence. (c) Detail of an intermediate aplite-pegmatite body showing a simple layering pattern with comb-texture in feldspar. (d) Characteristic view of a discordant contact between a dyke of Li-rich aplite-pegmatite and the host rock. (e) Detail of a sample section of a Li-rich aplite-pegmatite showing the main rock-forming minerals (feldspar, quartz and Li-micas), two cassiterite crystals are highlighted. (f) Representative photograph of a complex layering in a Li-rich aplite-pegmatite sample.
Figure 2. (a) Outcrop view of a layered barren aplite-pegmatite with alternating aplitic and pegmatitic bands. (b) Subhorizontal intermediate aplite-pegmatite dykes cutting the host metasedimentary sequence. (c) Detail of an intermediate aplite-pegmatite body showing a simple layering pattern with comb-texture in feldspar. (d) Characteristic view of a discordant contact between a dyke of Li-rich aplite-pegmatite and the host rock. (e) Detail of a sample section of a Li-rich aplite-pegmatite showing the main rock-forming minerals (feldspar, quartz and Li-micas), two cassiterite crystals are highlighted. (f) Representative photograph of a complex layering in a Li-rich aplite-pegmatite sample.
Minerals 10 01008 g002
Figure 3. Backscattered electron images of Nb-Ta oxides. (a) Elongated prismatic and (b) prismatic crystals from the barren aplite-pegmatites with irregular zones. (c) and (d) crystals from intermediate aplite-pegmatites showing a concentric zoning and irregular subrounded “patches”, as well as abundant corrosion gulfs. (e) A tabular concentrically zoned crystal and (f) an elongated prismatic slightly zoned crystal from Li-rich aplite-pegmatites.
Figure 3. Backscattered electron images of Nb-Ta oxides. (a) Elongated prismatic and (b) prismatic crystals from the barren aplite-pegmatites with irregular zones. (c) and (d) crystals from intermediate aplite-pegmatites showing a concentric zoning and irregular subrounded “patches”, as well as abundant corrosion gulfs. (e) A tabular concentrically zoned crystal and (f) an elongated prismatic slightly zoned crystal from Li-rich aplite-pegmatites.
Minerals 10 01008 g003
Figure 4. Optical photomicrographs showing the mineral assemblage of the Li-rich aplite-pegmatites, with (a) a strongly pleochroic subhedral cassiterite crystal, and (b) a cassiterite crystal with a chromatic zoning alternating different shades of brown standing out. Mineral abbreviations from Whitney and Evans [59].
Figure 4. Optical photomicrographs showing the mineral assemblage of the Li-rich aplite-pegmatites, with (a) a strongly pleochroic subhedral cassiterite crystal, and (b) a cassiterite crystal with a chromatic zoning alternating different shades of brown standing out. Mineral abbreviations from Whitney and Evans [59].
Minerals 10 01008 g004
Figure 5. Quadrilateral representation of Mn/(Mn+Fe) vs. Ta/(Ta+Nb) for the minerals of the CT group from the different three aplite-pegmatite types of Tres Arroyos. The reddish arrow shows the overall chemical trend for the aplite-pegmatites from the Tres Arroyos area, whereas the grey arrow indicates the normal fractionation tendency in LCT pegmatites based on Černý [60].
Figure 5. Quadrilateral representation of Mn/(Mn+Fe) vs. Ta/(Ta+Nb) for the minerals of the CT group from the different three aplite-pegmatite types of Tres Arroyos. The reddish arrow shows the overall chemical trend for the aplite-pegmatites from the Tres Arroyos area, whereas the grey arrow indicates the normal fractionation tendency in LCT pegmatites based on Černý [60].
Minerals 10 01008 g005
Figure 6. Covariation of (a) Nb+Ta vs. TiO2 in cassiterite (values in apfu and wt%, respectively). (b) Nb+Ta vs. Fe+Mn (apfu) in cassiterite. The rare-element pegmatites field and the hydrothermal and epithermal field modified from [61]. (c) Nb+Ta vs. Sn (apfu) for the cassiterites from the aplite-pegmatites from Tres Arroyos.
Figure 6. Covariation of (a) Nb+Ta vs. TiO2 in cassiterite (values in apfu and wt%, respectively). (b) Nb+Ta vs. Fe+Mn (apfu) in cassiterite. The rare-element pegmatites field and the hydrothermal and epithermal field modified from [61]. (c) Nb+Ta vs. Sn (apfu) for the cassiterites from the aplite-pegmatites from Tres Arroyos.
Minerals 10 01008 g006
Figure 7. Concordia diagram for the CT group minerals from the Tres Arroyos aplite-pegmatites.
Figure 7. Concordia diagram for the CT group minerals from the Tres Arroyos aplite-pegmatites.
Minerals 10 01008 g007
Figure 8. (a) B-A diagram of Debon and Lefort [63], modified by Villaseca et al. [64], showing whole-rock composition of all distinguished lithotypes in Tres Arroyos. (b) Upper continental crust-normalized [65] spider diagrams for the studied samples from the Tres Arroyos area. (ch) Whole-rock variation diagrams of selected elements, Fe, Mg and Mn in millications, other elements in ppm. In all diagrams, the unfilled symbols correspond to compositions from Gallego-Garrido [46].
Figure 8. (a) B-A diagram of Debon and Lefort [63], modified by Villaseca et al. [64], showing whole-rock composition of all distinguished lithotypes in Tres Arroyos. (b) Upper continental crust-normalized [65] spider diagrams for the studied samples from the Tres Arroyos area. (ch) Whole-rock variation diagrams of selected elements, Fe, Mg and Mn in millications, other elements in ppm. In all diagrams, the unfilled symbols correspond to compositions from Gallego-Garrido [46].
Minerals 10 01008 g008
Figure 9. Binary diagrams of CT oxides compositions in apfu. (a) Mn2+ vs. Fe2+. (b) Ta vs. Nb. (c) 2(Sn+Ti)+W+vacancy vs. 2(Nb+Ta)+2(Fe+Mn).
Figure 9. Binary diagrams of CT oxides compositions in apfu. (a) Mn2+ vs. Fe2+. (b) Ta vs. Nb. (c) 2(Sn+Ti)+W+vacancy vs. 2(Nb+Ta)+2(Fe+Mn).
Minerals 10 01008 g009
Table 1. Mineralogy and petrographic characteristics of the three aplite-pegmatite types from Tres Arroyos.
Table 1. Mineralogy and petrographic characteristics of the three aplite-pegmatite types from Tres Arroyos.
LithologyMain MineralsAccessory MineralsOxides Crystal SizeOxides TextureMain Petrographic Textures
Barren aplite-pegmatitesQuartz, plagioclase,
K-feldspar, muscovite, tourmaline
‘Zinnwaldite’ mineral series, topaz, Fe-Mn-(± Mg ± Al) phosphates, fluorapatite, Sn-Nb-Ta oxidesCT: Very fine < 2 mm
Cassiterite: Very fine < 2 mm
Sub-anhedral, elongated-prismatic to acicular CT
Slight concentric zoning, irregular patches and corrosion gulfs in CT
Layered texture, alternating pegmatitic and aplitic layers; locally greenish colour in hand sample due to Fe-Mn phosphates alteration; feldspars display comb-texture
Intermediate aplite-pegmatitesQuartz, plagioclase, K-feldspar, muscoviteTopaz, Li-Al phosphates, fluorapatite, Sn-Nb-Ta oxidesCT: Very fine < 2 mm
Cassiterite: Very fine < 2 mm
Sub-anhedral
Well-develop primary faces in CT
Concentric zoning, irregular patches and corrosion gulfs in CT
Layering: alternating albite-rich and quartz-rich layers, grain size also may change; aplitic albite: wedged crystals; coarse feldspars display comb-texture
Li-rich aplite-pegmatitesQuartz, plagioclase, K-feldspar, Li-Al-micaTopaz, Li-Al phosphates, Sn-Nb-Ta oxides, fluorapatiteCT: Very fine < 3 mm
Cassiterite: Very fine < 5 mm
Subhedral elongated-prismatic to acicular CT
Anhedral-subhedral cassiterite with chromatic zoning
Concentric zoning, slight irregular patches and less corrosion gulfs in CT
Rhythmic layering, some with a complex pattern, with alternating albite-rich and Li-mica rich layers; feldspars display comb-texture
CT—Columbite-tantalite group minerals.
Table 2. Selected microprobe analyses of columbite-tantalite group mineral from the Tres Arroyos aplite-pegmatites.
Table 2. Selected microprobe analyses of columbite-tantalite group mineral from the Tres Arroyos aplite-pegmatites.
LithologyBarren Aplite-PegmatitesIntermediate Aplite-Pegmatites
MineralTnfClfClfClfClfTnmTnmTnfClmClm
Number156891213151718
wt%
TiO2 2.191.371.354.771.26-0.000.020.150.96
MnO0.230.090.160.140.199.778.357.8810.3013.92
FeO13.1718.5418.1316.4318.126.498.468.146.955.78
Nb2O523.0169.1669.1250.8168.9123.7225.1021.6033.8861.61
SnO20.860.000.040.920.020.110.090.190.160.19
Ta2O552.976.647.5320.666.4259.4053.9959.1846.7513.74
WO31.343.542.963.744.330.891.210.920.672.55
MgO0.000.030.000.000.02----0.00
Total93.7799.3799.2897.4899.26100.3897.2097.9398.8598.79
apfu
Ti0.1270.0600.0590.2240.0550.0000.0000.0010.0080.044
Mn0.0150.0040.0080.0080.0090.6080.5290.5070.6120.713
Fe0.8450.9020.8840.8570.8830.3990.5290.5170.4080.292
Nb0.7971.8191.8221.4331.8170.7880.8480.7411.0751.684
Sn0.0260.0000.0010.0230.0000.0030.0030.0060.0040.005
Ta1.1040.1050.1190.3500.1021.1871.0981.2220.8920.226
W0.0270.0530.0450.0600.0650.0170.0230.0180.0120.040
Mg0.0000.0030.0000.0000.0020.0000.0000.0000.0000.000
XTa0.5810.0550.0610.1970.0530.6010.5640.6220.4540.118
XMn0.0170.0050.0090.0090.0110.6040.5000.4950.6000.709
Ʃ Cations2.9412.9432.9382.9552.9323.0013.0303.0123.0123.004
Ʃ Charges11.99811.99711.99711.99611.99611.99911.99911.99911.99911.994
LithologyIntermediate Aplite-PegmatitesLi-Rich Aplite-Pegmatites
MineralClmClmClfClfClfClmClmClmClmClm
Number19202327303132333437
wt%
TiO2 0.360.22-0.340.790.030.02-0.020.01
MnO10.4016.317.986.387.4517.3618.6719.1218.9818.99
FeO7.861.229.6811.3711.080.730.130.160.210.21
Nb2O545.2236.7938.3735.0653.5250.8555.8656.6156.3854.51
SnO20.100.370.660.191.190.490.300.250.130.12
Ta2O534.2144.0538.5944.9822.9629.2223.5022.5922.8825.00
WO30.650.365.450.572.060.600.580.440.420.41
MgO0.00---0.00-----
Total98.8199.31100.7298.8999.0999.2899.0599.1799.0099.26
apfu
Ti 0.0180.0110.0000.0180.0370.0010.0010.0000.0010.000
Mn0.5810.9500.4550.3760.3970.9430.9911.0101.0051.013
Fe0.4340.0700.5440.6610.5820.0390.0070.0090.0110.011
Nb1.3501.1431.1661.1031.5211.4741.5821.5961.5931.551
Sn0.0030.0100.0180.0050.0300.0130.0070.0060.0030.003
Ta0.6140.8240.7050.8510.3920.5100.4000.3830.3890.428
W0.0110.0060.0950.0100.0340.0100.0090.0070.0070.007
Mg0.0000.0000.0000.0000.0000.0000.0000.0000.0000.000
XTa0.3130.4190.3770.4360.2050.2570.2020.1940.1960.216
XMn0.5730.9310.4550.3620.4050.9600.9930.9920.9890.989
Ʃ Cations3.0103.0152.9823.0252.9932.9902.9983.0113.0093.014
Ʃ Charges11.99812.00011.99411.99911.99411.99911.99912.00012.00012.000
-: not analyzed. Structural formula in the basis of 6 O atoms. Tnf: tantalite-(Fe); Clf: columbite-(Fe); Tnm: tantalite-(Mn); Clm: columbite-(Mn); XTa: Ta/(Ta+Nb); XFe: Mn/(Mn+Fe).
Table 3. Selected microprobe analyses of cassiterite from the Tres Arroyos aplite-pegmatites.
Table 3. Selected microprobe analyses of cassiterite from the Tres Arroyos aplite-pegmatites.
LithologyBarren Aplite-
Pegmatites
Intermediate
Aplite-Pegmatites
Li-Rich Aplite-Pegmatites
Number12457161718192429
wt%
TiO20.850.610.220.140.080.010.090.210.000.010.06
MnO0.010.000.000.100.170.010.600.020.010.020.09
FeO0.860.390.030.731.550.010.090.000.000.020.06
Nb2O50.720.270.270.723.450.223.000.110.040.311.13
SnO295.2197.5397.9593.6290.0198.9092.1898.8699.5099.2096.52
Ta2O53.301.610.554.435.800.203.060.610.600.060.17
WO30.140.010.000.110.020.090.060.000.050.340.98
MgO0.000.000.000.000.000.000.000.000.000.000.00
BaO0.000.000.000.000.000.000.000.000.000.000.00
Total101.07100.4299.3399.84101.0799.4499.1699.82100.2199.9599.00
apfu
Ti 0.0310.0230.0080.0050.0030.0000.0030.0080.0000.0000.002
Mn0.0000.0000.0000.0040.0070.0000.0250.0010.0000.0010.004
Fe0.0350.0160.0010.0310.0640.0000.0040.0000.0000.0010.003
Nb0.0160.0060.0060.0160.0770.0050.0680.0020.0010.0070.026
Sn1.8731.9341.9621.8791.7681.9881.8401.9781.9871.9821.940
Ta0.0440.0220.0080.0610.0780.0030.0420.0080.0080.0010.002
W0.0020.0000.0000.0010.0000.0010.0010.0000.0010.0040.013
Mg0.0000.0000.0000.0000.0000.0000.0000.0000.0000.0000.000
Ba0.0000.0000.0000.0000.0000.0000.0000.0000.0000.0000.000
XTa0.7340.7840.5490.7870.5030.3560.3800.7770.9030.0950.083
XMn0.0080.0000.0000.1220.0990.4530.8680.7950.7910.6030.577
Ʃ Cations2.0022.0011.9851.9981.9971.9981.9831.9981.9971.9971.990
Ʃ Charges8.0008.0007.9528.0008.0008.0007.9868.0007.9988.0007.999
Structural formula in the basis of 4 O atoms. XTa: Ta/(Ta+Nb); XMn:Mn/(Mn+Fe).
Table 4. U-Pb isotope data (LA-Q-ICP-MS) of columbite-tantalite group minerals from the Tres Arroyos aplite-pegmatites.
Table 4. U-Pb isotope data (LA-Q-ICP-MS) of columbite-tantalite group minerals from the Tres Arroyos aplite-pegmatites.
Spot NameData for Wetherill PlotData for Tera-Wasserburg PlotElemental Concentration (µg g−1) a
207Pb/235U2s (abs)206Pb/238U2s (abs)Rho238U/206Pb2s (abs)207Pb/206Pb2s (abs)RhoUThPb
CT-010.3610.0120.0460.0010.7421.90.5270.0560.0020.272540.721.83
CT-020.4140.0170.0510.0020.7219.80.6640.0580.0020.263081.559.6
CT-030.3810.0120.0480.0010.7120.80.4780.0570.0020.362541.064.09
CT-040.4140.0150.0520.0020.8419.10.5480.0560.0010.424883.1412.6
CT-050.3670.0150.0480.0010.6121.10.6200.0550.0020.244932.157.51
CT-060.3850.0130.0490.0010.7920.60.5500.0560.0020.233591.044.20
CT-070.3600.0120.0470.0010.7721.40.5480.0550.0010.246262.316.50
CT-080.3890.0150.0490.0010.7020.50.5880.0570.0020.253481.393.50
CT-090.4940.0260.0530.0020.7219.00.7980.0680.0030.061231.289.6
CT-100.4480.0190.0530.0020.2018.80.5280.0610.0030.232640.4530.0
CT-111.650.0690.0610.0020.7816.40.4300.1940.005−0.361380.20139
CT-125.790.1500.1120.0030.868.970.2330.3760.0050.177832.043200
CT-138.950.3800.1410.0050.917.110.2630.4530.006−0.2626558.8916,700
CT-146.740.2000.1170.0030.878.520.2180.4140.005−0.02220714.010,350
CT-151.550.0940.0600.0020.3316.60.6350.1860.0110.331030.3193.6
CT-162.900.3000.0740.0040.1613.60.6460.2880.0310.101240.20199
CT-175.260.1900.1030.0030.829.680.2720.3700.006−0.144332.711737
a concentration uncertainty ca. 20%.
Table 5. Whole-rock major- (wt%) and trace-element (ppm) compositions of representative samples from the Tres Arroyos granite-aplite-pegmatite field.
Table 5. Whole-rock major- (wt%) and trace-element (ppm) compositions of representative samples from the Tres Arroyos granite-aplite-pegmatite field.
Sample3AR-13AR-23AR-33AR-43AR-53AR-63AR-73AR-8 **3AR-93AR-103AR-113AR-12 **3AR-13 **3AR-14 **
Lithotype11223444444555
SiO272.8572.3772.6374.0072.1370.8168.8673.7472.4072.0170.6468.8867.1771.13
Al2O314.1714.7514.8614.8616.2316.5516.5016.1317.0416.9716.2917.5418.5016.64
Fe2O3t2.172.300.961.380.660.630.300.720.450.830.410.700.060.23
MnO0.030.030.060.020.070.030.020.020.020.020.030.070.090.08
MgO0.470.360.040.090.040.050.050.030.040.020.030.010.020.02
CaO0.840.890.220.390.210.930.640.250.280.230.390.170.310.35
Na2O3.303.553.533.676.366.355.247.547.297.475.855.504.044.27
K2O4.935.134.944.161.381.413.340.751.131.062.791.904.002.80
TiO20.310.300.030.070.010.010.010.000.010.000.010.010.010.01
P2O50.480.331.040.601.170.992.080.460.490.461.792.500.800.57
F0.080.070.130.100.480.240.200.040.080.080.241.251.801.44
LOI1.290.631.400.841.151.702.020.900.971.240.922.113.882.60
Total100.92100.7199.84100.1899.8999.7099.26100.58100.20100.3999.39100.64100.68100.14
Ag0.8<0.5<0.5<0.5<0.5<0.5<0.5<0.5<0.5<0.5<0.5<0.5<0.5<0.5
As17198610290<5<5<5<5<5<5<5<5<5
Ba21641210685274320.522.525.62210167
Be5623475686710711
Bi0.80.41.41.50.61.00.80.15<0.1<0.10.2<0.1<0.10.1
Co34264<124<126<1<1<1<1<13332
Cr<2050<20120<2090<20<2040<2090<20<20<20
Cs2219.64829.226.115.423.88.4316.811.861.7136260341
Cu<10<10<10<10<10<10<10<10<10<10<10<10<10<10
Ga242435323652.84541.546.542.245506550
Ge2.02.33.73.48.55.2654.334.54.87.47.410.09.0
Hf3.94.11.11.01.54.22.24.734.43.041.12.73.22.2
In0.20.20.40.60.30.341.10.150.250.220.61.23.61.7
Li170210200150190848108575102290526082305730
Mo<2<2<2<2<23<2<2<2<25<2<2<2
Nb11.012.510.515.953.962.348.093.911951.240.280.793.171.5
Ni<20<20<20<20<20<20<20<20<20<20<20<20<20<20
Pb2022811511142124201268<5
Rb35231488156845931811302484213421130199047103450
Sb0.4<0.20.5<0.20.70.40.50.40.40.44<0.21.10.40.6
Sc5446<1<1<1<1<1<1<1<1<1<1
Sn1620383775983113556601503701150530
Sr4561640582.411366.38983.2285169171156
Ta2.191.813.944.8622.986.451.611115180.654.99171.664.8
Th12.212.00.620.670.491.890.470.770.730.710.580.831.210.95
Tl1.691.864.113.312.241.46.490.921.801.2410.18.0620.515.3
U5.337.8513.86.2217.624.37.593.987.8314.812.929.011.77.72
V2218<5<5<5<5<5<5<5<5<5<5<5<5
W265344911.923812.51236.087.84.546.234.6249301
Y17.214.83.62.30.71.680.80.90.50.65<0.5<0.5<0.5<0.5
Zn708080607030<30<30<30<30<3014080150
Zr144169222619501939322813283624
La20.226.10.891.260.340.730.240.3130.2750.250.40.310.230.19
Ce44.155.22.012.360.430.9180.340.4030.3150.2780.70.360.260.24
Pr5.536.310.240.260.050.0840.050.0630.0450.040.080.040.030.02
Nd22.824.10.870.850.140.2740.30.2950.1950.1840.230.170.10.1
Sm4.655.080.340.240.060.0840.090.0930.0450.0720.040.070.020.03
Eu0.490.7150.0140.036<0.0050.1870.050.0270.020.0190.0050.0110.0150.01
Gd3.83.90.440.290.050.090.120.0980.0650.0920.040.060.040.03
Tb0.560.640.10.060.010.020.020.0170.010.014<0.01<0.01<0.01<0.01
Dy2.983.240.550.410.080.1080.130.0780.0650.0780.040.050.030.04
Ho0.50.480.090.070.020.0240.020.020.010.014<0.01<0.01<0.01<0.01
Er1.361.150.250.180.050.0820.060.0430.0250.0440.030.030.020.02
Tm0.1970.160.0490.0380.0120.0200.0110.0110.0060.008<0.0050.007<0.005<0.005
Yb1.240.960.330.250.090.1640.060.050.030.052<0.010.060.030.01
Lu0.1720.140.0390.0340.0130.0260.0070.0070.0040.007<0.0020.0080.004<0.002
A/CNK1.151.131.291.321.321.211.231.181.241.231.221.541.601.56
K/Rb11613646.660.825.036.824.525.122.325.720.57.937.056.74
Nb/Ta5.026.912.673.272.350.7220.9300.8440.7870.6350.7320.8871.301.10
Zr/Hf36.941.220.026.012.711.98.648.107.279.1511.810.411.310.9
(La/Lu)N12.219.42.373.852.722.873.564.818.163.82ND4.025.97ND
Eu/Eu *0.3550.4900.1100.416ND6.561.470.8511.130.6970.3810.5171.621.02
ΣREE1081286.216.341.352.811.501.511.111.151.571.180.7790.690
** Data from Roda-Robles et al. [20]; ND: not determined; lithotype: 1; NA monzogranite, 2; marginal granitic facies, 3; barren aplite-pegmatites, 4; intermediate aplite-pegmatites, 5; Li-rich aplite-pegmatites; Eu/Eu* = Eu anomaly (EuN/[(SmN*GdN)1/2].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Garate-Olave, I.; Roda-Robles, E.; Gil-Crespo, P.P.; Pesquera, A.; Errandonea-Martin, J. The Tres Arroyos Granitic Aplite-Pegmatite Field (Central Iberian Zone, Spain): Petrogenetic Constraints from Evolution of Nb-Ta-Sn Oxides, Whole-Rock Geochemistry and U-Pb Geochronology. Minerals 2020, 10, 1008. https://doi.org/10.3390/min10111008

AMA Style

Garate-Olave I, Roda-Robles E, Gil-Crespo PP, Pesquera A, Errandonea-Martin J. The Tres Arroyos Granitic Aplite-Pegmatite Field (Central Iberian Zone, Spain): Petrogenetic Constraints from Evolution of Nb-Ta-Sn Oxides, Whole-Rock Geochemistry and U-Pb Geochronology. Minerals. 2020; 10(11):1008. https://doi.org/10.3390/min10111008

Chicago/Turabian Style

Garate-Olave, Idoia, Encarnación Roda-Robles, Pedro Pablo Gil-Crespo, Alfonso Pesquera, and Jon Errandonea-Martin. 2020. "The Tres Arroyos Granitic Aplite-Pegmatite Field (Central Iberian Zone, Spain): Petrogenetic Constraints from Evolution of Nb-Ta-Sn Oxides, Whole-Rock Geochemistry and U-Pb Geochronology" Minerals 10, no. 11: 1008. https://doi.org/10.3390/min10111008

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop