Next Article in Journal
Experimental Investigation on the Energy Properties and Failure Process of Thermal Shock Treated Sandstone Subjected to Coupled Dynamic and Static Loads
Next Article in Special Issue
Efflorescent Sulphates with M+ and M2+ Cations from Fumarole and Active Geothermal Fields of Mutnovsky Volcano (Kamchatka, Russia)
Previous Article in Journal
Downdip Development of the Ni-Cu-PGE-Bearing Mafic to Ultramafic Uitkomst Complex, Mpumalanga Province, South Africa
Previous Article in Special Issue
Kaolinite-to-Chlorite Conversion from Si,Al-Rich Fluid-Origin Veins/Fe-Rich Carboniferous Shale Interaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mineralogy, Fluid Inclusions, and Oxygen Isotope Geochemistry Signature of Wolframite to Scheelite and Fe,Mn Chlorite Veins from the W, (Cu,Mo) Ore Deposit of Borralha, Portugal

by
Iuliu Bobos
1,*,
Carlos Marques de Sá
2 and
Fernando Noronha
1
1
ICTerra-Porto, Faculdade de Ciências, Universidade do Porto, 4168-007 Porto, Portugal
2
Grupo de Pesquisa em Recursos Minerais, Departamento de Geologia, Universidade Federal de Sergipe, Aracaju 49060-108, SE, Brazil
*
Author to whom correspondence should be addressed.
Minerals 2022, 12(1), 24; https://doi.org/10.3390/min12010024
Submission received: 18 September 2021 / Revised: 16 December 2021 / Accepted: 21 December 2021 / Published: 23 December 2021
(This article belongs to the Special Issue Advances in Low-Temperature Mineralogy and Geochemistry)

Abstract

:
Scheelitization of Mn-bearing wolframite, scheelite, quartz, and Fe,Mn-chlorite veins was identified in the W, (Cu,Mo) ore deposits of Borralha, by optical microscopy, electron-microprobe analysis, and stable isotope geochemistry. Fluid inclusions derived scheelite crystallization temperature was compared with the oxygen isotope temperature estimated. Scheelite was formed mainly during stage I from a low salinity aqueous-carbonic fluid dominated by CO2, where the homogenization temperature (Th) decreased from 380 °C to 200 °C (average of 284 °C). As temperature decreased further, the aqueous-carbonic fluid became dominated by CH4 (Stage II; (average Th = 262 °C)). The final stage III corresponds to lower temperature mineralizing aqueous fluid (average Th = 218 °C). In addition, salinity gradually decreased from 4.8 wt.% to 1.12 wt.%. The δ18OFluid values calculated for quartz-water and wolframite-water fractionation fall within the calculated magmatic water range. The ∆quartz-scheelite fractionation occurred at about 350–400 °C. The ∆chlorite-water fractionation factor calculated is about +0.05‰ for 330 °C, dropping to −0.68‰ and −1.26‰ at 380 °C and 450 °C, respectively. Estimated crystallizing temperatures based on semi-empirical chlorite geothermometers range from 373 °C to 458 °C and 435 °C to 519 °C. A narrower temperature range of 375 °C to 410 °C was estimated for Fe,Mn-chlorite crystallization.

1. Introduction

Tungsten mineralization in magmatic-hydrothermal deposits occurs predominantly as wolframite ([Fe,Mn]WO4) and scheelite (CaWO4). Wolframite is the main dominated tungsten mineral in granite-related systems, whereas in skarn deposits, it is scheelite [1,2,3,4]. Given the variability of the geological context in which scheelite occurs in nature- scheelite-quartz vein mineralizations and greisen in granite [5,6,7,8], quartz-scheelite veins in metabasite [9], scheelite in orogenic gold [10], and skarn deposits [11], published scheelite fluid inclusions (FIs) are somewhat limited.
Scheelite related to quartz veins and greisen (e.g., Yangjiashan, China) include vapor- and liquid-rich Fis types with detectable CH4, CO2, and N2, where the Th ranges (mean values) from 230 °C to 250 °C and salinity (mean values) from 9.3 wt.% to 6.6 wt.% NaCl equiv. [12]. In addition, ore-forming fluids show intermediate to low temperatures that belong to the H2O-NaCl ± CO2 system in the disseminated scheelite associated with potassic feldspathization and silicification of porphyritic biotite granite from the Wuyi metallogenic belt (China) [13].
In the quartz-scheelite veins hosted in metabasites (e.g., Metaggitsi area, Greece), the most abundant FIs consist of H2O-CO2 with highly variable CO2 contents (20 ± 90 vol.%) and salinity between 0.2 wt.% and 8.3 wt.% NaCl equiv., where the homogenization temperature of aqueous and gas-rich type inclusions has been estimated to about 220 ± 25 °C (mean values) [9]. Furthermore, scheelite has a distinctive trace elements composition from various deposit types and geological environments, being considered an indicator mineral for orogenic gold [14]. Microthermometric and Raman spectroscopic studies of FIs have documented low salinity aqueous-gaseous (H2O-CO2-CH4-NaCl) ore fluids, where gold and scheelite precipitated in orogenic gold deposits from the Dharwar craton (South India) [15].
The FIs of scheelite crystallized during the cooler period of skarn development (e.g., Kara, Tasmania) consist of primary H2O-CO2 with moderate salinity (12.0 wt.% and 17.8 wt.% NaCl equiv.) and homogenization temperatures of 350–580 °C [16]. A similar thermal history and scheelite formation were observed for other W skarn deposits (e.g., CanTung Mine, Northwest Territories, Canada) [17,18]. Recently, scheelite-bearing albitite (no FI data are known) was identified in the Zhuxi (China) skarn tungstate deposit as a product of a silicate-poor, H2O-rich melt that formed by melt-melt-liquid immiscibility in an extremely fractionated residual magma [19].
The scheelite occurrences mentioned above showed a connection with a magmatic-hydrothermal, metamorphic, or hydrothermal-meteoric fluid of low to moderate salinity, where homogenization temperatures ranged from 220 °C to 580 °C. Concerning the source and evolution of ore-forming fluids, the mechanisms and thermodynamic conditions of scheelite deposition have yet to be addressed [20,21]. In addition, in most related cases, scheelite was found associated with quartz veins, greisen, albitite, or secondary K-feldspar. Nevertheless, there are no published data available in the literature regarding the association of scheelite with chlorite.
Scheelite from the W, (Cu,Mo) ore deposit of Borralha is associated with wolframite, Fe,Mn-chlorite, and quartz, where Mn-bearing wolframite has been replaced by scheelite. The relationships “chlorite vs. metal assemblages” have been revisited and explored, aimed towards using chlorite as a “proxy” for detecting porphyry ore deposits [22] because the large variety of chlorite minerals can serve as a tool to determine the direction towards a mineralized area, to estimate the temperatures under which the ore processes took place [23,24], or to estimate the heat flux away from a magmatic-hydrothermal centre [25].
The stability of the Fe2+ end-member wolframite (ferberite) is much more sensitive to the high variation of fO2 than the Mn2+ end-member wolframite (hübnerite) and the composition of ferberite might also be strongly affected by the oxidation state of the system [26]. The physicochemical parameters (e.g., temperature, Mn2+/Fe2+ ratio in the mineralizing solution) during wolframite crystallization influence the hübnerite vs. ferberite proportions of wolframite [27].
Furthermore, redox state exchange between Fe- and Mn-distributions in chlorite associated with wolframite and scheelite from the Borralha deposit showed that its composition is directly influenced by thermal gradients of the evolution of different stages within the magmatic-hydrothermal system [28], where the diversity of chlorite minerals associated with W-mineralization serves as an excellent indicator to identify mineralization areas and to estimate crystallization temperatures. Geothermometers have been developed for the Al-, Fe-, or Mg-rich chlorites and little is known about the specific chlorite systems, like Mn-rich chlorite, where a new question arises on Mn-redox, analogously to Fe-redox processes [29].
The application of empirical geothermometers to ore deposit research depends on complementary data regarding the mineralogical reaction mechanisms, estimation of P-V-T-X parameters of the ore-forming fluids or ones circulated through rocks, and the stable isotope geochemistry of neoformed minerals, helping to understand mineral history. Understanding the principles governing isotopic variations in nature is still limited due to the magnitude of isotopic fractionation that accompanies various geological processes and the diversity of fluids circulation mechanisms in the oceanic or continental crust, where geochemical parameters and the kinetic isotope effects accompanying a conversion of one mineral to another under a variety of conditions.
The main aim of this study is to characterize the mineralogy of scheelite + Fe,Mn-chlorite ± wolframite scheelitized assemblages found in large crystalline Fe,Mn-chlorite masses in a W veins system, the chemistry of the fluids responsible for scheelite crystallization, where FIs data are coupled with Fe,Mn-chlorite geothermometry and the oxygen isotope geochemistry of wolframite, scheelite, quartz, and Fe,Mn-chlorite.

2. Geological Background

The Variscan orogeny is currently explained by an obduction-collision orogenic model [30,31,32], which distinguishes three main deformation phases (D1, D2, and D3) in northern Portugal [33,34]. The maximum crustal thickening was achieved by the end of the D1/D2 phases, whereas the D3 phase (intra-Westphalian age) corresponds to the final stages of the continental collision process. Most of the granitic intrusions and associated thermal metamorphic peaks are coeval with the D3 phase [30,35,36]. The syn- to late-orogenic magmatic activity generated voluminous granitoid batholiths often with contrasting compositions, where their emplacement was in most cases controlled by major Variscan structures.
The W, (Cu,Mo) deposit of Borralha is located in the Central Iberian Zone in northwestern part of the Iberian Peninsula of the (Figure 1a), where different syn-orogenic Variscan granites (e.g., two mica peraluminous, syn-D3, and biotite rich syn-D3) intruded Palaeozoic metasedimentary units. The ore deposits occur at the contact between metasedimentary formations (Silurian) and the syn-D3 porphyritic Borralha biotite granite (~315 Ma) and a syn-D3 two-mica granite (Figure 1b). The Borralha ore deposit was an important W deposit in Portugal explored for wolframite, scheelite, and argentiferous chalcopyrite during 1903–1985. W mineralization occurs in quartz veins and in the two breccia pipes known as Santa Helena and Venise [37]. The Santa Helena breccia (SHB) occurs in outcrops (Figure 1b), whereas the Venise breccia was found underground during mining works. Rock fragments of variable size, cemented with mineralized coarse-grained quartz aggregates, characterize both breccias. The underground works exploited vertical quartz veins (more than 45° dip, VV) and sub-horizontal veins or “flat veins” (<30° dip, FV). The main vertical veins mineralized with tungstates occur in the most central deposit part, located in the Santa Helena veins (e.g., Santo António, Santa Helena north, and Santa Helena south) and cross several lithologies including porphyroid biotite granite, tonalites, and schists. The dominant micaschist rocks to the east part are cross-cut by the main W veins subdivided into vertical veins sub-concordant with the schistosity, namely: The three north (3N) and three south (3S) veins. However, these lodes contain less mineralizations of tungstates and more sulphides (i.e., chalcopyrite). The ore mineral assemblages consist of tungstate (wolframite and scheelite), various sulphide minerals (chalcopyrite, pyrite, pyrrhotite, molybdenite, sphalerite, galena, bismuthinite, marcasite), native bismuth, and Pb-Bi-Ag sulphosalts. Fluid inclusion studies of quartz related to W mineralization have identified aqueous-carbonic fluids (300 °C < T < 400 °C and 50 MPa < P < 100 MPa) and aqueous fluids (250 °C < T < 300 °C and 30 MPa < P < 50 MPa) in sulphide mineralization [38,39].

3. Materials and Methods

Sample collection. Wolframite to scheelite alteration and scheelite samples (SHS and 3N “vertical veins”) with crystal sizes of a few centimeters associated with Fe,Mn-chlorite in large crystalline masses collected from the Santa Helena veins (SHS) and northern gallery (3N) from the exploitation level (−60 m underground) of the Borralha ore deposit were selected for mineralogy, FI, and oxygen isotope geochemistry analyses.
Optical microscopy. The petrography of gangue and ore minerals was conducted by optical microscopy (transmitted and reflected light) using a Nikon microscope ECLIPSE model E400POL (Minato, Tokyo, Japan), coupled with an AxioCam camera ZEISS model MRc, connected to a computer with an AxioVision imaging software (SE64 Rel.4.9).
Electron microprobe analysis. Mineral compositions were obtained with an electron probe micro-analyzer (EPMA) Jeol Hyperprobe JXA-8500F (Tokyo, Japan) operated at a 15 kV accelerating voltage and with a 10 nA beam current in the case of silicates; the operation mode for the tungstates involves 20 kV and 20 nA, with counting times of 10 s on the element peaks and 5 s on the background position. The major elements of the minerals were determined using detection limits that were acceptably (σ) above the mean background of 0.02 wt.% for most sulphides and silicates with mean counting times of 80 s. The standards used for the silicates were: Albite (NaKα), orthoclase (Al, Si, K), apatite (Ca, P), MgO (Mg), MnTiO3 (Mn), TiO2 (Ti), and FeO (Fe); for W minerals: Cassiterite (Sn ), pure Ta (), Mo (), Ni (), Co (), Cu (), Ag (), and Au () metals, SrBaNbO4O12 (Nb ), scheelite (W), Sc2O3 (Sc ), UO2 (U ), ThO2 (Th Mβ), galena (Pb ), LaP5O14 (La ), CeP5O14 (Ce ), YAg (Y ), Cd (), Sb2S3 (Sb), AsGa (As ), and Bi2Se3 (Bi). The calculation of the structural formula was based on four oxygens for tungstate minerals and 14 oxygens for chlorite.
Fluid inclusion microthermometry. Four bi-polished plate thin sections (two from SHS and the other two from 3N “vertical veins”) were prepared and studied. Fluid petrography was performed on four scheelite (SHS1, SHS2, 3N1, and 3N2) bi-polished slide samples, where about 50 FIs were observed per bi-polished sample plate. The petrographic study was carried out by transmission microscopy, using a Leica DM/LSP microscope (Wetzlar, Germany) equipped with objectives up to 50× and a computer-controlled ICC50HD imaging system using the LAS EZ4 W software. The classification of FIs was performed according to the genetic criteria of Roedder [41], where the petrographic concepts of generations of FIs and fluid phase petrography were exposed by Van den Kerkhof and Hein [42], using Shepperd’s et al. [43] abacuses. Microthermometric analyses were carried out using an Olympus microscope (Tokyo, Japan) with a maximum objective lens of 80× and a 12×, equipped with a Chaixmeca stage [44] for cryometry, and a Linkam TH600 stage, coupled to a NIKON POL-Optiphot microscope, for heating. The calibration of the equipment was done regularly using a set of SynFlinc standards. The estimated accuracy was −0.1 °C for the Chaixmeca stage and 1 °C in the Linkam stage. The salinity of the fluids was calculated with the help of the software developed by Bakker [45].
Raman spectroscopy. The study was conducted with a LabRam Dilor-JobinYvon-Spex (Lille, France) equipped with a He-Ne 632.8 nm laser source, coupled to an Olympus microscope with a 100× objective. The spectral decomposition of the bands and Raman parameters was carried out with Dilor LabSpec-JobinYvon-Spex Raman software (version 1.1). Raman spectrometry analysis of aqueous-carbonic inclusions allowed the estimation of the bulk composition and density of the fluids using the methods and equations of Touret [46] and Ramboz et al. [47]. The calibration and other technical details are described in Prieto et al. [48].
Oxygen isotopes. The scheelite, quartz, and chlorite samples were selected for oxygen isotope analysis carried out at the Laboratory for Stable Isotopes at the University of Salamanca (Salamanca, Spain). A conventional procedure was used for the extraction of oxygen from scheelite, chlorite, and quartz with BrF5 and the quantitative conversion to CO2 prior to the analysis by stable-gas mass spectrometry [49]. The δ18O values were reported in ‰ relative to the Vienna-Standard Mean Ocean Water (V-SMOW) and normalized to the NBS-30 reference (+5.17 per mil vs. V-SMOW) and NBS-28 (+9.34 per mil vs. V-SMOW). The accuracy of δ18O measurements was ±0.2‰ (1σ). The following isotope fractionation equations were used in this work between minerals and water in the system: Wolframite-H2O [50], scheelite-H2O [51,52], chlorite-H2O [53], and quartz-H2O [54].

4. Results

4.1. Mineralogy

The selected samples (3N and SHS) are characterized by the following mineral assemblages: (a) 3N: Wolframite, scheelite, Fe,Mn-chlorite, quartz, and (b) SHS: Fe,Mn-chlorite, scheelite. Fe,Mn-chlorite is represented by the following samples used for EPMA: FMN1, FMN2, FMN3, FMN8, FMN9, and FMN12.
Wolframite scheelitization. The wolframite crystals have been partially or totally replaced by scheelite in the samples analyzed. A net contact may be observed between wolframite (white field) replaced by scheelite (grey) (Figure 2a, NII). Small scheelite crystals occur along fissures or fractures and cleavages in wolframite, where the replacement of wolframite by scheelite developed a shredded texture (Figure 2b, NII) until complete replacement. Wolframite scheelitization implies a later circulation of a rich alkaline fluid (Ca2+) (Figure 2c, NII).
In addition, small inclusions of sulphide (pyrite or chalcopyrite) were identified along fractures in wolframite (Figure 2d, N+). EPMA data of wolframite show a mixed composition between ferberite (F) and hübnerite (H), where the Fe amount of 0.53 atoms per formula unit (apfu) is higher than Mn (0.43 apfu) (Table 1).
The hübnerite and ferberite ratio (H/F) calculated shows values from 46.25% to 46.7%. Small amounts of Nb2O5 (0.18%), Ta2O5 (0.01%), and SnO2 (0.01%) were also measured. EPMA data of scheelite show traces of Fe, Mn and also Nb, Ta, Mo, and Sn (Table 1). X-ray element maps of wolframite replaced by scheelite show the Fe, Mn, Ca, and W distributions in wolframite and scheelite crystals (Figure 3).
Scheelite. Scheelite (CaWO4) aggregates adjacent to wolframite or Fe,Mn-chlorite were identified. The size of the scheelite aggregates reaches from 1 to 1.5 × 0.5 to 0.75 cm2. The color is orange to yellow-brown and fluoresces with a blue-white color, indicating low MoO3 content (average <0.09%). Traces of Sn, Pb, Bi, and Mo were measured in scheelite by EPMA (Table 1).
The average chemical composition of the scheelite associated with Fe,Mn-chlorite corresponds to W1.00Ca0.99O4. The X-ray element maps show the W and Ca distributions in scheelite (Figure 4). Scheelite has a structure with tetrahedral [WO4]2− and irregular dodecahedral [CaO8]14− groups, where the hexavalent Mo can substitute for W, forming a complete solid solution series extending up to powellite (CaMoO4) [55].
Fe,Mn-chlorite. Fe,Mn-chlorite aggregates observed in transmitted light under the microscope can be well identified by their green color pleochroism, varying from pale green for some massive aggregates to a dark pleochroism, a basal cleavage, and anomalous birefringence in cross-polarized light.
A strong pleochroism in green colors with an atypical dark orange interference color was observed (Figure 5), which is in accordance with the octahedral occupancy of heavy atoms (e.g., Mn2+) [56] and the Fe + Mn/Fe + Mn + Mg ratio [57]. The elongation sign is opposite to the optical one and the optical plane is almost parallel to (010). The Fe,Mn-chlorite aggregates are radial with a radius size ranging from 500 μm to 1000 μm.
The crystal chemistry of Fe,Mn-chlorite (Table 2) studied by EPMA shows an average crystal chemical composition (140 points analysed) corresponding to: (Al1.385Fe3.76Mn0.59Mg0.24Ca0.01)5.99(Si2.625Al1.375)4O10(OH)8. There is no tetrahedral substitution [Si⇔Al(IV)], confirmed also by the calculated vacancy site (0.005) [58,59]. The R2+ exchanges (Mn2+ for Fe2+) are evidenced in the octahedral sheet, also confirmed by the vacancy site calculated after Inoue et al. [59]. Small amounts of Zn2+, Ca2+, Na+, and K+ were identified in Fe,Mn-chlorite. The ratio of Fe/(Fe + Mg) shows values from 0.92 to 0.94, whereas the Mg/(Mg + Fe) ratio is about 0.06. We also calculated the vacancy sites (AlVI-AlIV-Na-K)/2 [58], (AlVI-AlIV)/2 [59], and 6-/Fe2+ + Fe3+-Mg + AlVI [60] (see Table 2).
Compositionally, the Fe,Mn-chlorite falls in the middle field between daphnite to amesite (Figure 6) in the diagram of R2+ vs. Si4+ [61].
The oxidation state of Fe,Mn-chlorite has previously been estimated [62], where the Fe3+ exhibited values between 0.00 and 0.08, with an average value of 0.02 for 145 spots analyzed.

4.2. Fluid Phase Petrography and Microthermometry

Petrography. The scheelite crystals exhibited fractures, making difficult the distinction between primary and secondary and pseudo-secondary inclusions.
The FIs occur in three preferred modes, namely: (i) In trails; (ii) isolated inclusions, and (iii) zonate ones (Figure 7; Table 3). Isolated inclusions are generally larger than 20 µm, whereas trails and zoning plane inclusions are smaller (average ~10 μm). The biggest measured FIs has a size about 160 μm with a negative crystal or oval shape, making the crystallographic form of the tetragonal crystal easy to identify. The FIs appear mostly dark due to internal reflections and occur in random clusters (primary inclusions) or in intergranular arrangements (pseudo-secondary inclusions).
Generally, the distinguished FIs are bi-phasic. Isolated tri-phasic primary FIs with a size exceeding 100 μm (sample SHS1), primary isolated FI, and intra-granular trails of pseudo-secondary FIs are shown in Figure 8a. Isolated FIs with a negative-crystal shape (sample SHS1), primary ones with a negative-crystal shape, and pseudo-secondary oval-shaped FIs in intra-granular trails are shown in Figure 8b. Large variations in the degree of fill (Φw) were also observed, especially in the 3N1-3N2 primary FIs and in both primary and pseudosecondary FIs of SHS1 and SHS2 (Table 3).
Freezing temperature, microthermometry, and salinity. The freezing temperature data include the first melting point, freezing depression, and melting of clathrate. Three different types of fluids in the FIs hosted by scheelite were identified: (i) Aqueous-carbonic fluids with CO2 as the dominant carbonic phase, sometimes containing CH4 and N2; (ii) aqueous-carbonic fluids with CH4 as the dominant carbonic phase, containing small amounts of CO2; and (iii) aqueous fluids generally with a low to medium salinity. Aqueous-carbonic FIs with a dominant CH4 phases (sample 3N1) are shown in Figure 9a,b; those with dominantly CO2 phases (samples 3N2.1c, SHS1.3a, and SHS2.5b) are shown in Figure 9c,d.
Microthermometric data of aqueous-carbonic fluids dominated by CO2 are shown in Table 4. The Tmice of aqueous-carbonic FIs consisting mainly of CO2 ranges from −7.2 °C to 0 °C; Tclat from 0.2 °C to 11.5 °C; Th from 380 °C to 200 °C; salinity from 3.5 wt.% to 5.5 wt.% NaCl equiv.; and 80 to 100 mol% XCO2. N2 was only identified in six of the CO2-dominant aqueous-carbonic inclusions and always with XN2 < 8 mol%. The Tmice of aqueous-carbonic FIs with dominant CH4 ranges from −6.3 °C to 0 °C; Tclat from 7 °C to 10.4 °C; Th from 323 °C to 220 °C; salinity from 4.3 to 5.1 wt.% NaCl equiv.; and 100 mol% CH4 (Table 4). The data for aqueous FIs are: TmIce from −5.7 °C to 0 °C; Th from 265 °C to 130 °C; and salinity from 0 to 6.2 wt.% NaCl equiv. (Table 4).
Salinity (wt.% NaCl equiv.) measured for each type of fluid was plotted against Th (°C) where three distinct evolution stages were distinguished (Figure 10). The first stage corresponds to the evolution of the aqueous-carbonic fluid with CO2 dominating at high temperatures (<400 °C). The separation of carbonic phases shown in the Th histogram occurred at about 330 °C through boiling.
At this stage, the fluids lose their volatile components and start to precipitate their metals at the boiling point. Evidence for the boiling is also shown by the variations of liquid/vapor ratios observed in the FI and homogenization in liquid and vapor phase. The CH4 and water were separated during decompression at temperatures below 300 °C and then mixed with an aqueous fluid of lower salinity. The Th, Tmice, and salinity (wt.% NaCl equiv.) histograms corresponding to the whole set of FIs are shown in Figure 11.
The Tmi and salinity histograms show that the hotter fluids were saltier (with about 5 wt.% NaCl equiv.) than the aqueous colder fluids which mixed at the end of the process. In addition, averages of the bulk compositions and densities of the FIs studied were calculated (see Table 4). No distinction between the primary and pseudosecondary inclusions could be established using the microthermometry of the fluid inclusion assemblages, but rather on the composition of the fluids. Raman spectrometry analyses were performed on representative aqueous-carbonic type (CO2 or CH4 dominant) FIs, where differences in the CO2/CH4 ratio were identified (Figure 12 and Table S1 in the Supplementary Materials).

4.3. Oxygen Isotope Analysis

Oxygen isotope measurements were obtained for quartz, wolframite, scheelite, and Fe,Mn-chlorite (Table 5), and previously published data on Fe,Mn chlorite [28] were re-analysed. The δ18O isotope (V-SMOW) compositions of quartz (Qwf) associated with wolframite shows a mean value of 12.6‰ (1σ) and a mean value of 11.5‰ (1σ) for the quartz join to scheelite.
The δ18OFluid of the quartz-water pair (103lnαQuartz–water = 3.38 × 106T−2−3.40; [49]) yielded the following values for the precipitated quartz (Qwf) associated with wolframite: +5.71‰ (1σ) at 300 °C, +7.30‰ (1σ) at 350 °C, and 8.54 (1σ) at 400 °C. For the precipitated quartz (Qsch) associated with scheelite yielded: +4.61‰ (1σ) at 300 °C, +6.20‰ (1σ) at 350 °C, and +7.44‰ (1σ) at 400 °C (Table 5). The δ18O (V-SMOW) measured for wolframite yielded a value of 3.8‰ (1σ) (i.e., much lighter than quartz). The δ18OFluid of the wolframite-water pair (103lnαWf–water = 1.03 × 106T−2−4.96; and 103lnαWf–water = 0.21 × 106T−2−2.91 for temperature intervals between 250–370 °C and 370–420 °C [50]) yielded a value of 6.11‰ (1σ) for an estimated temperature of 300 °C and 6.31‰ (1σ) for 400 °C.
Combining the quartz-water fractionation equation [54] with the wolframite-water pair [50], the following oxygen isotope fractionation equations were proposed for the quartz-wolframite pair [63]: (i) 250–370 °C: 103lnαq-Wf = 2.35 × 106T−2 + 1.56 and (ii) 370–420 °C: 103lnαq-Wf = 3.17 × 106T−2–0.49. The quartz-wolframite oxygen isotope fractionation estimated for temperatures ranging from 300 °C to 400 °C yielded 9 and 7.11‰ (1σ). The data plotted (not shown) in the 103lnα vs. T °C [63] fall between 300 °C and 400 °C.
The δ18O isotope (V-SMOW) composition measured on scheelite yielded values of 2.8‰, 3‰, and 3.8‰ (1σ). The δ18OFluid of the scheelite-water pair (103lnαscheelite-water =1.39 × 106T−2−5.87; [52]) yielded 4.44‰, 4.64‰, and 5.44‰ (1σ) for 300 °C, 5.09‰, 5.29‰, and 6.09‰ (1σ) for 350 °C, and 5.60‰, 5.80‰, and 6.60‰ (1σ) for 400 °C (Table 5. The oxygen isotope fractionation in the quartz-scheelite and quartz-wolframite pairs is large, offering a much more sensitive geothermometer [64].
Combining the quartz-water fractionation equation [54] with the scheelite-water equation [52], the following fractionation equation was obtained for quartz-scheelite fractionation: 103lnαq-scheelite = 1.99 × 106T−2 + 2.47. The ∆quartz-scheelite calculated was 6.86‰ (1σ) for a T °C of 400 °C and 7.54‰ (1σ) for 350 °C. The ∆quartz-scheelite values plotted in the diagram of 103lnα vs. T °C (not shown) intersected the quartz-scheelite fractionation curve at about 350–400 °C.
The δ18O isotope (V-SMOW) signatures measured for Fe,Mn-chlorite yielded values of +3.1 and +3.2‰ ± 1.2‰, where the δ18OFluid calculated for the chlorite-water pair (103lnαchlorite-water = 1.56 × 106T−2 − 4.70; [53]) yielded +3.27‰ for a T °C of 350 °C, +4.46‰ for 400 °C (Table 5), and +4.92‰ for 450 °C. The ∆chlorite-water fractionation factor calculated was about +0.05‰ for 330 °C and values of −0.68‰ and −1.26‰ corresponded to temperatures of 380 °C and 450 °C, respectively. After plotting the data in the 106 × T−2 (°K) vs. 103lnαchlorite-water diagram (Figure 13) [65], an average crystallization temperature of 400 °C was obtained.

4.4. Chlorite Geothermometry

Several chlorite thermometers based on empirical calibrations (linking T to AlIV content) or semi-empirical models (linking T to the chlorite + quartz + water equilibrium constant) have been developed for Al-, Fe-, and Mg-rich chlorites over the last decade [58,59,60,67,68,69]. In a recent review on chlorite geothermometry, Bourdelle [29] noted that it could be worth applying these models for Mn-rich or Al-free chlorites to Mn-redox behavior, like Fe-redox.
In this study, 146 chemical analyses of Fe,Mn-chlorite (Table 2) were used for evaluating both empirical and semi-empirical models known in the literature. The average amount for FeO was about 39.05% and 6.05% for MnO, where the estimated Fe3+ was about 0.02 apfu. High contents of Fe [Fe/(Fe + Mg + Mn) = 0.4–0.95] and Mn [Mn/(Fe* + Mg) > 0.05] in chlorite minerals are also diagnostic of chemical environments related to ore-forming processes [70]. The Fe,Mn-chlorite temperature values obtained using semi-empirical geothermometers [58,59,60,71] are shown in Table 6.
The average calculated temperature of Fe,Mn-chlorite [59,70] was estimated at about 228 °C (Fetotal = FeO + Fe2O3) and 196 °C (Fetotal = FeO). The temperatures estimated for samples FMN 8 and FMN 9 were respectively about 435 °C and 519 °C (Fetotal = FeO) after Bourdelle [58] and 373 °C and 458 °C (Fetotal = FeO) after Lanari [59]. Here, the geothermometers of Bourdelle and Lanari [58,59] are not applicable (Table 7). The average temperature calculated for Fe,Mn-chlorite crystallization was about 373 °C, after Cathelineau [67], and 392 °C after Jowett [68]. The chemical composition of Fe,Mn-chlorite plotted in the Si4+ vs. R2+ diagram [61] falls into a field with temperatures ranging from 400 °C to 550 °C (Figure 6).

5. Discussion

The successful application of chlorite minerals and stable isotope geothermometry to ore deposits requires several conditions to satisfy any pair of minerals: (i) Fractionation factors between them are well-calibrated and strongly temperature dependent; (ii) the mineral(s) species were in isotopic equilibrium at the time of deposition; and (iii) they have retained their isotopic composition since formation. Therefore, the sequence mineral pairs identified in the W, (Cu,Mo) ore deposit of Borralha satisfy all the requirements to evaluate the isotope geochemistry relationship between tungstates and neoformed silicates (quartz and Fe,Mn-chlorite) and empirical chlorite geothermometers, based also on FI microthermometry data.

5.1. Wolframite to Scheelite and Fe,Mn-Chlorite Crystallization

The scheelite samples collected are excellent samples due to their size, purity, and their genetic relationship with Mn-bearing wolframite and Fe,Mn-chlorite. Wolframite crystallized from fluids with slightly acidic to neutral pH, where WO42− may be easily complexed with Fe2+, Mn2+, and Ca2+. It should be noted that the stability of the Fe2+ end-member (ferberite) is much more sensitive to the high variations of fO2 than the Mn2+ end-member wolframite (hübnerite) since the composition of ferberite might be strongly affected by the oxidation state [26]. Only Mn-bearing wolframite is associated with the late evolving W mineralization, possibly triggered by an alkalinity rising in the fluid as a result of an increasing S activity and the availability of Ca2+ in the system, which is correlated with a late alkaline metasomatic reaction common in geochemical settings [73]. Several authors have suggested that Mn-rich wolframites precipitated at higher temperatures than Fe-rich wolframites [73,74,75,76]. Nevertheless, the H/F ratio used as a geothermometer [77] failed, because temperature is not the only factor controlling the composition of wolframite [27]. The H/F ratio can be used as an indicator of contrasting W deposition environments in perigranitic ore-forming systems, where hübnerite can be viewed as a relatively simple indicator of a strong magmatic control on W deposition [76]. The H/F ratio measured in wolframite samples from Borralha ranged from 46.25% to 46.70%, as a consequence of the high Mn/Fe ratio in the magmatic-hydrothemal fluid. By contrast, a H/F ratio < 40 and down to 0 indicates a W deposition environment where a significant non-magmatic contribution to the ore fluid and W deposition should be necessary [78].
Experimental studies on the stability of H4[Si(W3O10)4]2H2O at high P-T conditions and variable pH in the presence of Ca2+ have confirmed the separation of silico-tungstic acid under the effect of solution neutralization [79], where [WO4]2− is preferentially combined with Ca2+ to form scheelite, favoring the incorporation of the available Fe2+ and Mn2+ from wolframite and SiO2 from the fluid into the chlorite structure. The CO2 loss from the hydrothermal fluid during boiling caused pH changes from acidic to alkaline conditions [80,81,82,83], accompanying the wolframite dissolution and scheelite deposition in the presence of Ca2+. Calcium for scheelite crystallization was supplied from the albitization and/or sericitization of plagioclase from granitic and pegmatite host rocks. Optical microscopy showed that the scheelitization of Mn-bearing wolframite occurred along fissures and fractures. This implies that high-pressure fluids are required to trigger fracture initiation and propagation, where the fluctuations of fluid pressure were recorded by FIs. Hence, Mn-bearing wolframite supported a dissolution process where [WO4]2− separation served for scheelite crystallization, whereas Mn and Fe for chlorite. It is important to note that the term “metamorphic fluid” is used here to describe a fluid that was partially equilibrated with metamorphic rocks and this does not necessarily imply that the fluid components were produced from prograde metamorphic reactions.

5.2. Characterization of the Mineralizing Fluid

Previous FIs studies on scheelite from Borralha revealed two generations of FIs (primary and pseudosecondary) characterized by the presence of an aqueous, low salinity fluid [38]. The fluids entrapped in scheelites showed a complex history with different fluids and boiling during W precipitation, where boiling (immiscibility), mixing, cooling, and fluid-rock interactions (e.g., Ca2+ needed for scheelite precipitation) played an important role in the scheelite precipitation. The primary fluid stage was characterized by a complex aqueous-carbonic fluids of low salinity, indicating a typical metamorphic fluid. The CO2 ubiquitous recorded in FIs from a variety of mineralized systems underwent boiling/unmixing presumably leading to ore deposition. Scheelite was mainly precipitated during the early stage from an aqueous-carbonic fluid with a CO2−dominated composition, where the Th of the ore-forming fluids decreased from 380 °C to 200 °C.
Fluid boiling/immiscibility has two fundamental effects on the fate of metals. Indeed, it is known to induce the precipitation of some metals from the liquid phase owing to enhanced metal concentrations and to change the redox conditions and ligand contents in the residual liquid as a result of the removal of volatile components (H2S, H2, HCl, and CO2) into the vapor [81,82,83]. The partitioning of CO2 and CH4 into the vapor phase occurred through boiling, where both participated in redox reactions. The coexistence of vapor- and liquid-rich inclusions indicates fluid boiling/immiscibility. The boiling resulted in the separation of the aqueous-carbonic fluid with predominant CH4 and mixing with an aqueous fluid of low salinity at about 330 °C, deduced from the histograms (Figure 8) and the diagram of salinity vs. Th (Figure 9). The salinity (wt.% NaCl equiv.) vs. Th (°C) diagram shows two distinct fields corresponding to a magmatic-hydrothermal vs. metamorphic fluid (Stages I and II) and a hydrothermal-meteoric fluid (Stage III) (Figure 9).
Stage I occurred around the maximum Th measured, corresponding to the beginning of separation between CO2-CH4 fluids during boiling (immiscibility). The L-rich FIs appeared to be in equilibrium with the CO2-CH4 inclusions (see Figure 10 and Figure 11a), which is another indication of boiling. Boiling also increased oxygen fugacity, leading to the immiscibility of CH4 where a disequilibrium isotopic fractionation occurred between CO2 and CH4 below 400 °C [83]. The process ended during the final of Stage II, when the aqueous saline FIs were generated. Changes in the aqueous-carbonic fluid composition were recognized as temperature decreased, where CH4 became dominant (Stage II) and the Th decreased (average of 262 °C). Stage III corresponded to the dilution of metamorphic fluids and cooling by the circulation of meteoric fluids. Exhaustion of deposition led to low salinity FIs. This stage corresponded to an aqueous fluid of lower temperature (average Th = 218.5 °C). In addition, the salinity gradually decreased as temperature decreased (Figure 10 and Figure 11c). The large dispersal of Th values in the histogram and the reduced salinity observed for the aqueous inclusions was indicative of mixing.
Experimental studies have demonstrated that the solubility of scheelite decreases substantially with decreasing temperatures [73], where tungstate destabilization was influenced by the pH changes of the fluid [84]. This process removed CO2 from the hydrothermal fluid, a component that can not only be a buffer fluid pH [73,85], but also favors W transport [86]. The trace amount of N2 addition to the mineralizing fluids derived from the devolatilization of metasediments [87].

5.3. Oxygen Isotopes

The δ18OFluid calculated for the quartz-water pair (sample Qwf) yielded values of + 5.7‰ at about 300 °C, + 7.30‰ at 350 °C, and + 8.54‰ at 400 °C, where the calculated values fall within the magmatic water field [88,89] from 5.5‰ to 10.0‰ [90,91,92]. In fact, from the isotopic point of view, quartz (Qwf) and wolframite (+ 6.11 to + 6.31‰ at 300 and 400 °C, respectively) crystallized directly from a magmatic-hydrothermal fluid, which may represent primary magmatic water or water from any source whose isotopic composition is controlled by an exchange with a large volume of igneous rock. Nevertheless, the δ18OFluid calculated for quartz (Qsch) and scheelite yielded values from 5.55‰ to 6.6‰ for sample SHS, for temperatures ranging from 300 °C to 400 °C; for sample 3N values of 4.44‰ to 5.8‰ were obtained in the same temperature range. The samples analyzed suggest several scheelite generations crystallized from a magmatic-hydrothermal vs. metamorphic fluid (sample SHS) at temperatures below 400 °C, which were later diluted by a hydrothermal meteoric fluid (sample 3N). In fact, the differentiation between magmatic and metamorphic fluids in terms of δ18OFluid is limited, creating difficult interpretations concerning the mixing of these two fluid phases. According with the FIs data, scheelite crystallization could be largely influenced by metamorphic fluids which equilibrated with the magmatic-hydrothermal fluid during the first wolframite crystallization. δ18OFluid data support an input of hydrothermal-meteoric waters, confirming an exchange and re-equilibration with the primary fluid at elevated temperatures and low water/rock ratios. This assumption is supported by FIs data obtained from the studied scheelite samples, where an input of aqueous FIs after boiling was observed (Figure 10). Based on δ18O data, it seems that Fe,Mn-chlorite crystallized in equilibrium with scheelite, but from a diluted hydrothermal-meteoric fluid contribution at a high temperature. The crystallization temperatures estimated from the 106 × T−2 (°K) vs. 103lnαchlorite-water diagram (Figure 13) [65,66] were between 350 °C to 450 °C, where the chlorite-water pair fractionation yielded values from +3.17‰ at 350 °C to +4.46‰ at 400 °C, confirming the presence of a hydrothermal-meteoric fluid in equilibrium with scheelite crystallization.
The mixing of magmatic and metamorphic fluids with different origin in granite systems has been inferred in the literature, where the external fluids (i.e., metamorphic, hydrothermal-meteoric) can mix and dilute the primary magmatic fluid, even when enriched by metals extracted from wall-rocks in the frame of the resulting convective circulation [93,94,95].

5.4. Chlorite Geothermometer

Chlorite consists of a 2:1 layer with a negative charge [(R2 +, R2 +)3(xSi4xR2 +y)O10OH2] balanced by a positively charged interlayer octahedral sheet [(R2 +, R3 +)3(OH)6]+. R2+,3+ is represented by Mg, Fe(II), and Fe(III) with Mg-rich chlorite (clinochlore) generally being metamorphic (high temperature), while the Fe-rich chlorite (chamosite) is typically diagenetic (low temperature). Additionally, Ni-rich chlorite (nimite) and Mn-rich chlorite (pennantite) are the other two less common varieties of chlorite. Most chlorites are trioctahedral in both sheets, i.e., the ferric iron content is low. Furthermore, it is well known that the ubiquitous and large thermodynamic stability of chlorite is in metamorphic, diagenetic, and hydrothermal systems, where its chemical composition is influenced by temperature and pressure. Thus, chlorite may crystallize over a large range of temperatures from 80 °C to more than 700 °C and pressures from 1 Kbar to 12 Kbar [96]. For example, cookeite, a Li-di,trioctaedral chlorite, is an example of a polymorph that has crystallized at low to high pressures [97,98].
The Fe,Mn-chlorite associated with W mineralizations from Borralha crystallized directly from the hydrothermal fluid and isotopically re-equilibrated with respect to scheelite. Hydrothermal fluids in the crust generally have a low oxidation potential and dissolved Fe predominantly occurs in its +2 oxidation state [99]. Under these reduced conditions, the amount of Fe3+ in hydrothermal fluids is negligible compared to that of Fe2+ [71]. Both Fe and Mn were available in the hydrothermal fluid from wolframite replaced by scheelite. Considering the textural relationships observed between wolframite to scheelite and Fe,Mn-chlorite, and the high amount of Fe and Mn in the hydrothermal fluid, chlorite crystallized after scheelite preserving the redox conditions during the fluid migration. The redox conditions are supported by the obtained FI data, where CO2 and CH4 subsequently reduced oxidation. By contrast, adequate oxidizing conditions were needed to buffer the oxygen activity in the case of a hydrothermal fluid with a Fe- or Mn-poor or -rich character and a low fluid/rock ratio. In this case, both Fe3+ and Mn3+ had a limited stability field in aqueous solution, being rapidly trapped during chlorite crystal growth.
Several authors have suggested that empirical or semi-empirical geothermometers have only been developed for the most common chlorites (i.e., Al-, Fe-, or Mg-rich chlorites) and it could worth extending to rarer chlorites, like Mn-rich chlorite [29], because the actual geothermometers do not include a Mn-rich end-member, such as pennantite. Nevertheless, the systematic presence of measurable amounts of Mn has been tested as a new pathfinder for chlorites developed in the course of ore-forming processes. A valid example is related to Fe,Mg,Mn-chlorite identified in Pb-Zn, Bi, and Ag ore veins from the Toyoha geothermal system, Japan [70]. In this case, the structural formula calculation, similar to ferric iron, will arise without the quantification of the total Mn3+/Mn ratio, but the ratio of Mn3+/Mn2+ would probably theoretically modify the chlorite’s structural formula [29]. Nevertheless, the presence of Mn3+ in chlorite, which would eventually be substituted with Al3+ in the octahedral site, is expected to lead to the structural distortion of chlorite due to the different ionic radii of Al3+ (0.535 Å) and Mn3+ (0.645 Å) [100].
The chemical compositions of Fe,Mn-chlorite are plotted in the field of daphnite (average Si = 5.30 apfu and Fe 7.51 apfu) in the Si vs. Fetot diagram (not shown) or in the field of daphnite-amesite according to the Si vs. R2+ diagram (Figure 6; [61]), which makes it valid for the four-thermodynamic-component solid-solution model (Mg-sudoite, Mg-amesite, clinochlore, and daphnite) used for semi-empirical geothermometers.
The estimated Fe,Mn-chlorite temperatures using the semi-empirical thermometer of Lanari [59] were about 373 °C and 458 °C (Fetotal = FeO) and 435 °C and 519 °C (Fetotal = FeO) (Bourdelle [58]). This represents average values of 415 and 477 °C. If we consider the uncertainly of 50–60 °C for T °C > 300 °C suggested by Bourdelle [58], we propose a temperature between 375 °C to 410 °C. The estimated temperatures calculated from semi-empirical thermometers seem to be supported by the δ18OFluid calculated using the fractionation equation for the chlorite-water pair [53] which confirmed a temperature of chlorite crystallization estimated at around 400 °C.
The magmatic-hydrothermal vs. metamorphic fluid circulation evidences was proposed in this work based on FI data and the oxygen stable isotope fingerprints obtained on tungstates, quartz, and Fe,Mn-chlorite (Figure 14). The presence of a metamorphic fluid with an aqueous-carbonic composition and low salinity was contributed to wolframite destabilization and scheelite precipitation in the presence of Ca2+ derived from plagioclase alteration. During boiling, the CO2 was removed from the aqueous-carbonic fluid, favoring W transport [84], where Fe,Mn-chlorite crystallized after scheelite crystallization at a temperature close to the crystallization of scheelite.

6. Conclusions on Fe,Mn-Chlorite Thermometry

The diversity of chlorite minerals in fossils or active hydrothermal systems is connected in most cases with the alteration types of magmatic rocks, where chlorite may crystallize directly from a solution or replace the primary minerals. The sequence of mineral pairs identified in the W, (Cu,Mo) ore deposit of Borralha satisfies all conditions to evaluate the oxygen isotope signature between tungstates and silicates. Besides, the question arises on Mn-oxidation, the effect of fluid chemistry on the mineral-water fractionation factor in chlorite may represent an uncertainty however, fractionation factors between minerals and aqueous solutions are independent of solution chemistry.
Our multiproxy study based on FIs, stable isotopes, and semi-empirical chlorite thermometers showed good consistency between the results obtained. Combining FIs and oxygen isotope geochemistry data, it was revealed that a metamorphic fluid supported the crystallization of scheelite and quartz at temperatures below 400 °C. The geothermometers applied according to Lanari and Bourdelle for Fe,Mn-chlorite produced temperature estimates between 373 °C and 458 °C, and 435 °C and 519 °C, respectively. The ∆chlorite-water calculated fractionation factor corresponded to temperatures of 380 °C and 450 °C, respectively. An average chlorite crystallization temperature of 400 °C was assumed considering an uncertainly of 50–60 °C for T °C > 300 °C, as suggested by the authors of the geothermometers.

Supplementary Materials

The following are available online: https://www.mdpi.com/article/10.3390/min12010024/s1. Table S1. The CO2, CH4 and N2 (mol%) of aqueous-carbonic fluid measured by Raman spectroscopy analysis (50 analyses).

Author Contributions

Conceptualization, I.B., C.M.d.S. and F.N.; Formal analysis, I.B.; Investigation, I.B. and C.M.d.S.; Methodology, I.B., C.M.d.S. and F.N.; Writing—original draft, I.B. and C.M.d.S.; Writing—review and editing, I.B., C.M.d.S. and F.N. All authors have read and agreed to the published version of the manuscript.

Funding

POCTEP-ESMIMET 2018–2020 Interreg Spain-Portugal Financed by POCTEP.

Acknowledgments

This work was developed in the frame of projects POCTEP-ESMIMET 2018–2020 Interreg Spain-Portugal Financed by POCTEP and UIDB/04683/2020 ICTerra, FCT -Lisboa. Carlos Sá acknowledge the post-doc scholarship attributed in the frame of this project during 2019–2020. Thanks also to ICTerra-Polo Porto for using the fluid inclusions lab. The first author is greatly indebted to F. Nieto and Franck Bourdelle for their thorough reading and helpful suggestions carried out on an early manuscript. The authors are grateful to the Academic Editor and to all three reviewers who contributed with many suggestions and greatly improved the final manuscript version. Also thanks to Mark Ryan who reviewed the English presentation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ren, Y.S.; Zhao, H.L.; Lei, E.; Wang, H.; Ju, N.; Wu, C.Z. Trace element and rare earth element geochemistry of the scheelite and ore genesis of the Yangjingou large scheelite deposit in Yanbian area, northeastern China, China. Acta Petrol. Sin. 2010, 26, 3720–3726. [Google Scholar]
  2. Song, G.; Qin, K.; Li, G.; Evans, N.G.; Chen, L. Scheelite elemental and isotopic signatures: Implications for the genesis of skarn-type W-Mo deposits in the Chizhou Area, Anhui Province, Eastern China. Am. Miner. 2014, 99, 303–317. [Google Scholar] [CrossRef]
  3. Poulin, R.S.; McDonald, A.M.; Kontak, D.J.; McClenaghan, M.B. On the relationship between cathodoluminescence and the chemical composition of scheelite from geologically diverse ore-deposit environments. Can. Miner. 2016, 54, 1147–1173. [Google Scholar] [CrossRef]
  4. Poulin, R.S.; Kontak, D.J.; McDonald, A.; McClenaghan, M.B. Assessing scheelite as an ore-deposit discriminator using its trace-element and REE chemistry. Can. Miner. 2018, 56, 265–302. [Google Scholar] [CrossRef]
  5. Guo, S.; Chen, Y.; Liu, C.-Z.; Wang, J.-G.; Su, B.; Gao, Y.-J.; Wu, F.-Y.; Sein, K.; Yang, Y.-H.; Mao, Q. Scheelite and coexisting F-rich zoned garnet, vesuvianite, fluorite, and apatite in calc-silicate rocks from the Mogok metamorphic belt, Myanmar: Implications for metasomatism in marble and the role of halogens in W mobilization and mineralization. J. Asian Earth Sci. 2016, 117, 82–106. [Google Scholar] [CrossRef]
  6. Wang, H.; Feng, C.; Zhao, Y.; Zhang, M.; Chen, R.; Chen, J. Ore genesis of the Lunwei granite-related scheelite deposit in the Wuyi Metallogenic belt, southeast China: Constraints from geochronology, fluid inclusions, and H-O–S isotopes. Res. Geol. 2016, 66, 240–258. [Google Scholar] [CrossRef]
  7. Wang, Y.C.; Wang, K.Y.; Konare, Y. N2-rich fluid in the vein-type Yangjingou scheelite deposit, Yanbian, NE China. Sci. Rep. 2018, 8, 5662. [Google Scholar] [CrossRef]
  8. Chen, J.; Sheng, D.; Shao, Y.; Zhang, J.; Liu, Z.; Wei, H.; Yang, Q.; Luo, X.; Du, Y. Silurian S-type granite-related W-(Mo) mineralization in the Nanling Range, south China: A case study of the Pingtan W-(Mo) deposit. Ore Geol. Rev. 2019, 107, 186–200. [Google Scholar] [CrossRef]
  9. Kilias, S.P.; Konnerup-Madsen, J. Fluid inclusion and stable isotope evidence for the genesis of quartz-scheelite veins, Metaggitsi area, central Chalkidiki Peninsula, N. Greece. Miner. Depos. 1997, 32, 581–595. [Google Scholar] [CrossRef]
  10. Sciuba, M.; Beaudoin, G.; Grzela, D.; Makvandi, S. Trace element composition of scheelite in orogenic gold deposits. Miner. Depos. 2019, 55, 1149–1172. [Google Scholar] [CrossRef]
  11. Meinert, L.D.; Dipple, G.M.; Nicolescu, S. World Skarn Deposits. In Economic Geology 100th Anniversary Volume; Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P., Eds.; Society of Economic Geologists, Inc.: Littleton, CO, USA, 2005; pp. 299–336. [Google Scholar]
  12. Li, W.; Xie, G.Q.; Cook, N.J.; Mao, J.W.; Li, C.; Ciobanu, C.L.; Zhang, Z.Y. Tracking dynamic hydrothermal processes: Textures, in-situ Sr-Nd isotopes and trace element analysis of scheelite from the Yangjiashan vein-type W deposit, South China. Am. Miner. 2021, 106, 1987–2002. [Google Scholar] [CrossRef]
  13. Wang, G.W.; Li, R.X.; Carranza, E.J.M.; Zhang, S.T.; Yan, C.H.; Zhu, Y.Y.; Qu, J.N.; Hong, D.M.; Song, Y.W.; Han, J.W. 3D geological modeling for prediction of subsurface Mo targets in the Luanchuan district, China. Ore Geol. Rev. 2015, 71, 270–274. [Google Scholar] [CrossRef]
  14. McClenaghan, M.B.; Cabri, L.J. Review of gold and platinum group element (PGE) indicator minerals methods for surficial sediment sampling. Geochem. Explor. Environ. Anal. 2011, 11, 251–263. [Google Scholar] [CrossRef]
  15. Mishra, B.; Pruseth, K.L.; Hazarika, P.; Chinnasamy, S.S. Nature and source of the ore-forming associated with orogenic gold-deposits in the Dharwar Craton. Geosci. Front. 2018, 9, 715–726. [Google Scholar] [CrossRef]
  16. Singoyi, B.; Zaw, K. A petrological and fluid inclusion study of magnetite–scheelite skarn mineralization at Kara, Northwestern Tasmania: Implications for ore genesis. Chem. Geol. 2001, 173, 239–253. [Google Scholar] [CrossRef]
  17. Zaw, K. The CanTung E-Zone Orebody, Northwest Territories: A Major Scheelite Skarn Deposit. Unpublished Master’s Thesis, Queen’s University, Kingston, ON, Canada, 1976. [Google Scholar]
  18. Zaw, K.; Clark, A.H. Fluoride–hydroxyl ratios of skarn silicates, CanTung E-zone scheelite orebody, Tungsten, Northwest Territories. Can. Miner. 1978, 16, 207–221. [Google Scholar]
  19. Song, S.W.; Mao, J.W.; Xie, G.Q.; Jian, W.; Chen, G.H.; Rao, J.F.; Ouyang, Y.P. Petrogenesis of scheelite-bearing albitite as an indicator for the formation of a world-class scheelite skarn deposit: A case study of the Zhuxi tungsten deposit. Econ. Geol. 2021, 116, 91–121. [Google Scholar] [CrossRef]
  20. Yuvan, J.; Shelton, K.; Falck, H. Geochemical investigations of the high-grade quartz-scheelite veins of the Cantung mine, Northwest Territories. Geol. Surv. Can. Open File 2007, 5344, 177–190. [Google Scholar]
  21. Xie, G.; Mao, J.; Li, W.; Fu, B.; Zhang, Z. Granite-related Yangjiashan tungsten deposit, southern China. Miner. Depos. 2019, 54, 67–80. [Google Scholar] [CrossRef]
  22. Wilkinson, J.J.; Zhaoshan Chang, Z.; Cooke, D.R.; Baker, M.J.; Wilkinson, C.C.; Inglis, S.; Chen, H.; Gemmell, J.B. The chlorite proximitor: A new tool for detecting porphyry ore deposits. J. Geochem. Explor. 2015, 152, 10–26. [Google Scholar] [CrossRef] [Green Version]
  23. Walshe, J.L. A six-component chlorite solid solution model and the conditions of chlorite formation in hydrothermal and geothermal systems. Econ. Geol. 1986, 81, 681–703. [Google Scholar] [CrossRef]
  24. Cathelineau, M.; Nieva, D. A chlorite slid solution geothermometer the Los Azufres (Mexico) geothermal system. Contrib. Miner. Petrol. 1985, 91, 235–244. [Google Scholar] [CrossRef]
  25. Sillitoe, R.H. Porphyry copper systems. Econ. Geol. 2010, 105, 3–41. [Google Scholar] [CrossRef] [Green Version]
  26. Hsu, L.C. The stability relations of the wolframite series. Am. Miner. 1976, 61, 944–955. [Google Scholar]
  27. Amossé, J. Physicochemical Study of the Hubnerite-Ferberite (MnWO4-FeWO4) zonal distribution in wolframite (MnxFe(1-x)WO4) Deposits. Application to the Borralha Mine. Phys. Chem. Miner. 1978, 3, 331–341. (In Portuguese) [Google Scholar] [CrossRef]
  28. Bobos, I.; Noronha, F.; Mateus, A. Fe-, Fe,Mn- and Fe,Mg-chlorite: A genetic linkage to W, (Cu,Mo)-mineralization in the magmatic-hydrothermal system of Borralha, Northern Portugal. Miner. Mag. 2018, 82, 259–279. [Google Scholar] [CrossRef] [Green Version]
  29. Bourdelle, F. Low-Temperature chlorite geothermometry and related recent analytical advances: A review. Minerals 2021, 11, 130. [Google Scholar] [CrossRef]
  30. Ribeiro, A.; Pereira, E.; Dias, R. Structure in the Iberian Peninsula. In Pre-Mesozoic Geology of Iberia; Dallmeyer, R.D., Martinez, G., Eds.; Springer: Berlin, Germany, 1990; pp. 220–236. [Google Scholar]
  31. Ribeiro, A.; Munhá, J.; Dias, R.; Mateus, A.; Peirera, E.; Ribeiro, L.; Fonseca, P.; Araújo, A.; Oliveira, T.; Romão, J.; et al. Geodynamic evolution of the SW Europe Variscides. Tectonics 2007, 26, 1–24. [Google Scholar] [CrossRef]
  32. Matte, P. Accretionary history and crustal evolution of the Variscan Belt in Western Europe. Tectonophysics 1991, 196, 309–337. [Google Scholar] [CrossRef]
  33. Ribeiro, A. Contribuition à l’étude tectonique de Trás-os-Montes Oriental. Comunic. Serv. Geológicos Port. 1974, 24, 1–168. [Google Scholar]
  34. Noronha, F.; Ramos, J.M.F.; Rebelo, J.; Ribeiro, A.; Ribeiro, M.L. Essai de corrélation des phases de déformation hercyniennes dans le NW de la Péninsule Ibérique. Leidse Geol. Meded. 1981, 52, 89–91. [Google Scholar]
  35. Ferreira, N.; Iglesias, M.; Noronha, F.; Pereira, E.; Ribeiro, A.; Ribeiro, M.L. Granitoides da Zona Centro Ibérica e seu enquadramento geodinâmico. In Geologia de Los Granitoides y Rocas Asociadas del Macizo Hesperico; Bea, F., Carnicero, A., Gonzalo, J., Lopez Plaza, M., Rodriguez Alonso, M., Eds.; Libro de Homenaje a García de Figuerola L.C.; Rueda: Madrid, Spain, 1987; pp. 37–51. [Google Scholar]
  36. Dias, R.; Ribeiro, A. The Ibero-Armorican Arc: A collision effect against an irregular continent? Tectonophysics 1995, 246, 113–128. [Google Scholar] [CrossRef]
  37. Noronha, F. Les brèches du gisement de tungstène de Borralha (Nord du Portugal). Quelques données pour leur classification. Public. Domus. Laboratório Mineral. Geológico Fac. Ciências Porto 1979, 4, 187–212. [Google Scholar]
  38. Noronha, F. Caractéristiques physico-chimiques des fluides associés à la génese du gisement de tungsten de Borralha (Nord Portugal). Bull. Miner. 1984, 107, 273–284. [Google Scholar] [CrossRef]
  39. Noronha, F.; Vindel, E.; López, J.A.; Dória, A.; García, E.; Boiron, M.C.; Cathelineau, M. Fluids related to tungsten ore deposits in Northern Portugal and Spanish Central system: A comparative study. Rev. Soc. Geol. España 1999, 12, 397–403. [Google Scholar]
  40. Noronha, F. Metalogenic Study of the Wore Deposit of Borralha. Ph.D. Thesis, Universidade do Porto, Porto, Portugal, 1983. Unpublished work (In Portuguese). [Google Scholar]
  41. Roedder, E. Fluid inclusions. Rev. Miner. Geochem. 1984, 12, 644. [Google Scholar]
  42. Van den Kerkhof, A.M.; Hein, U.F. Fluid Inclusion Petrography. Lithos 2001, 55, 27–47. [Google Scholar] [CrossRef]
  43. Shepherd, T.J.; Rankin, A.H.; Alderton, D.H.M. A Practical Guide to Fluid Inclusion Studies; Blackie & Son Ltd.: Glasgow, UK, 1985; 239p. [Google Scholar]
  44. Poty, B.; Leroy, J.; Jachimowicz, L. Un nouveil appareil pour la mesure des temperatures sous le microscope: I: Installation de microthermometrie Chaixmeca. Bull. Miner. 1976, 99, 182–186. [Google Scholar]
  45. Bakker, R.J. Package FLUIDS Computer programs for analysis of fluid inclusion data and modeling bulk fluid properties. Chem. Geol. 2003, 194, 3–23. [Google Scholar] [CrossRef]
  46. Touret, J. The significance of fluid inclusionsin metamorphic rocks. In Thermodynamics in Geology; Fraser, D.G., Ed.; Amsterdam Reidel Pub. Co.: Dordrecht, The Netherlands, 1977; pp. 203–227. [Google Scholar]
  47. Ramboz, C.; Schnapper, D.; Dubessy, J. The P-V-T-X-FO2 evolution of H2O-CO2-CH4 bearing fluid in a wolframite vein: Reconstruction from fluid inclusion studies. Geochim. Cosmochim. Acta 1985, 49, 205–219. [Google Scholar] [CrossRef]
  48. Prieto, A.C.; Guedes, A.; Doria, A.; Noronha, F.; Jimenez, J. Quantitative determination of gaseous phases in fluid inclusions by Raman microspectrometry. Spectr. Lett. 2012, 45, 1–5. [Google Scholar] [CrossRef]
  49. Clayton, R.N.; Mayeda, T.K. The use of bromide pentafluoride in the extraction of oxygen from oxides and silicates for isotopic analysis. Geochim. Cosmochim. Acta 1963, 27, 43–52. [Google Scholar] [CrossRef]
  50. Zhang, L.; Liu, J.; Zhou, H.; Chen, Z. Oxygen isotope fractionation in the quartz-water-salt system. Econ. Geol. 1989, 84, 1643–1650. [Google Scholar] [CrossRef]
  51. Landis, G.P.; Rye, R.O. Geologic, fluid inclusion, and stable isotope studies of the Pasto Bueno tungsten-base metal ore deposit northern Peru. Econ. Geol. 1974, 69, 1025–1059. [Google Scholar] [CrossRef]
  52. Wesolowski, D.; Ohmoto, H. Calculated oxygen isotope fractionation factors between water and the minerals scheelite and powellite. Econ. Geol. 1986, 81, 471–477. [Google Scholar] [CrossRef]
  53. Wenner, D.B.; Taylor, H.R., Jr. Temperatures of serpentinization of ultramafic rocks based on 18O/16O fractionation between coexisting serpentine and magnetite. Contrib. Miner. Petrol. 1971, 32, 165–168. [Google Scholar] [CrossRef]
  54. Clayton, R.N.; O’Neil, J.R.; Mayeda, T.K. Oxygen isotope exchange between quartz and water. Geophys. Res. 1972, 77, 3057–3067. [Google Scholar] [CrossRef]
  55. Brugger, J.; Giere, R.; Grobe, B.; Uspenski, E. Scheelite -powellite and paraniite-(Y) from the Fe-Mn deposit at Fianel, Eastern Swiss Alps. Am. Miner. 1998, 83, 1100–1110. [Google Scholar] [CrossRef]
  56. Bailey, S.W. Clinochlores: Structures and crystal chemistry, In: Hydrous Phyllosilicates (S.W. Bailey, editor), Mineralogical Society of America, Washington. Rev. Miner. 1988, 19, 398–404. [Google Scholar]
  57. Nieto, F. Chemical composition of metapelitic chlorites: X-ray diffraction and optical property approach. Eur. J. Miner. 1997, 9, 829–841. [Google Scholar] [CrossRef] [Green Version]
  58. Bourdelle, F.; Parra, T.; Chopin, C.; Beyssac, O. A new chlorite geothermometer for diagenetic to low-grade metamorphic conditions. Contrib. Miner. Petrol. 2013, 165, 723–735. [Google Scholar] [CrossRef]
  59. Lanari, P.; Wagner, T.; Vidal, O. A thermodynamic model for di-trioctahedral chlorite from experimental and natural data in the system MgO–FeO–Al2O3–SiO3–H2O: Applications to P–T sections and geothermometry. Contrib. Miner. Petrol. 2014, 167, 968–984. [Google Scholar] [CrossRef] [Green Version]
  60. Inoue, A.; Meunier, A.; Patrier-Mas, P.; Rigault, C.; Beaufort, D.; Vieillard, P. Application of chemical geothermometry to low-temperature trioctahedral chlorites. Clays Clay Miner. 2009, 57, 371–382. [Google Scholar] [CrossRef]
  61. Wiewióra, A.; Weiss, Z. Crystallochemical classifications of phyllosilicates based on the unified system of projection of chemical composition: II. The chlorite group. Clay Miner. 1990, 25, 83–92. [Google Scholar] [CrossRef]
  62. Jacobson, C.E. Estimation of Fe3+ from electron microprobe analyses: Observations on calcic amphibole and chlorite. J. Metamorf. Geol. 1989, 7, 507–513. [Google Scholar] [CrossRef]
  63. Shieh, Y.N.; Zhang, G.X. Stable isotope studies of quartz-vein type tungsten deposits in Dajistan mine, Jiangxi Province, Southeast China. In Stable Isotope Geochemistry: A Tribute to Samuel Epstein; Taylor, H.P., O’Neil, J.R., Kaplan, I.R., Eds.; SPE Publ.: London, UK, 1991; Volume 3, pp. 425–435. [Google Scholar]
  64. Zheng, Y.-F. Oxygen isotope fractionation in wolframite. Eur. J. Miner. 1992, 4, 1133–1135. [Google Scholar] [CrossRef]
  65. Wenner, D.B.; Taylor, H.P., Jr. Oxygen and hydrogen isotope studies of the serpentinization of ultramafic rocks in oceanic environments and continental complexes. Am. J. Sci. 1973, 273, 207–239. [Google Scholar] [CrossRef]
  66. Cole, D.R. A preliminary evaluation of oxygen isotopic exchange between chlorite–water. Geol. Soc. Am. Annu. Meet. 1985, 17, 550. [Google Scholar]
  67. Cathelineau, M. Cation site occupancy in chlorites and illites as a function of temperature. Clay Miner. 1988, 23, 471–485. [Google Scholar] [CrossRef]
  68. Jowett, E.C. Fitting Iron and Magnesium into the Hydrothermal Chlorite Geothermometer. In Proceedings of the GAC/MAC/SEG Joint Annual Meeting, Toronto, ON, Canada, 27–29 May 1991. [Google Scholar]
  69. Bourdelle, F.; Cathelineau, M. Low-temperature chlorite geothermometry: A graphical representation based on a T–R2+–Si diagram. Eur. J. Miner. 2015, 27, 617–626. [Google Scholar] [CrossRef]
  70. Inoue, A.; Kurokawa, K.; Hatta, T. Application of chlorite geothermometry to hydrothermal alteration in Toyoha geothermal system, southwestern Hokkaido, Japan. Res. Geol. 2010, 60, 52–70. [Google Scholar] [CrossRef]
  71. Inoue, A.; Inoue, S.; Utada, M. Application of chlorite thermometry to estimation of formation temperature and redox conditions. Clay Miner. 2018, 53, 143–158. [Google Scholar] [CrossRef]
  72. Verdecchia, S.O.; Collo, G.; Zandomeni, P.S.; Wunderlin, C.; Fernhrmann, M. Crystallochemical index and geothermobarometric calcultions as a multiproxy approach to P-T condition of the low-grade metamorphism: The case of he San Luis Formation, eastern Sierras Pampeanas. Lithos 2019, 324–325, 385–401. [Google Scholar] [CrossRef]
  73. Wood, S.A.; Samson, I.M. The hydrothermal geochemistry of tungsten in granitoid environments: I. Relative solubilities of ferberite and scheelite as a function of T, P, pH and mHCL. Econ. Geol. 2000, 95, 143–182. [Google Scholar] [CrossRef] [Green Version]
  74. Oelsner, O. Über Erzgebirgische Wolframite. Ber. Freib. Geol. Ges. 1944, 20, 44–49. [Google Scholar]
  75. Leutwein, F. Die chemische Zusammensetzung der Wolframite und ihre lagerstttenkundliche Bedeutung. Acta Geol. Akad. Sci. Hungary 1952, 1, 133–141. [Google Scholar]
  76. Taylor, R.G.; Hosking, K.F.G. Manganese-iron ratios in wolframites, South Crofty Mine, Cornwall. Econ. Geol. 1970, 65, 47–53. [Google Scholar] [CrossRef]
  77. Moore, F.; Howie, R.A. On the application of the hubnerite ferberite ratio as a Geothermometer. Miner. Depos. 1978, 13, 391–397. [Google Scholar] [CrossRef]
  78. Michaud, J.A.; Pichavant, M. The H/F ratio as an indicator of contrasted wolframite deposition mechanisms. Ore Geol. Rev. 2019, 104, 266–272. [Google Scholar] [CrossRef] [Green Version]
  79. Lecumberri-Sanchez, P.; Vieira, R.; Heinrich, C.A.; Pinto, F.; Wälle, M. Fluid-rock interaction is decisive for the formation of tungsten deposits. Geology 2017, 45, 579–582. [Google Scholar] [CrossRef]
  80. Baumer, A.; Caruba, R.; Guy, B. Experimental study of hydrothermal transformation scheelite in ferberite: Preliminary results. Bull. Miner. 1985, 108, 15–20. [Google Scholar] [CrossRef]
  81. Reed, M.H.; Spycher, N. Calculation of high temperature pH and mineral equilibria in hydrothermal waters, with application to geothermometry and studies of boiling and dilution. Geochim. Cosmochim. Acta 1984, 48, 1479–1492. [Google Scholar] [CrossRef]
  82. Reed, M.H.; Spycher, N. Boiling, cooling, and oxidation in epithermal systems: A numerical modeling approach. In Geology and Geochemistry of Epithermal Systems; Berger, B.R., Bethke, P.M., Eds.; Soc. Econ. Geologists: Littleton, CO, USA, 1985; Volume 2, pp. 249–272. [Google Scholar]
  83. Drummond, S.E.; Ohmoto, H. Chemical evolution and mineral depositionin boiling hydrothermal systems. Econ. Geol. 1985, 80, 126–147. [Google Scholar] [CrossRef]
  84. Wood, S.A.; Samson, I.M. Solubility of ore minerals and complexation of ore metals in hydrothermal solutions. Rev. Econ. Geol. 1998, 10, 33–80. [Google Scholar]
  85. Phillips, G.N.; Evans, K.A. Role of CO2 in the formation of gold deposits. Nature 2004, 429, 860–863. [Google Scholar] [CrossRef] [PubMed]
  86. Higgins, N.C. Fluid inclusion evidence for the transportof tungsten by carbonate complexes in hydrothermal solutions. Can. J. Earth Sci. 1980, 17, 823–830. [Google Scholar] [CrossRef]
  87. Kreulen, R.; Schuiling, R.D. N2- CH4-CO2 during formation of the Dome de l’Agoul, France. Geochim. Cosmochim. Acta 1981, 46, 193–203. [Google Scholar] [CrossRef]
  88. Taylor, H.P., Jr. The application of oxygen and hydrogen isotope studies to problems of hydrothermal alteration and ore deposition. Econ. Geol. 1974, 69, 843–883. [Google Scholar] [CrossRef]
  89. Taylor, H.P., Jr. Oxygen and hydrogen isotope relationships in hydrothermal mineral deposits. In Geochemistry of Hydrothermal Ore Deposits, 2nd ed.; Barnes, H.L., Ed.; Wiley: New York, NY, USA, 1979; pp. 236–277. [Google Scholar]
  90. Ohmoto, H. Stable isotope geochemistry of ore deposits. In Stable Isotopes in High Temperature Geological Processes; Valley, J.W., Taylor, H.P., Jr., O’Neil, J.R., Eds.; Mineralogical Society of America: Chantilly, VA, USA, 1986; Volume 16, pp. 491–559. [Google Scholar]
  91. Sheppard, S.M.F. Characterization and isotopic variations in natural water. In Stable Isotopes in High Temperature Geological Processes; Valley, J.W., Taylor, H.P., Jr., O’Neil, J.R., Eds.; De Gruyter: Berlin, Germany, 1986; Volume 16, pp. 165–183. [Google Scholar]
  92. Taylor, H.P., Jr. Oxygen and hydrogen isotope relationships in hydrothermal mineral deposits. In Geochemistry of Hydrothermal Ore Deposits, 3rd ed.; Barnes, H.L., Ed.; John Wiley & Sons: New York, NY, USA, 1997; pp. 229–302. [Google Scholar]
  93. Marignac, C.; Cathelineau, M. The Nature Ore-Forming Fluids in Peri-Batholitic Sn-W Deposis; James Cook University: Townsville, Australia, 2018. [Google Scholar]
  94. Polya, D.A.; Foxford, K.A.; Stuart, F.; Boyce, A.; Fallick, A.E. Evolution and paragenetic content of low D hydrothermal fluids from Panasqueira W-Sn deposit, Portugal: New evidence from microthermometric, stable isotope, nobe gas and halogen analyses of primary fluid inclusions. Geochim. Cosmochim. Acta 2000, 64, 3357–3371. [Google Scholar] [CrossRef]
  95. Wilkinson, J.J. The role of metamorphic fluids in the development of the Cornubian prefield: Fluid inclusion evidence from south Cornwall. Miner. Mag. 1990, 54, 219–230. [Google Scholar] [CrossRef]
  96. Vidal, O.; Goffeé, B. Cookeite experimental study and thermodynamic analysis of its compatibility relations in the Li2O-Al2O3-SiO2-H2O system. Contrib. Miner. Petrol. 1991, 108, 72–81. [Google Scholar] [CrossRef]
  97. Bobos, I.; Vieillard, P.; Charoy, B.; Noronha, F. Alteration of spodumene to cookeite and its thermodynamic stability. Clays Clay Miner. 2007, 57, 296–304. [Google Scholar]
  98. Goffeé, B.; Azanõn, J.M.; Bouybaouene, M.L.; Jullien, M. Metamorphic cookeite in Alpine metapelites from the Rif (Morocco) and Betic chain (Spain). Eur. J. Miner. 1996, 8, 335–348. [Google Scholar] [CrossRef]
  99. Heinrich, C.A.; Seward, T.M. A spectrophotometric study of aqueous iron (II) chloride complexing from 25 to 200 °C. Geochim. Cosmochim. Acta 1990, 54, 2207–2221. [Google Scholar] [CrossRef]
  100. Burns, R.G. Mineralogical Applications of Crystal Field Theory; Cambridge University Press: New York, NY, USA, 1993. [Google Scholar]
Figure 1. (a) Geological map of the northern part of Portugal and location of the W, (Cu,Mo) ore deposit of Borralha, Gerês Mountains [35]. (b) Geological map of the Borralha region [40]. (Both figures are reproduced from Bobos et al. [28] with permission of the Mineralogical Society of Great Britain and Ireland).
Figure 1. (a) Geological map of the northern part of Portugal and location of the W, (Cu,Mo) ore deposit of Borralha, Gerês Mountains [35]. (b) Geological map of the Borralha region [40]. (Both figures are reproduced from Bobos et al. [28] with permission of the Mineralogical Society of Great Britain and Ireland).
Minerals 12 00024 g001
Figure 2. Photomicrographs showing textural relationships between wolframite and scheelite: (a) Wolframite (white field) replaced by scheelite (gray); (b,c) precipitation of scheelite crystals along to fissures or fractures of wolframite; and (d) inclusions of sulphide (pyrite) were identified along to fractures of wolframite.
Figure 2. Photomicrographs showing textural relationships between wolframite and scheelite: (a) Wolframite (white field) replaced by scheelite (gray); (b,c) precipitation of scheelite crystals along to fissures or fractures of wolframite; and (d) inclusions of sulphide (pyrite) were identified along to fractures of wolframite.
Minerals 12 00024 g002
Figure 3. X-ray maps of Fe, Mn, and Ca distribution during wolframite scheelitization.
Figure 3. X-ray maps of Fe, Mn, and Ca distribution during wolframite scheelitization.
Minerals 12 00024 g003
Figure 4. X-ray maps of W and Ca distribution in scheelite (reproduced from Bobos et al. [28] with permission of the Mineralogical Society of Great Britain and Ireland).
Figure 4. X-ray maps of W and Ca distribution in scheelite (reproduced from Bobos et al. [28] with permission of the Mineralogical Society of Great Britain and Ireland).
Minerals 12 00024 g004
Figure 5. Photomicrograph showing radial aggregates of Fe,Mn-chlorite (NX).
Figure 5. Photomicrograph showing radial aggregates of Fe,Mn-chlorite (NX).
Minerals 12 00024 g005
Figure 6. Chemical compositions of Fe,Mn-chlorite plotted in the R2+-Si diagram [61]. Temperature estimation after Bourdelle [58].
Figure 6. Chemical compositions of Fe,Mn-chlorite plotted in the R2+-Si diagram [61]. Temperature estimation after Bourdelle [58].
Minerals 12 00024 g006
Figure 7. Photomicrographs showing: zonings and trails of fluid inclusions (a,b) and isolated fluid inclusions in sample SHS2 (c).
Figure 7. Photomicrographs showing: zonings and trails of fluid inclusions (a,b) and isolated fluid inclusions in sample SHS2 (c).
Minerals 12 00024 g007
Figure 8. (a) Isolated bi-phasic primary fluid inclusions with a size of over 100 μm (sample SHS2, objective 5×). Other primary isolated fluid inclusions (back and front of the big isolated fluid inclusions) and intra-granular trails of pseudo-secondary fluid inclusions. (b) Isolated fluid inclusions in a negative-crystal shape (sample SHS1, objective 50×). Primary negative-crystal shape fluid inclusions and pseudo-secondary oval-shaped fluid inclusions in intra-granular trails are observed in the back side.
Figure 8. (a) Isolated bi-phasic primary fluid inclusions with a size of over 100 μm (sample SHS2, objective 5×). Other primary isolated fluid inclusions (back and front of the big isolated fluid inclusions) and intra-granular trails of pseudo-secondary fluid inclusions. (b) Isolated fluid inclusions in a negative-crystal shape (sample SHS1, objective 50×). Primary negative-crystal shape fluid inclusions and pseudo-secondary oval-shaped fluid inclusions in intra-granular trails are observed in the back side.
Minerals 12 00024 g008
Figure 9. Aqueous-carbonic fluid inclusions: (a) Sample 3N1.6a and b predominantly CH4 phase. (b) Sample 3N2.1c predominantly CO2 phase. (c) Sample SHS1.3a predominantly CO2 phase. (d) Sample SHS2.5b predominantly CO2 phase.
Figure 9. Aqueous-carbonic fluid inclusions: (a) Sample 3N1.6a and b predominantly CH4 phase. (b) Sample 3N2.1c predominantly CO2 phase. (c) Sample SHS1.3a predominantly CO2 phase. (d) Sample SHS2.5b predominantly CO2 phase.
Minerals 12 00024 g009
Figure 10. The diagram of salinity (wt.% NaCl eq.) vs. Th (°C) corresponding to fluids identified in scheelite.
Figure 10. The diagram of salinity (wt.% NaCl eq.) vs. Th (°C) corresponding to fluids identified in scheelite.
Minerals 12 00024 g010
Figure 11. Histograms of Th (a), Tm (b), and salinity of fluid inclusions (c) in scheelite from Borralha.
Figure 11. Histograms of Th (a), Tm (b), and salinity of fluid inclusions (c) in scheelite from Borralha.
Minerals 12 00024 g011aMinerals 12 00024 g011b
Figure 12. The ternary diagram of CO2-CH4-N2.
Figure 12. The ternary diagram of CO2-CH4-N2.
Minerals 12 00024 g012
Figure 13. Estimation the oxygen isotope fractionation ratio between Fe,Mn-chlorite and water using the diagram of 106 × T−2 (°K) vs. 103lnαchlorite-water [53,66].
Figure 13. Estimation the oxygen isotope fractionation ratio between Fe,Mn-chlorite and water using the diagram of 106 × T−2 (°K) vs. 103lnαchlorite-water [53,66].
Minerals 12 00024 g013
Figure 14. Synthesis of microthermometry and oxygen isotope data correlated with semi-empirical chlorite thermometers [58,59].
Figure 14. Synthesis of microthermometry and oxygen isotope data correlated with semi-empirical chlorite thermometers [58,59].
Minerals 12 00024 g014
Table 1. Electron microprobe data of wolframite and scheelite and calculus of structural formula based on four oxygens.
Table 1. Electron microprobe data of wolframite and scheelite and calculus of structural formula based on four oxygens.
Oxides10 Analyses MinimumFe,Mn-W MaximumFe,Mn-W AverageSigma14 Analyses MinimumScheelite MaximumScheelite AverageSigma
WO375.1276.3575.941.2379.2980.8580.342.14
FeO 12.8812.510.7 0.070.020.0738
MnO 11.0310.620.66 0.130.030.00994
CaO 0.0100.0024 19.3219.180.68
Sc2O3 0.040.010.0274 0.020.010.02
Nb2O5 0.260.180.13 0.20.040.164
SnO2 0.030.010.031 0.130.010.126
Ta2O5 0.140.040.012 0.30.060.2951
TiO2 000 0.040.010.0172
V2O3 0.040.020.02 0.080.020.0581
Bi2O3 0.030.010.018 0.060.020.0453
MoO3 0.2410.090.0321
Sb2O5 0.050.020.02 000
Total 100.8699.36 100.4799.81
W 0.991 0.991
Fe 0.540.53 00
Mn 0.470.46 00
Ca 00 0.980.99
Sc 00 00
Nb 0.010 00
Sn 00 00
Ta 00 00
Ti 00 00
V 00 00
Bi 00 00
Mo 0.020
Sb 00 00
Total 2.0071.99 1.9941.994
Table 2. Electron microprobe data of Fe,Mn-chlorite and calculus of structural formula based on 14 oxygens (Fe2+/Fe3+ and OH calculated assuming full site occupancy).
Table 2. Electron microprobe data of Fe,Mn-chlorite and calculus of structural formula based on 14 oxygens (Fe2+/Fe3+ and OH calculated assuming full site occupancy).
OxidesFMN 1FMN2FMN 3FMN 12FMN 8FMN 9Averages 146 Analyses
N = 31N = 29N = 19N = 22N = 34N = 11
SiO223.1822.723.223.1123.7322.4622.82
TiO20.0800.020.010.040.040.06
Al2O319.8819.9519.8219.8220.1420.8620.37
FeO39.5239.8639.5339.1339.5538.5939.055
MnO6.235.035.826.215.285.816.02
MgO1.871.441.881.671.751.011.44
CaO0.040.010.030.050.050.060.05
Na2O0.030.0300.0100.010.02
K2O0.0100000.010.01
Total corrected by Cl + F90.8489.0290.390.0290.5488.8589.845
Si2.642.642.662.662.692.612.62
Al IV1.361.361.341.341.311.391.37
Total T site4444444
Al VI1.311.371.331.341.391.461.385
Ti0.01000000.005
Fe2+3.773.873.783.763.753.753.76
Mn0.60.50.560.60.510.570.585
Mg0.320.250.320.290.30.170.245
Ca0000.010.010.010.005
Na0.010.0100000.005
K0000000
Total O site6.0266.0165.965.975.995
Total Cations10.021010.01109.969.979.995
OH8888888
Mg/(Fe + Mg)0.080.060.080.070.070.040.06
Wt% H2O calculated10.5210.3210.4810.4310.5710.3310.425
Total Wt% (plus H2O)101.3699.34100.78100.45101.1199.18100.27
Vacancies
(AlVI-AlIV)/2 [58]−0.020−0.0100.040.030.005
(AlVI Al IV-Na-K)/2 [59]−0.020−0.0100.040.030.005
Calculation with Fe3+ estimated00000.080.060.02
Mg/(Fe2 + Mg)0.080.060.080.070.070.040.06
Vacancies
6-(Fe2 + Fe3 + Mg + AVI) [60]0.60.50.560.610.560.620.61
(AlVI-AlIV + Fe3-Na-K)/2 [59]−0.020−0.0100.040.030.005
Table 3. Petrographic characteristics of fluid inclusions observed in scheelite. Degree of fill by liquid water is an approximate range. Size is generalized, the biggest fluid inclusions is 160 μm in the 3N1 sample (see photo Figure 8a). Total number of studied fluid inclusions was n = 190.
Table 3. Petrographic characteristics of fluid inclusions observed in scheelite. Degree of fill by liquid water is an approximate range. Size is generalized, the biggest fluid inclusions is 160 μm in the 3N1 sample (see photo Figure 8a). Total number of studied fluid inclusions was n = 190.
SamplesTypeMode of OcurrencesShapeNumber of PhasesSize (μm)Φw (%)
3N1 and 3N2PrimaryIsolatedIrregular or rectangular2>2050–80
3N1 and 3N2PseudosecondaryIntragranular trails and clustersNegative crystal, rectangular or oval2<3070–80
SHS1 and SHS2PrimaryIsolatedNegative crystal and irregular2>2050–70
SHS1 and SHS2PseudosecondaryIntragranular trails and clustersRectangular and negative crystal2<3070–95
Table 4. Microthermometric data for aqueous-carbonic fluid. AqCO2—aqueous CO2 FI; AqCH4—aqueous CH4 FI; N—number of analysis; Φw—degree of fill liquid water; Te—eutectic temperature; Tmice—ice melting temperature; Tclat—clathrate melting temperature; Th—homogenization temperature; Sal.—salinity in weight percent equivalents of NaCl; Xgas—amount of CO2 or CH4 in mol%; D—density in g/cc; Min-Max Av.—minimum, maximum, average.
Table 4. Microthermometric data for aqueous-carbonic fluid. AqCO2—aqueous CO2 FI; AqCH4—aqueous CH4 FI; N—number of analysis; Φw—degree of fill liquid water; Te—eutectic temperature; Tmice—ice melting temperature; Tclat—clathrate melting temperature; Th—homogenization temperature; Sal.—salinity in weight percent equivalents of NaCl; Xgas—amount of CO2 or CH4 in mol%; D—density in g/cc; Min-Max Av.—minimum, maximum, average.
Type Φw (%)Te (°C)Tmice (°C)Tclat (°C)Th (°C)Sal (wt.eq. NaCl)Xgas (mol%)D (g/cc)
AqCO2 (n = 79)Min-max50 to 80−75.7 to −50−7.2 to 00.2 to 11.5200 to 3803.5 to 5.580 to 1000.55 to 1
Av.70−59−2.35.92844.894.20.77
AqCH4 (n = 11)Min-max60 to 80−68.7 to −52−6.3 to 07 to 10.4220 to 3234.3 to 5.11000.6 to 0.8
Av.75−59−2.59.42624.71000.8
Aqueous (n = 96)Min-max50 to 95−67 to −23.1−5 to 0 130 to 2650 to 6.2 1 to 1.0
Av.82−54.8−0.75 218.51.18 1.01
Table 5. Oxygen isotope data of quartz, wolframite, scheelite and Fe,Mn-chlorite minerals.
Table 5. Oxygen isotope data of quartz, wolframite, scheelite and Fe,Mn-chlorite minerals.
SamplesLocationδ18O (0) μεασυρεδQuartz-Water PairWolframite-Water PairScheelite-Water PairChlorite-Water Pair
300 °C350 °C400 °C300 °C350 °C400 °C300 °C350 °C400 °C300 °C350 °C400 °C
Quartz (wf)SHS
48 m
12.65.717.38.54---------
Quartz (sch)3N11.54.616.27.44---------
WolframiteSHS-
48 m
3.8---6.11-6.31------
Scheelite3N2.8------4.445.095.6---
Scheelite3N3------4.645.295.8---
ScheeliteSHS
48 m
3.8------5.446.096.6---
Fe,Mn-
chlorite
3N3.1---------3.093.174.36
Fe,Mn-
chlorite
SHS3.2---------3.153.274.46
Table 6. Calculated temperatures using semi-empirical thermometry.
Table 6. Calculated temperatures using semi-empirical thermometry.
Samples
Calculated Temperatures (°C)
FMN1FMN2FMN3FMN12FMN8FMN9
Semi-Empirical ThermometryT °CT °CT °CT °CT °CT °C
Inoue et al. [60,71].
FeTotal = FeO + Fe2O3. Quadratic equation thermometer (valid for <350°, may be <400 °C)
229246230220212228
Inoue et al. [71].
FeTotal = FeO (valid for <350 °C, may be <400 °C)
197210197190183196
Bourdelle et al. [58]
FeTotal = FeO. (Valid for <350 °C and <4 kb)
----435519
Lanari et al. [59].
Chl1: FeTotal = FeO + Fe2O3 (valid between 100–500 °C)
----718984
Lanari et al. [59]
Chl2: FeTotal = FeO (valid between 100–500 °C and 1–20 kbar)
----373458
Table 7. Calculated temperatures using empirical thermometry *.
Table 7. Calculated temperatures using empirical thermometry *.
Samples
Empirical Thermometry
(FeTotal = FeO)
FMN1
T °C
FMN2
T °C
FMN3
T °C
FMN12
T °C
FMN8
T °C
FMN9
T °C
Cathelineau [66]. (valid for <350 °C) 375376371371359387
Jowett [67]. For Si < 3.3 apfu and Ca < 0.07 apfu. 393395389389377406
Fe/(Fe + Mg) < 0.6, and T° < 325 °C
Restrictions of each thermometer calibration
(1 = Ok; 0 = not valid)
Bourdelle et al. [58].NOT APPLICABLE NOT APPLICABLE NOT APPLICABLE NOT APPLICABLE NOT APPLICABLE NOT APPLICABLE
Uncertainty of 30 °C for T < 300 °C and 50–60 °C for T > 300 °C
Na2O + K2O + CaO <1? (wt.%)111111
Si (apfu) between 2 to 4 apfu111111
Vacancies >0.05 apfu?000000
Lanari et al. [59].NOT APPLICABLENOT APPLICABLENOT APPLICABLENOT APPLICABLENOT APPLICABLENOT APPLICABLE
Uncertainty of 50 °C.
Si < 3 apfu111111
Na2O + K2O + CaO <1? (wt.%)111111
Vacancies (Fe Total = FeOT) > 0.05
(for <0.05 = Thermometer Chl2 not applicable)
T° Chl2 should be not usedT° Chl2 should be not usedT° Chl2 should be not usedT° Chl2 should be not usedT° Chl2 should be not usedT° Chl2 should be not used
Inoue et al. [60,71].Composition Valid Composition Valid Composition Valid Composition Valid Composition Valid Composition Valid
Vacancies apfu < 1111111
Na2O + K2O + CaO < 0.5111111
* Thermometry calculations was carried out using the database of Verdecchia et al. [72].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bobos, I.; de Sá, C.M.; Noronha, F. Mineralogy, Fluid Inclusions, and Oxygen Isotope Geochemistry Signature of Wolframite to Scheelite and Fe,Mn Chlorite Veins from the W, (Cu,Mo) Ore Deposit of Borralha, Portugal. Minerals 2022, 12, 24. https://doi.org/10.3390/min12010024

AMA Style

Bobos I, de Sá CM, Noronha F. Mineralogy, Fluid Inclusions, and Oxygen Isotope Geochemistry Signature of Wolframite to Scheelite and Fe,Mn Chlorite Veins from the W, (Cu,Mo) Ore Deposit of Borralha, Portugal. Minerals. 2022; 12(1):24. https://doi.org/10.3390/min12010024

Chicago/Turabian Style

Bobos, Iuliu, Carlos Marques de Sá, and Fernando Noronha. 2022. "Mineralogy, Fluid Inclusions, and Oxygen Isotope Geochemistry Signature of Wolframite to Scheelite and Fe,Mn Chlorite Veins from the W, (Cu,Mo) Ore Deposit of Borralha, Portugal" Minerals 12, no. 1: 24. https://doi.org/10.3390/min12010024

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