Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-25T04:09:00.344Z Has data issue: false hasContentIssue false

Crystal chemistry of povondraite by single-crystal XRD, EMPA, Mössbauer spectroscopy and FTIR

Published online by Cambridge University Press:  28 November 2022

Ferdinando Bosi*
Affiliation:
Department of Earth Sciences, Sapienza University of Rome, Piazzale A. Moro, 5, I-00185 Rome, Italy
Henrik Skogby
Affiliation:
Department of Geosciences, Swedish Museum of Natural History, Box 50007, SE-10405 Stockholm, Sweden
Guy L. Hovis
Affiliation:
Department of Geology and Environmental Geosciences, Lafayette College, Easton, PA 18042, Pennsylvania, USA
*
*Author for correspondence: Ferdinando Bosi, Email: ferdinando.bosi@uniroma1.it
Rights & Permissions [Opens in a new window]

Abstract

Five povondraite crystals from San Francisco Mine, Villa Tunari, Bolivia, have been structurally and chemically characterised by single-crystal X-ray diffraction and electron microprobe analysis. For the first time, this characterisation is accompanied by Mössbauer spectroscopic and single-crystal infrared spectroscopic data, which show the exclusive presence of Fe3+ at both the octahedrally-coordinated Y and Z sites as well as slight disorder of (OH) and O over the O(1) and O(3) sites.

The data obtained along with those for earlier-studied bosiite and oxy-dravite oxy-tourmalines show a complete substitution series described by the reaction YFe3+3 + ZMg + ZFe3+4YAl2 + YMg + ZAl5 (i.e. Fe3+Al–1) with variation of the structural parameters dominated by Fe3+ (or Al). Povondraite is the tourmaline member having the largest unit-cell parameters due to the larger size of Fe3+ relative to other trivalent cations (V > Cr > Al). In the tourmaline-supergroup minerals, the a and c unit-cell parameters vary from ~15.60 Å to ~16.25 Å and ~7.00 Å to ~7.50 Å, respectively. Their values increase with increasing Fe3+ or decreasing Al. End-member compositions related to the smallest and largest a and c parameters are, respectively, NaAl3Al6(Si3B3O18)(BO3)3(OH)3(OH) (synthetic tourmaline) and NaFe3+3(Fe3+4Mg2)(Si6O18)(BO3)3(OH)3O (povondraite).

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press on behalf of The Mineralogical Society of Great Britain and Ireland

Introduction

Tourmalines are complex borosilicates whose general chemical formula may be written as: XY3Z6T6O18(BO3)3V3W, where X = Na+, K+, Ca2+ and ▫ (= vacancy); Y = Al3+, Fe3+, Cr3+, V3+, Mg2+, Fe2+, Mn2+ and Li+; Z = Al3+, Fe3+, Cr3+, V3+, Mg2+ and Fe2+; T = Si4+ Al3+ and B3+; B = B3+; V = (OH) and O2–; W = (OH), F and O2– (Henry et al., Reference Henry, Novák, Hawthorne, Ertl, Dutrow, Uher and Pezzotta2011). In this representation unitalicised letters X, Y, Z, T and B represent groups of cations hosted at the [9]X, [6]Y, [6]Z, [4]T and [3]B crystallographic sites (letters italicised), whereas V and W represent groups of anions accommodated at the [3]-coordinated O(3) and O(1) crystallographic sites, respectively.

Tourmaline has been studied extensively in terms of crystal structure and crystal chemistry (e.g. Foit, Reference Foit1989; Grice and Ercit, Reference Grice and Ercit1993; Hawthorne and Henry, Reference Hawthorne and Henry1999; Ertl et al., Reference Ertl, Hughes, Pertlik, Foit, Wright, Brandstatter and Marler2002; Novák et al., Reference Novák, Povondra and Selway2004; Bosi and Lucchesi, Reference Bosi and Lucchesi2007; Bosi, Reference Bosi2018; Henry and Dutrow, Reference Henry and Dutrow2011; Henry et al., Reference Henry, Novák, Hawthorne, Ertl, Dutrow, Uher and Pezzotta2011; Cempírek et al., Reference Cempírek, Houzar, Novák, Groat, Selway and Šrein2013; Bačík and Fridrichová, Reference Bačík and Fridrichová2020). Results show that the tourmaline structure is remarkably flexible in a chemical sense, accommodating ions of a wide range of size and charge, which in turn leads to Mg–Fe–Al–Cr–V disorder over the Y and Z sites.

Povondraite, ideally NaFe3+3(Fe3+4Mg2)(Si6O18)(BO3)3(OH)3O, was described by Grice et al. (Reference Grice, Ercit and Hawthorne1993). Until now, however, a chemical and structural study dealing with a statistically significant dataset of Fe3+-dominant oxy-tourmalines is missing. To explore the crystal-chemical aspects and their implications on the tourmaline supergroup, single-crystal structure refinements and electron microprobe data have been collected on five crystals from the type locality for povondraite: San Francisco Mine, Villa Tunari, Alto Chapare, Cochabamba, Bolivia (Walenta and Dunn, Reference Walenta and Dunn1979; Grice et al., Reference Grice, Ercit and Hawthorne1993; Žáček et al., Reference Žáček, Frýda, Petrov and Hyršl2000). All five single crystals were extracted from povondraite sample 110379 from the American Museum of Natural History, New York, USA. This is the first time that Mössbauer spectroscopic and single-crystal infrared spectroscopic data have been presented for this mineral.

Analytical methods and results

General comment

Initially, several crystal fragments of sample 110379 were analysed by electron microprobe; these proved to be chemically inhomogeneous, as reported in Hovis et al. (Reference Hovis, Tribaudino, Altomare and Bosiin press). More recent analyses, reported in the present study, were obtained on different relatively small fragments, which proved to be quite homogeneous as shown by the relatively low standard deviation values of the analysed elements (Table 4).

Single-crystal structure refinement (SREF)

Single-crystal X-ray diffraction (XRD) was undertaken on five crystal fragments of povondraite by mounting on a Bruker KAPPA APEX-II single-crystal diffractometer (Sapienza University of Rome, Earth Sciences Department) equipped with a CCD area detector (6.2 × 6.2 cm active detection area, 512 × 512 pixels) and a graphite-crystal monochromator using MoKα radiation from a fine-focus sealed X-ray tube. The sample-to-detector distance was 4 cm. A total of 1296 exposures (step = 0.4° and time/step = 20 s) covering a full reciprocal sphere were collected using ω and φ scan modes. Final unit-cell parameters were refined using the Bruker AXS SAINT program on reflections with I > 10 σI in the range 5° < 2θ < 75°. The intensity data were processed and corrected for Lorentz polarisation and background effects using the APEX2 software program of Bruker AXS. The data were corrected for absorption using a multi-scan method (SADABS). The absorption correction led to an improvement in R int. No violation of R3m symmetry was detected.

Structure refinement was done using the SHELXL-2013 program (Sheldrick, Reference Sheldrick2015). Starting coordinates were taken from Grice et al. (Reference Grice, Ercit and Hawthorne1993). Variable parameters included scale factor, extinction coefficient, atom coordinates, site-scattering values (for X, Y and Z sites) and atomic-displacement factors. Each structure was refined as a two-component inversion twin. Fully ionised-oxygen scattering factor and neutral-cation scattering factors were used. In detail, the X site was modelled using the Na vs. K scattering factors. The occupancy of the Y site was obtained by considering the presence of Fe vs. Mg, and the Z site with Fe vs. Al. Although there is more Mg than Al content, the latter was preferred to ZMg, having produced slightly better statistical indices and standard uncertainties. The T, B and anion sites were modelled, respectively, with Si, B and O scattering factors and with a fixed occupancy of 1, as refinement with unconstrained occupancies showed no significant deviations from this value. The position of the H atom bonded to the oxygen at the O(3) site in the structure was taken from the difference-Fourier map and incorporated into the refinement model; the O(3)–H(3) bond length was restrained (by DFIX command) to be 0.97 Å with isotropic displacement parameter constrained to be equal to 1.2 times that obtained for the O(3) site. Table 1 lists crystal data, data-collection information, and refinement details. Table 2 gives the fractional atom coordinates and equivalent isotropic-displacement parameters of a typical povondraite crystal (Pov1). Table 3 reports selected bond lengths for all studied crystals. The crystallographic information files showing all structural data have been deposited with the Principal Editor of Mineralogical Magazine and are available as Supplementary material (see below).

Table 1. Single-crystal X-ray diffraction data details for the povondraite crystals studied.*

* Notes: R int = merging residual value; R 1 = discrepancy index, calculated from F-data; wR 2 = weighted discrepancy index, calculated from F 2-data; GoF = goodness of fit; diff. peak = maximum and minimum residual electron density. Space-group type = R3m and Z = 3. Radiation: MoKα = 0.71073 Å. Range for data collection (2θ) = 5°–75°. Data collection temperature = 293 K. Axis = phi-omega, frame width = 0.4°, time per frame = 20 s. Absorption correction method: multi-scan (SADABS, Bruker). Refinement method: full-matrix last-squares on F 2. Structural refinement program: SHELXL-2013 (Sheldrick, Reference Sheldrick2015).

Table 2. Fractional atom coordinates and isotropic (*) or equivalent-isotropic displacement parameters (in Å2) for crystal Pov1.

* Isotropic-displacement parameters (U iso) for H(3) constrained to have a U iso 1.2 times the U eq value of the O(3) oxygen atom, respectively.

Table 3. Selected bond lengths (Å) and cation site occupancy (s.o.) for the povondraite crystals studied.*

* Note: T-site occupancy = Si1.00 and B-site occupancy = B1.00

Electron microprobe analysis (EMPA)

The crystals used for X-ray diffraction refinement were analysed by wavelength dispersive spectrometry (WDS mode) using a Cameca SX50 instrument (CNR-Istituto di Geologia Ambientale e Geoingegneria, Rome, Italy) operating at an accelerating potential of 15 kV and a sample current of 15 nA, with a 10 μm beam diameter. The following standards, X-ray Kα lines and analyser crystals were used: jadeite (Na; TAP), periclase (Mg; TAP), orthoclase (K; PET), rutile (Ti; PET), wollastonite (Si, Ca; PET), metallic Zn and Mn (Zn, Mn; LIF), vanadinite (V; PET), fluorophlogopite (F; TAP), metallic Cr (Cr; PET), corundum (Al; TAP) and magnetite (Fe; LIF). The ‘PAP’ routine was applied (Pouchou and Pichoir Reference Pouchou and Pichoir1991). The results (Table 4) represent mean values of several spot analyses. Vanadium, Cr, Mn and Zn were below detection limits (< 0.03 wt.%).

Table 4. Chemical composition for the povondraite crystals studied.*

* Notes: wt.% = weighted percent; apfu = atoms per formula unit (normalised to 31 anions); epfu = electrons per formula unit; b.d.l. = below detection limit; errors for oxides and fluorine are standard deviations (in brackets).

a Calculated by stoichiometry, (Y+Z+T) = 15.00 apfu.

b Fe oxidation state determined by Mössbauer spectroscopy.

Mössbauer spectroscopy

Several crystal fragments from povondraite sample 110379 were used for 57Fe Mössbauer spectroscopy performed at the Swedish Museum of Natural History, Stockholm, Sweden, using a conventional spectrometer system operated in constant-acceleration mode. Data were collected over 1024 channels; these were folded and calibrated against the spectrum of α-Fe foil. The spectrum (Fig. 1) was fit using the software MossA (Prescher et al., Reference Prescher, McCammon and Dubrowinsky2012) with two absorption doublets consistent with Fe3+ (Table 5). No indications of absorption due to Fe2+ was observed. In line with the site population results from SREF (see below, Table 6), the two doublets can be related to the occurrence of Fe3+ at both the Y and Z sites. However, the low resolution of the two doublets does not allow a definite site assignment.

Fig. 1. Mössbauer spectrum of povondraite obtained at room temperature. The fitted absorption doublets assigned to Fe3+ are indicated in blue. Diamonds denote the measured spectrum and the black curve represents summed fitted doublets.

Table 5. Mössbauer parameters for povondraite obtained at room temperature.*

* δ = centre shift, ΔE Q = quadrupole splitting, FWHM = full width at half-maximum.

Table 6. Site populations (atoms per formula unit) and mean atomic number (man) for the povondraite crystals studied.*

* Notes: All crystals have the B and O(2,4,5,6,7,8) sites fully occupied by B3+ and O2–, respectively; obs = observed, calc = calculated from the site population.

The Mössbauer spectroscopy results are consistent with a synchrotron XANES study that reported 87–100% Fe as Fe3+ (Levy et al., Reference Levy, Henry, Roy and Dutrow2018).

Single-crystal infrared spectroscopy

Polarised Fourier-transform infrared (FTIR) absorption spectra were measured on a 33 μm thick doubly polished single-crystal section oriented parallel to the c-axis. A Bruker Vertex spectrometer attached to a Hyperion 2000 microscope and equipped with a halogen lamp source, CaF2 beamsplitter, ZnSe wiregrid polariser and InSb detector was used to collect spectra in the range 2000–13000 cm–1 at a resolution of 4 cm–1. Spectra recorded in polarised mode parallel to the crystallographic c-axis (E||c) show a weak band at 3440 cm–1, a very intense band around 3550 cm–1, a significant band at 3593 cm–1 and a weak band at 3699 cm–1 (Fig. 2). As observed typically for polarised tourmaline spectra in the (OH) range, the main band is off-scale for the E||c direction due to excessive absorption. Spectra obtained perpendicular to the c-axis show considerably weaker bands.

Fig. 2. Polarised FTIR spectra of povondraite, off-set vertically for clarity. The main band is truncated at ~2 absorbance units in the E||c direction due to excessive absorption. Note the comparatively low intensity of the band at ~3699 cm–1 corresponding to very small (OH) contents at W [≡ the O(1) site]. Sample thickness = 33 μm.

Bands above 3600–3650 cm–1 are normally considered to be due to (OH) at the W position [≡ O(1) site] (e.g. Gonzalez-Carreño et al., Reference Gonzales-Carreño, Fernández and Sanz1988; Bosi et al., Reference Bosi, Skogby, Lazor and Reznitskii2015). For the samples studied, the comparatively weak intensity of the band at 3699 cm–1 indicates low amounts of W(OH). On the basis of studies by Bosi et al. (Reference Bosi, Skogby and Balić-Žunić2016), Watenphul et al. (Reference Watenphul, Burgdorf, Schlüter, Horn, Malcherek and Mihailova2016) and Gatta et al. (Reference Gatta, Bosi, McIntyre and Skogby2014), the main FTIR bands at ~3440 cm–1, ~3550 and ~3593 cm–1 are probably caused by the occurrence of the atomic arrangements 3[Y(Fe3+)Z(Fe3+,Al)Z(Al)]–O(3)(OH)3, {2[Y(Fe3+)Z(Fe3+)Z(Mg)]–[Y(Fe3+)Z(Fe3+)Z(Fe3+)]}–O(3)(OH)3 and 3[Y(Fe3+)Z(Fe3+)Z(Mg)]–O(3)(OH)3, respectively, whereas the band at ~3699 cm–1 may be caused by the arrangements Y(Fe3+MgMg)–O(1)(OH)–X(Na,K).

Determination of number of atoms per formula unit (apfu)

In agreement with the structure-refinement results, the boron content was assumed to be stoichiometric (B3+ = 3.00 apfu). In fact, both the site-scattering results and the bond lengths of B and T are consistent with the B site fully occupied by B3+ and with the T site free of B3+ (e.g. Bosi and Lucchesi, Reference Bosi and Lucchesi2007). Iron oxidation state was determined by Mössbauer spectroscopy, which shows the exclusive presence of Fe3+. In accordance with Pesquera et al. (Reference Pesquera, Gil-Crespo, Torres-Ruiz, Torres-Ruiz and Roda-Robles2016), Li concentrations were considered insignificant as MgO > 2 wt.% in the povondraite crystals studied. The (OH) content and the formula were then calculated by charge balance with the assumption (T + Y + Z) = 15 apfu and 31 anions. The excellent agreement between the number of electrons per formula unit (epfu) derived from EMPA and SREF (within 1 epfu for all studied crystals) supports the stoichiometric assumptions.

Site populations

The povondraite site populations at the X, B, T, O(3) (≡ V) and O(1) (≡ W) sites of crystals Pov1,2,3,4,5 follow the standard site preference suggested for tourmaline (Henry et al., Reference Henry, Novák, Hawthorne, Ertl, Dutrow, Uher and Pezzotta2011) and are coherent with the information from FTIR absorption spectra. In particular, the presence of ~0.10 Al apfu at the T site is consistent with observed <T–O> distances ranging from 1.621–1.625 Å, which are larger than the expected value for <TSi–O> = 1.619(1) Å (Bosi and Lucchesi, Reference Bosi and Lucchesi2007). The Fe3+, Al and Mg site populations at the octahedrally coordinated Y and Z sites were optimised according to the procedure of Bosi et al. (Reference Bosi, Reznitskii, Hålenius and Skogby2017b) and Wright et al. (Reference Wright, Foley and Hughes2000), as well as by fixing the minor elements Ti4+ at the Y site. The resulting site populations are reported in Table 6, which also includes a comparison between the values of observed mean atomic number (as defined by Hawthorne et al., Reference Hawthorne, Ungaretti and Oberti1995) and those calculated from the site populations. The agreement between the refined and calculated values is very good and validates the distribution of cations over the X, Y, Z and T sites in the crystals studied. This site population is also supported by the comparison of weighted bond-valence sums (BVS) and weighted atomic valence (or mean formal charge) calculated from the site populations (Table 7). It is worth noting that the presence of W(OH) at the O(1) site, revealed by FTIR spectra, has been quantified by using the empirical equation W(OH) = {2 – [1.01⋅BVS(O1)] – 0.21 – F} of Bosi (Reference Bosi2013). As a result, O and (OH) are partially disordered over the O(1) and O(3) sites.

Table 7. Weighted bond-valence sum (BVS, in valence units) and weighted atomic valence (WAV) calculated from site population for the povondraite crystals studied.*

* Note: The O(2,4,5,6,7,8) sites are fully occupied by O2–. Bond valence parameters from Gagné and Hawthorne (Reference Gagné and Hawthorne2015).

Discussion

All the crystals studied can by identified as povondraite (Table 6). More specifically, they are consistent with oxy-tourmalines belonging to the alkali group (Henry et al., Reference Henry, Novák, Hawthorne, Ertl, Dutrow, Uher and Pezzotta2011), Na-dominant at the X position and oxy-dominant at the W position with O2– > (F + OH) in the tourmaline general formula. The Y position is dominated by Fe3+ and the Z position requires a double site-occupancy (Fe3+4Mg2) for formula electroneutrality. Collectively, these constituents lead to the povondraite end-member NaFe3+3(Fe3+4Mg2)(Si6O18)(BO3)3(OH)3O.

The five analysed crystals show a substitution series dominated by Fe3+ and Al, which leads to bosiite, ideally NaFe3+3(Al4Mg2)(Si6O18)(BO3)3(OH)3O, by the substitution ZFe3+ZAl, and to oxy-dravite, ideally Na(Al2Mg)(Al5Mg)(Si6O18)(BO3)3(OH)3O, by the substitution YFe3+3 + ZMg ↔ 2YAl2 + YMg + ZAl. As a result, the comprehensive substitution reaction along the povondraite–bosiite–oxy-dravite series is: YFe3+3 + ZMg + ZFe3+4YAl2 + YMg + ZAl5. The latter can be summarised as Fe3+Al–1, as shown in Fig. 3 where the substitution of Fe3+ for Al defines a line having a slope of ~45°. Any deviation from this line may be ascribed to other substitutional mechanisms such as coupled substitutions related to O–(OH) at the O(1) (= W) and O(3) (= V) sites. A similar Fe3+Al–1 substitution is reported in Žáček et al. (Reference Žáček, Frýda, Petrov and Hyršl2000).

Fig. 3. Plot of Fe3+ vs. Al. The dashed black line is a linear regression based on 8 data points, which gives a slope of ~45°. Data show the occurrence of the Fe3+Al–1 substitution in the povondraite–bosiite–oxy-dravite series. Black circles represent povondraite (Pov1-5) from this study, grey circle represents povondraite (Pov) from Grice and Ercit (Reference Grice and Ercit1993), blue circle represents bosiite (Bos) from Ertl et al. (Reference Ertl, Baksheev, Giester, Lengauer, Prokofiev and Zorina2016), red circle represents oxy-dravite (Odrv) from Bosi and Skogby (Reference Bosi and Skogby2013). Abbreviations list according to Warr (Reference Warr2021).

Sodium and K show similar variations, 0.79–0.89 apfu and 0.12–0.24 apfu, respectively, which are strongly correlated with each other (coefficient of determination, r 2 = 0.997) and affect the <X–O> mean bond-length variation (2.729–2.744 Å). In particular, the relatively high K content is related to the increase in Fe3+ (Table 4). As noted by Bačík et al. (Reference Bačík, Uher, Sykora and Lipka2008), the incorporation of a relatively large cation such as K (+ Na) into the povondraite structure should be favoured by the larger unit-cell of Fe3+ – relative to Al-dominant tourmalines such as bosiite or oxy-dravite (see below). This mechanism is different from that involved in maruyamaite (K- and Al-dominant tourmaline) in which the substitution K → Na occurs only under high-pressure conditions (Berryman et al., Reference Berryman, Wunder and Rhede2014; Lussier, et al., Reference Lussier, Ball, Hawthorne, Henry, Shimizu, Ogasawara and Ota2016).

As for the octahedrally coordinated cations, Mg varies from 1.84 to 2.02 apfu and occupies both the Y and Z sites, whereas Al varies from 0.18 to 1.50 apfu and is ordered at the Z site. Ferric iron varies from 5.44 to 7.07 apfu, showing a rather disordered distribution over the Y and Z sites.

Despite the significant YFe3+ variations in the present povondraite crystals, the <Y–O> values are practically constant; the increase in contents of the smaller cation YFe3+ (2.26–2.69 apfu), accompanied by decrease in contents of the larger cation YMg (0.31–0.66 apfu), do not produce any decrease in <Y–O> (2.043–2.045 Å). Therefore, we may infer that the accommodation of Fe at the Y site should produce a <Y–O> expansion that compensates for the differences in size between Fe3+ and Mg2+ substituent ions. This expansion may be shown by the smaller values of bond-valence sum at Y (2.69–2.74 vu) with respect to the weighted atomic valence at Y (2.81–2.90 vu) (Table 7), indicating that the Y-cation is underbonded and bond lengths in the YO6 polyhedron are stretched (Bosi, Reference Bosi2014).

In general, the variation of the structural parameters is dominated by Fe3+ (or Al). No significant correlation occurs between <Y–O> and YFe3+ (Fig. 4a), whereas the <Z–O> variation (1.997–2.017 Å) is positively correlated with ZFe3+ (Fig. 4b) and negatively correlated with ZAl (not shown; r 2 = 0.98). Similarly, the a- and c-parameter are positively related to ZFe3+ (Fig. 5).

Fig. 4. Plot of <Y–O> vs. Fe3+ at the Y site (a) and plot of <Z–O> vs. Fe3+ at the Z site (b). The latter shows a much stronger correlation between the parameters than does the former. The dashed black line is a linear regression (number of data = 8 data). Sources of data as in Fig. 3.

Fig. 5. Plot of a and c unit-cell parameters vs. Fe3+ at the Z site. The dashed black line is a linear regression (number of data = 8 data). Sources of data as in Fig. 3.

The a- and c-parameters show similar variations in the studied crystals, 16.1679(2)–16.2366(3) Å and 7.4122(1)–7.4688(2) Å, respectively, which are positively correlated (r 2 = 0.98). Povondraite has relatively large unit-cell parameters with respect to other tourmalines due to the larger size of Fe3+ compared to other trivalent cations V3+ > Cr3+ > Al3+ (Bosi, Reference Bosi2018). The plot of a against c (Fig. 6) shows the variation of these parameters in the tourmaline-supergroup minerals (a range ~15.60–16.25 Å and c range ~7.00–7.50 Å) and their increase with increasing Fe3+ or decreasing Al. In particular, the smallest a- and c-parameters are those of synthetic Al-B-tourmalines, whose compositions lead to the end-members NaAl3Al6(Si3B3O18)(BO3)3(OH)3(OH) (Schreyer et al., Reference Schreyer, Wodara, Marler, van Aken, Seifert and Robert2000; Marler et al., Reference Marler, Borowski, Wodara and Schreyer2002) and NaAl3Al6(Si4B2O18)(BO3)3(OH)3O (Kutzschbach et al., Reference Kutzschbach, Wunder, Rhede, Koch-Mueller, Ertl, Giester, Heinrich and Franz2016), whereas the largest ones are of povondraite crystal pov5 of the present study.

Fig. 6. Plot of a against c showing the whole variation of the unit-cell parameters in the tourmaline-supergroup minerals. Plot obtained using 326 data sets with structure refinement. In detail, black circles represent povondraite from this study, a grey circle represents povondraite from Grice and Ercit (Reference Grice and Ercit1993), a blue circle represents bosiite from Ertl et al. (Reference Ertl, Baksheev, Giester, Lengauer, Prokofiev and Zorina2016), a red circle represents oxy-dravite from Bosi and Skogby (Reference Bosi and Skogby2013), white circles represent data from literature (see figure 3 of Bosi, Reference Bosi2018), grey squares represent samples from Marler et al. (Reference Marler, Borowski, Wodara and Schreyer2002), a grey triangle symbol represents samples from Kutzschbach et al. (Reference Kutzschbach, Wunder, Rhede, Koch-Mueller, Ertl, Giester, Heinrich and Franz2016), and grey diamonds represent the ideal value from Epprecht (Reference Epprecht1953). Text symbols: Elb = elbaite, Drv = dravite, Srl = schorl. Text symbols highlighted in pale green refer to holotypes data of Y(Fe2+,Fe3+), Cr- and V-dominant tourmalines: Ovdrv = oxy-vanadium-dravite (Bosi et al., Reference Bosi, Reznitskii and Sklyarov2013), Ocdrv = oxy-chromium-dravite (representing also vanadio-chromium-oxy-dravite; Bosi et al., Reference Bosi, Reznitskii and Skogby2012, Reference Bosi, Reznitskii, Skogby and Hålenius2014a), Capov = chromo-alumino-povondraite (Reznitskii et al., Reference Reznitskii, Clark, Hawthorne, Grice, Skogby, Hålenius and Bosi2014), Vodv = Vanadio-oxy-dravite (Bosi et al., Reference Bosi, Skogby, Reznitskii and Hålenius2014b), Lcc = lucchesiite (Bosi et al., Reference Bosi, Skogby, Ciriotti, Gadas, Novák, Cempírek, Všianský and Filip2017a), Fbu = fluor-buergerite (Donnay et al., Reference Donnay, Ingamells and Mason1966), Adc = Adachiite (Nishio-Hamane et al., Reference Nishio-Hamane, Minakawa, Yamaura, Oyama, Ohnishi and Shimobayashi2014), Ofoi = oxy-foitite (Bosi et al., Reference Bosi, Skogby and Hålenius2017c), Foi = foitite (MacDonald et al., Reference MacDonald, Hawthorne and Grice1993), Osrl = oxy-schorl (Bačík et al., Reference Bačík, Cempírek, Uher, Novák, Ozdín, Filip, Škoda, Breiter, Klementová and Ďuďa2013). Abbreviations list according to Warr (Reference Warr2021).

Acknowledgements

The authors are grateful to George E. Harlow (American Museum of Natural History, New York, USA) for kindly furnishing povondraite sample 110379 to G.H. (Hovis et al., Reference Hovis, Tribaudino, Altomare and Bosiin press) along with well wishes for the further study reported here. Chemical analyses were done with the kind assistance of Beatrice Celata to whom the authors express their gratitude. Comments by the Structural Editor (P. Leverett) and reviewers Peter Bačík and Darrell Henry are very much appreciated. F.B. acknowledges funding by Sapienza University of Rome (Prog. Università 2020) and by the Italian Ministry of Education (MIUR)–PRIN 2020, ref. 2020WYL4NY.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1180/mgm.2022.132

Competing interests

The authors declare none.

Footnotes

Associate Editor: Charles A. Geiger

References

Bačík, P. and Fridrichová, G. (2020) Cation partitioning among crystallographic sites based on bond-length constraints in tourmaline-supergroup minerals. American Mineralogist, 106, 851861, https://doi.org/10.2138/am-2021-7804.CrossRefGoogle Scholar
Bačík, P., Uher, P., Sykora, M. and Lipka, J. (2008) Low-Al tourmalines of the schorl – dravite – povondraite series in redeposited tourmalinites from the Western Carpathians, Slovakia. The Canadian Mineralogist, 46, 11171129.CrossRefGoogle Scholar
Bačík, P., Cempírek, J., Uher, P., Novák, M., Ozdín, D., Filip, J., Škoda, R., Breiter, K., Klementová, M. and Ďuďa, R. (2013) Oxy-schorl, Na(Fe22+Al) Al6Si6O18(BO3)3(OH)3O, a new mineral from Zlatá Idka, Slovak Republic and Přibyslavice, Czech Republic. American Mineralogist, 98, 485492.CrossRefGoogle Scholar
Berryman, E., Wunder, B. and Rhede, D. (2014) Synthesis of K-dominant tourmaline. American Mineralogist, 99, 539542.CrossRefGoogle Scholar
Bosi, F. (2013) Bond-valence constraints around the O1 site of tourmaline. Mineralogical Magazine, 77, 343351.CrossRefGoogle Scholar
Bosi, F. (2014) Mean bond-length variation in crystal structures: A bond-valence approach. Acta Crystallographica, B70, 697704.Google Scholar
Bosi, F. (2018) Tourmaline crystal chemistry. American Mineralogist, 103, 298306.CrossRefGoogle Scholar
Bosi, F. and Lucchesi, S. (2007) Crystal chemical relationships in the tourmaline group: structural constraints on chemical variability. American Mineralogist, 92, 10541063.CrossRefGoogle Scholar
Bosi, F. and Skogby, H. (2013) Oxy-dravite, Na(Al2Mg)(Al5Mg)(Si6O18)(BO3)3(OH)3O, a new mineral species of the tourmaline supergroup. American Mineralogist, 98, 14421448.CrossRefGoogle Scholar
Bosi, F., Reznitskii, L. and Skogby, H. (2012) Oxy-chromium-dravite, NaCr3(Cr4Mg2)(Si6O18)(BO3)3(OH)3O, a new mineral species of the tourmaline supergroup. American Mineralogist, 97, 20242030.CrossRefGoogle Scholar
Bosi, F., Reznitskii, L. and Sklyarov, E.V. (2013) Oxy-vanadium-dravite, NaV3(V4Mg2)(Si6O18)(BO3)3(OH)3O: crystal structure and redefinition of the “vanadium-dravite” tourmaline. American Mineralogist, 98, 501505.CrossRefGoogle Scholar
Bosi, F., Reznitskii, L., Skogby, H. and Hålenius, U. (2014a) Vanadio-oxy-chromium-dravite, NaV3(Cr4Mg2)(Si6O18)(BO3)3(OH)3O, a new mineral species of the tourmaline supergroup. American Mineralogist, 99, 11551162.CrossRefGoogle Scholar
Bosi, F., Skogby, H., Reznitskii, L. and Hålenius, U. (2014b) Vanadio-oxy-dravite, NaV3(Al4Mg2)(Si6O18)(BO3)3(OH)3O, a new mineral species of the tourmaline supergroup. American Mineralogist, 99, 218224.CrossRefGoogle Scholar
Bosi, F., Skogby, H., Lazor, P. and Reznitskii, L. (2015) Atomic arrangements around the O3 site in Al- and Cr-rich oxy-tourmalines: a combined EMP, SREF, FTIR and Raman study. Physics and Chemistry of Minerals, 42, 441453.CrossRefGoogle Scholar
Bosi, F., Skogby, H. and Balić-Žunić, T. (2016) Thermal stability of extended clusters in dravite: a combined EMP, SREF and FTIR study. Physics and Chemistry of Minerals, 43, 395407.CrossRefGoogle Scholar
Bosi, F., Skogby, H., Ciriotti, M.E., Gadas, P., Novák, M., Cempírek, J., Všianský, D. and Filip, J. (2017a) Lucchesiite, CaFe2+3Al6(Si6O18)(BO3)3(OH)3O, a new mineral species of the tourmaline supergroup. Mineralogical Magazine, 81, 114.CrossRefGoogle Scholar
Bosi, F., Reznitskii, L., Hålenius, U. and Skogby, H. (2017b) Crystal chemistry of Al–V–Cr oxy-tourmalines from Sludyanka complex, Lake Baikal, Russia. European Journal of Mineralogy, 29, 457472.CrossRefGoogle Scholar
Bosi, F., Skogby, H. and Hålenius, U. (2017c) Oxy-foitite, ▫(Fe2+Al2)Al6(Si6O18)(BO3)3(OH)3O, a new mineral species of the tourmaline supergroup. European Journal of Mineralogy, 29, 889896.CrossRefGoogle Scholar
Cempírek, J., Houzar, S., Novák, M., Groat, L.A., Selway, J.B. and Šrein, V. (2013) Crystal structure and compositional evolution of vanadium-rich oxy-dravite from graphite quartzite at Bítovánky, Czech Republic. Journal of Geosciences, 58, 149162.CrossRefGoogle Scholar
Donnay, G., Ingamells, C.O. and Mason, B.H. (1966) Buergerite, a new species of tourmaline. American Mineralogist, 50, 198199.Google Scholar
Epprecht, W. (1953) Die Gitterkonstanten der Turmaline. Schweizerische Mineralogische und Petrographische Mitteilungen, 33, 481505.Google Scholar
Ertl, A., Hughes, J.M., Pertlik, F., Foit, F.F. Jr., Wright, S.E., Brandstatter, F. and Marler, B. (2002) Polyhedron distortions in tourmaline. The Canadian Mineralogist, 40, 153162.CrossRefGoogle Scholar
Ertl, A., Baksheev, I.A., Giester, G., Lengauer, C.L., Prokofiev, V.Yu. and Zorina, L.D. (2016) Bosiite, NaFe3+3(Al4Mg2)(Si6O18)(BO3)3(OH)3O, a new ferric member of the tourmaline supergroup from the Darasun gold deposit, Transbaikalia, Russia. European Journal of Mineralogy, 28, 581591.CrossRefGoogle Scholar
Foit, F.F. Jr. (1989) Crystal chemistry of alkali-deficient schorl and tourmaline structural relationships. American Mineralogist, 74, 422431.Google Scholar
Gagné, O.C. and Hawthorne, F.C. (2015) Comprehensive derivation of bond- valence parameters for ion pairs involving oxygen. Acta Crystallographica, B71, 562578.Google Scholar
Gatta, G.D., Bosi, F., McIntyre, G.J. and Skogby, H. (2014) First accurate location of two proton sites in tourmaline: A single-crystal neutron diffraction study of oxy-dravite. Mineralogical Magazine, 78, 681692.CrossRefGoogle Scholar
Gonzales-Carreño, T., Fernández, M. and Sanz, J. (1988) Infrared and electron microprobe analysis of tourmaline. Physics and Chemistry of Minerals, 15, 452460.CrossRefGoogle Scholar
Grice, J.D. and Ercit, T.S. (1993) Ordering of Fe and Mg in the tourmaline crystal structure: the correct formula. Neues Jahrbuch für Mineralogie, Abhandlungen, 165, 245266.Google Scholar
Grice, J.D., Ercit, T.S. and Hawthorne, F.C. (1993) Povondraite, a redefinition of the tourmaline ferridravite. American Mineralogist, 78, 433436.Google Scholar
Hawthorne, F.C. and Henry, D. (1999) Classification of the minerals of the tourmaline group. European Journal of Mineralogy, 11, 201215.CrossRefGoogle Scholar
Hawthorne, F.C., Ungaretti, L. and Oberti, R. (1995) Site populations in minerals: terminology and presentation of results of crystal-structure refinement. The Canadian Mineralogist, 33, 907911.Google Scholar
Henry, D.J. and Dutrow, B.L. (2011) The incorporation of fluorine in tourmaline: Internal crystallographic controls or external environmental influences? Canadian Mineralogist, 49, 4156.CrossRefGoogle Scholar
Henry, D.J., Novák, M., Hawthorne, F.C., Ertl, A., Dutrow, B., Uher, P. and Pezzotta, F. (2011) Nomenclature of the tourmaline supergroup minerals. American Mineralogist, 96, 895913.CrossRefGoogle Scholar
Hovis, G.L., Tribaudino, M., Altomare, C. and Bosi, F. (in press) Thermal expansion of minerals in the tourmaline supergroup. American Mineralogist, doi.org/10.2138/am-2022-8580.Google Scholar
Kutzschbach, M., Wunder, B., Rhede, D., Koch-Mueller, M., Ertl, A., Giester, G., Heinrich, W. and Franz, G. (2016) Tetrahedral boron in natural and synthetic HP/UHP tourmaline: Evidence from Raman spectroscopy, EMPA, and single crystal XRD. American Mineralogist, 101, 93104.CrossRefGoogle Scholar
Levy, E.A., Henry, D.J., Roy, A. and Dutrow, B.L. (2018) Determination of ferrous-ferric iron contents in tourmaline using synchrotron-based XANES. Journal of Geosciences, 63, 167174.CrossRefGoogle Scholar
Lussier, A.J., Ball, N.A., Hawthorne, F.C., Henry, D.J., Shimizu, R., Ogasawara, Y. and Ota, T. (2016) Maruyamaite, K(MgAl2)(Al5Mg)Si6O18(BO3)3(OH)3O, from the ultrahigh-pressure Kokchetav massif, northern Kazakhstan: Description and crystal structure. American Mineralogist, 101, 355361.CrossRefGoogle Scholar
MacDonald, D.J., Hawthorne, F.C. and Grice, J.D. (1993) Foitite, ▫[Fe2+2(Al,Fe3+)]Al6Si6O18(BO3)3(OH)4, a new alkali-deficient tourmaline: Description and crystal structure. American Mineralogist, 78, 12991303.Google Scholar
Marler, B., Borowski, M., Wodara, U. and Schreyer, W. (2002) Synthetic tourmaline (olenite) with excess boron replacing silicon in the tetrahedral site: II. Structure analysis. European Journal of Mineralogy, 14, 763771.CrossRefGoogle Scholar
Nishio-Hamane, D., Minakawa, T., Yamaura, J., Oyama, T., Ohnishi, M. and Shimobayashi, N. (2014) Adachiite, a Si–poor member of the tourmaline supergroup from the Kiura mine, Oita Prefecture, Japan. Journal of Mineralogical and Petrological Sciences, 109, 7478.CrossRefGoogle Scholar
Novák, M., Povondra, P. and Selway, J.B. (2004) Schorl-oxy-schorl to dravite-oxydravite tourmaline from granitic pegmatites; examples from the Moldanubicum, Czech Republic. European Journal of Mineralogy, 16, 323333.CrossRefGoogle Scholar
Pesquera, A., Gil-Crespo, P.P Torres-Ruiz, F., Torres-Ruiz, J. and Roda-Robles, E. (2016) A multiple regression method for estimating Li in tourmaline from electron microprobe analyses. Mineralogical Magazine, 80, 11291133.CrossRefGoogle Scholar
Pouchou, J.L. and Pichoir, F. (1991) Quantitative analysis of homogeneous or stratified microvolumes applying the model “PAP.” Pp. 31–75 in: Electron Probe Quantitation (K.F.J. Heinrich and D.E. Newbury, editors). Plenum, New York.CrossRefGoogle Scholar
Prescher, C., McCammon, C. and Dubrowinsky, L. (2012) MossA: a program for analyzing energy-domain Mössbauer spectra from conventional and synchrotron sources. Journal of Applied Crystallography, 45, 329331.CrossRefGoogle Scholar
Reznitskii, L., Clark, C.M., Hawthorne, F.C., Grice, J.D., Skogby, H., Hålenius, U. and Bosi, F. (2014) Chromo-alumino-povondraite, NaCr3(Al4Mg2)(Si6O18)(BO3)3(OH)3O, a new mineral species of the tourmaline supergroup. American Mineralogist, 99, 17671773.CrossRefGoogle Scholar
Schreyer, W., Wodara, U., Marler, B., van Aken, P.A., Seifert, F. and Robert, J.L. (2000) Synthetic tourmaline (olenite) with excess boron replacing silicon in the tetrahedral site: I. Synthesis conditions, chemical and spectroscopic evidence. European Journal of Mineralogy, 12, 529541.CrossRefGoogle Scholar
Sheldrick, G.M. (2015) Crystal structure refinement with SHELXL. Acta Crystallographica, C71, 38.Google Scholar
Walenta, K. and Dunn, P.J. (1979) Ferridravite, a new mineral of the tourmaline group from Bolivia. American Mineralogist, 64, 945948.Google Scholar
Warr, L.N. (2021) IMA–CNMNC approved mineral symbols. Mineralogical Magazine, 85, 291320.CrossRefGoogle Scholar
Watenphul, A., Burgdorf, M., Schlüter, J., Horn, I., Malcherek, T. and Mihailova, B. (2016) Exploring the potential of Raman spectroscopy for crystallo- chemical analyses of complex hydrous silicates: II. Tourmalines. American Mineralogist, 101, 970985.CrossRefGoogle Scholar
Wright, S.E., Foley, J.A. and Hughes, J.M. (2000) Optimization of site occupancies in minerals using quadratic programming. American Mineralogist, 85, 524531.CrossRefGoogle Scholar
Žáček, V., Frýda, J., Petrov, A. and Hyršl, J. (2000) Tourmalines of the povondraite—(oxy)dravite series from the caps rocks of meta-evaporite in Alto Chapare, Cochamba, Bolivia. Journal of the Czech Geological Society, 45, 312.Google Scholar
Figure 0

Table 1. Single-crystal X-ray diffraction data details for the povondraite crystals studied.*

Figure 1

Table 2. Fractional atom coordinates and isotropic (*) or equivalent-isotropic displacement parameters (in Å2) for crystal Pov1.

Figure 2

Table 3. Selected bond lengths (Å) and cation site occupancy (s.o.) for the povondraite crystals studied.*

Figure 3

Table 4. Chemical composition for the povondraite crystals studied.*

Figure 4

Fig. 1. Mössbauer spectrum of povondraite obtained at room temperature. The fitted absorption doublets assigned to Fe3+ are indicated in blue. Diamonds denote the measured spectrum and the black curve represents summed fitted doublets.

Figure 5

Table 5. Mössbauer parameters for povondraite obtained at room temperature.*

Figure 6

Table 6. Site populations (atoms per formula unit) and mean atomic number (man) for the povondraite crystals studied.*

Figure 7

Fig. 2. Polarised FTIR spectra of povondraite, off-set vertically for clarity. The main band is truncated at ~2 absorbance units in the E||c direction due to excessive absorption. Note the comparatively low intensity of the band at ~3699 cm–1 corresponding to very small (OH) contents at W [≡ the O(1) site]. Sample thickness = 33 μm.

Figure 8

Table 7. Weighted bond-valence sum (BVS, in valence units) and weighted atomic valence (WAV) calculated from site population for the povondraite crystals studied.*

Figure 9

Fig. 3. Plot of Fe3+vs. Al. The dashed black line is a linear regression based on 8 data points, which gives a slope of ~45°. Data show the occurrence of the Fe3+Al–1 substitution in the povondraite–bosiite–oxy-dravite series. Black circles represent povondraite (Pov1-5) from this study, grey circle represents povondraite (Pov) from Grice and Ercit (1993), blue circle represents bosiite (Bos) from Ertl et al. (2016), red circle represents oxy-dravite (Odrv) from Bosi and Skogby (2013). Abbreviations list according to Warr (2021).

Figure 10

Fig. 4. Plot of <Y–O> vs. Fe3+ at the Y site (a) and plot of <Z–O> vs. Fe3+ at the Z site (b). The latter shows a much stronger correlation between the parameters than does the former. The dashed black line is a linear regression (number of data = 8 data). Sources of data as in Fig. 3.

Figure 11

Fig. 5. Plot of a and c unit-cell parameters vs. Fe3+ at the Z site. The dashed black line is a linear regression (number of data = 8 data). Sources of data as in Fig. 3.

Figure 12

Fig. 6. Plot of a against c showing the whole variation of the unit-cell parameters in the tourmaline-supergroup minerals. Plot obtained using 326 data sets with structure refinement. In detail, black circles represent povondraite from this study, a grey circle represents povondraite from Grice and Ercit (1993), a blue circle represents bosiite from Ertl et al. (2016), a red circle represents oxy-dravite from Bosi and Skogby (2013), white circles represent data from literature (see figure 3 of Bosi, 2018), grey squares represent samples from Marler et al. (2002), a grey triangle symbol represents samples from Kutzschbach et al. (2016), and grey diamonds represent the ideal value from Epprecht (1953). Text symbols: Elb = elbaite, Drv = dravite, Srl = schorl. Text symbols highlighted in pale green refer to holotypes data of Y(Fe2+,Fe3+), Cr- and V-dominant tourmalines: Ovdrv = oxy-vanadium-dravite (Bosi et al., 2013), Ocdrv = oxy-chromium-dravite (representing also vanadio-chromium-oxy-dravite; Bosi et al., 2012, 2014a), Capov = chromo-alumino-povondraite (Reznitskii et al., 2014), Vodv = Vanadio-oxy-dravite (Bosi et al., 2014b), Lcc = lucchesiite (Bosi et al., 2017a), Fbu = fluor-buergerite (Donnay et al., 1966), Adc = Adachiite (Nishio-Hamane et al., 2014), Ofoi = oxy-foitite (Bosi et al., 2017c), Foi = foitite (MacDonald et al., 1993), Osrl = oxy-schorl (Bačík et al., 2013). Abbreviations list according to Warr (2021).

Supplementary material: File

Bosi et al. supplementary material

Bosi et al. supplementary material

Download Bosi et al. supplementary material(File)
File 949.5 KB