Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-26T18:00:04.187Z Has data issue: false hasContentIssue false

Ferriandrosite-(Ce), a new member of the epidote supergroup from Betliar, Slovakia

Published online by Cambridge University Press:  07 August 2023

Martin Števko*
Affiliation:
Earth Science Institute v.v.i., Slovak Academy of Sciences, Dúbravská cesta 9, P.O. BOX 106, 840 05 Bratislava, Slovakia Department of Mineralogy and Petrology, National Museum, Cirkusová 1740, 193 00 Prague 9 – Horní Počernice, Czech Republic
Pavol Myšľan
Affiliation:
Earth Science Institute v.v.i., Slovak Academy of Sciences, Dúbravská cesta 9, P.O. BOX 106, 840 05 Bratislava, Slovakia Department of Mineralogy, Petrology and Economic Geology, Faculty of Natural Sciences, Comenius University, Ilkovičova 6, 842 15, Bratislava, Slovakia
Cristian Biagioni
Affiliation:
Dipartimento di Scienze della Terra, Università di Pisa, Via Santa Maria, 53, I-56126 Pisa, Italy Centro per l'Integrazione della Strumentazione Scientifica dell'Università di Pisa, Pisa, Italy
Daniela Mauro
Affiliation:
Dipartimento di Scienze della Terra, Università di Pisa, Via Santa Maria, 53, I-56126 Pisa, Italy
Tomáš Mikuš
Affiliation:
Earth Science Institute v.v.i., Slovak Academy of Sciences, Ďumbierska 1, 974 11 Banská Bystrica, Slovakia
*
Corresponding author: Martin Števko; Email: msminerals@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

A new member of the epidote supergroup, ferriandrosite-(Ce), ideally MnCeFe3+AlMn2+(Si2O7)(SiO4)O(OH), was found at the Július manganese ore occurrence near Betliar, Rožňava Co., Košice Region, Slovakia. It occurs as subhedral grains and polycrystalline aggregates, up to 0.3 mm in size, enclosed in pyroxmangite. Other associated minerals are spessartine, rhodochrosite, quartz, baryte and pyrosmalite-(Mn). Ferriandrosite-(Ce) is dark brown, with a light brown streak and vitreous lustre. The Mohs hardness is ~6½ to 7 and tenacity is brittle with no observable cleavage or fracture. The calculated density is 4.321 g⋅cm–3. Ferriandrosite-(Ce) is optically biaxial (+), with weak pleochroism, high surface relief and the mean calculated refractive index is 1.832. The empirical structural formula of ferriandrosite-(Ce), based on 13 anions per formula unit, is A1(Mn2+0.63Ca0.35Ce0.02)Σ1.00A2(Ce0.53La0.27Nd0.14Pr0.05REE*0.01)Σ1.00M1(Fe3+0.41Al0.12V3+0.01Mg0.40Ti0.05)Σ0.99M2Al1.00M3(Mn2+0.75Fe2+0.22Mg0.03)Σ1.00T1–3Si3.00O11O4(O0.67F0.33)(OH), where REE* are minor rare earth elements. Ferriandrosite-(Ce) is monoclinic, space group P21/m, a = 8.8483(4) Å, b = 5.7307(3) Å c = 10.0314(5) Å, β = 113.3659(15)°, V = 466.95(4) Å3 and Z = 2. The crystal structure of ferriandrosite-(Ce) was refined to a final R1 = 0.0210 for 1910 reflections with Fo > 4σ(Fo) and 127 refined parameters. Structural features of ferriandrosite-(Ce) are discussed and compared with other members of the androsite-series.

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), 2023. Published by Cambridge University Press on behalf of The Mineralogical Society of the United Kingdom and Ireland

Introduction

The epidote supergroup is composed of mixed-anion silicates with the general formula A 2M 3[T 2O7][TO4](O,F)(OH,O) (Armbruster et al., Reference Armbruster, Bonazzi, Akasaka, Bermanec, Chopin, Gieré, Heuss-Assbichler, Liebscher, Menchetti, Pan and Pasero2006). Currently, this supergroup is represented by 33 monoclinic species, divided into the allanite group (17 species), clinozoisite group (11), and dollaseite group (4); åskagenite-(Nd) is an additional member of this supergroup with O dominant over (OH) at the O10 site. Members of the allanite group are phases rich in rare earth elements (REE), with the following valences at key sites: A1 = M2+, A2 = M3+, M1 = M3+, M2 = M3+, M3 = M2+, O4 = O2–, and O10 = (OH) (Armbruster et al., Reference Armbruster, Bonazzi, Akasaka, Bermanec, Chopin, Gieré, Heuss-Assbichler, Liebscher, Menchetti, Pan and Pasero2006).

During examination of the mineral assemblage of the Július manganese ore occurrence, near Betliar, in Slovakia, a new allanite-group mineral was identified, corresponding to the ideal composition MnCeFe3+AlMn2+(Si2O7)(SiO4)O(OH). Such a composition was reported previously by Girtler et al. (Reference Girtler, Tropper and Hauzenberger2013) and Kolitsch et al. (Reference Kolitsch, Schachinger, Auer and Walter2021) from some occurrences of metamorphosed Mn ores in Austria but no formal proposal for its approval was submitted to the Commission on New Minerals, Nomenclature and Classification of the International Mineralogical Association (IMA-CNMNC). On the basis of the Slovak occurrence, this mineral has been named ferriandrosite-(Ce), in accordance with the existing nomenclature scheme for the epidote-supergroup minerals (Armbruster et al., Reference Armbruster, Bonazzi, Akasaka, Bermanec, Chopin, Gieré, Heuss-Assbichler, Liebscher, Menchetti, Pan and Pasero2006). The root-name ‘androsite’ was first used by Bonazzi et al. (Reference Bonazzi, Menchetti and Reinecke1996) for a REE-rich member from the Andros Island, Greece, having Mn2+ as the dominant cation at the A1 and M3 sites and the M1 site occupied by Mn3+; the prefix ‘ferri’ indicates the dominance of Fe3+ at the M1 site for the specimen from the Július manganese ore occurrence. The suffix, indicating the dominant cation at the A2 site, agrees with the nomenclature for REE mineral species (Bayliss and Levinson, Reference Bayliss and Levinson1988). The new mineral and its name have been approved by the IMA-CNMNC (IMA2023–022, Števko et al., Reference Števko, Myšľan, Biagioni, Mauro and Mikuš2023). The approved mineral symbol of ferriandrosite-(Ce) is Fea-Ce. The holotype specimen of ferriandrosite-(Ce) (polished section Be-5) is deposited in the collections of the Department of Mineralogy and Petrology, National Museum in Prague, Cirkusová 1740, 19300 Praha 9, Czech Republic under the catalogue number P1P 2/2023 and the grain used for the single-crystal X-ray diffraction study is kept in the collections of the Museo di Storia Naturale of the Università di Pisa, Via Roma 79, Calci (PI), Italy under catalogue number 20063.

Occurrence and physical properties

Occurrence

Ferriandrosite-(Ce) was found at the Július metamorphosed manganese ore occurrence, which is located on the NE slopes of the Turecká hill (953 m a.s.l.), ~1.4 km WSW of the Betliar village, Rožňava Co., Košice Region, Slovak Republic. Samples with ferriandrosite-(Ce) were collected on September 17, 2019, by one of the authors (MS) from the dump of the main exploration adit (48°41'53.3“N, 20°29'20.1”E).

The Július manganese ore occurrence is situated in the Spišsko-gemerské rudohorie Mts. Small bodies and lenses of metamorphosed carbonate–silicate manganese ore mineralisation are known from several localities in the Spišsko-gemerské rudohorie Mts. (e.g. Čučma-Čierna baňa, Smolník-Hekerová, Betliar-Július), and are hosted in Lower Palaeozoic black phyllites and lydites of the Gelnica Group (Gemeric Unit), metamorphosed in greenschist-facies conditions (Kantor, Reference Kantor1954; Bajaník et al., Reference Bajaník, Ivanička, Mello, Pristaš, Reichwalder, Snopko, Vozár and Vozárová1983; Faryad, Reference Faryad1994; Grecula et al., Reference Grecula, Abonyi, Abonyiová, Antaš, Bartalský, Bartalský, Dianiška, Ďuďa, Gargulák, Gazdačko, Hudáček, Kobulský, Lörincz, Macko, Návesňák, Németh, Novotný, Radvanec, Rojkovič, Rozložník, Varček and Zlocha1995; Rojkovič, Reference Rojkovič2001; Peterec and Ďuďa, Reference Peterec and Ďuďa2003). According to Rojkovič (Reference Rojkovič2001), mineral associations composed of Mn carbonates and silicates in the Spišsko-gemerské rudohorie Mts., mainly represented by rhodochrosite, rhodonite, pyroxmangite, spessartine, tephroite and magnetite, were formed during Variscan metamorphism of sedimentary–diagenetic manganese proto-ores. Subsequent metamorphic–hydrothermal veinlets with quartz, calcite and sulfides represent younger mineralisation stages, probably related to the Alpine orogeny. Ore mineralisation at the Július occurrence was explored in a small scale for manganese and iron in the second half of the 19th century (Maderspach, Reference Maderspach1875). In contrast with the nearby and very similar Čučma-Čierna baňa or Smolník-Hekerová manganese deposit, the mineralogy of the small Július manganese ore occurrence was never studied in detail. Maderspach (Reference Maderspach1875) briefly mentioned rhodonite, rhodochrosite, magnetite, pyrite, ‘wad’ and ‘psilomelane’ as principal ore minerals. Ferriandrosite-(Ce) was discovered in a large block of rhodochrosite–spessartine–pyroxmangite–magnetite ore, cut by abundant younger veinlets consisting mainly of coarse crystalline pyroxmangite and spessartine, with minor amounts of quartz, rhodochrosite, baryte and pyrosmalite-(Mn). It occurs rarely as aggregates embedded in pyroxmangite–spessartine veinlets together with rhodochrosite, quartz, baryte and pyrosmalite-(Mn). Its origin is probably due to an Alpine metamorphic–hydrothermal event, favouring the remobilisation of Mn and REE in the Július manganese ore deposit.

Physical and optical properties

Ferriandrosite-(Ce) occurs as subhedral grains and polycrystalline aggregates up to 0.3 mm in size (Figs 1, 2), enclosed in pyroxmangite–spessartine matrix. Ferriandrosite-(Ce) is dark brown, with a light-brown streak and vitreous lustre and it is non-fluorescent in shortwave and longwave ultraviolet light. The Mohs hardness is estimated at ~6½ to 7 based on analogy to other members of the allanite group. It is brittle, with no observable cleavage or fracture. A density of 4.321 g⋅cm–3 was calculated using the unit-cell volume refined from the single-crystal X-ray diffraction data and empirical chemical formula.

Figure 1. Dark brown aggregate of ferriandrosite-(Ce) enclosed in pyroxmangite (light pink)–spessartine (light yellow) matrix. Field of view is 2.06 mm. Photo L. Hrdlovič. Holotype material (P1P 2/2023, polished section Be-5).

Figure 2. Aggregates and subhedral grains of ferriandrosite-(Ce) (white) enclosed in pyroxmangite (light grey) associated with euhedral crystals of spessartine (dark grey). The red box indicates the area where the grain used for single-crystal X-ray diffraction was picked up. Back-scattered electron (BSE) photo T. Mikuš. Holotype material (P1P 2/2023, polished section Be-5).

Ferriandrosite-(Ce) is optically biaxial (+), with weak pleochroism and high surface relief. Other optical properties were not measured because of relatively strong compositional zoning of the studied aggregates and the presence of intimate intergrowths between the two distinct members of the epidote supergroup. In agreement with Armbruster et al. (Reference Armbruster, Bonazzi, Akasaka, Bermanec, Chopin, Gieré, Heuss-Assbichler, Liebscher, Menchetti, Pan and Pasero2006), it is recognised that optical data can be ambiguous and have some limitations in identifying some chemically complex minerals like those belonging to the allanite group. The mean refractive index of ferriandrosite-(Ce), obtained from the Gladstone–Dale relationship (Mandarino, Reference Mandarino1979, Reference Mandarino1981) using ideal end-member formula and calculated density is 1.832.

Chemical composition

Quantitative chemical (wavelength dispersive spectroscopy) analyses of ferriandrosite-(Ce) were carried out using a JEOL-JXA850F electron microprobe (Earth Science Institute, Slovak Academy of Sciences, Banská Bystrica, Slovakia). The following conditions were applied: accelerating voltage 15 kV, probe current 20 nA, counting time 20s on peak and 10s for background. The diameter of the electron beam ranged from 2–4 μm; ZAF correction was used. The following standards, X-ray lines and crystals were used: diopside (CaKα, PETL and MgKα, TAP); UO2 (UMβ, PETL); orthoclase (KKα, PETL and SiKα, TAP); thorianite (ThMα, PETL); tugtupite (ClKα, PETL); crocoite (PbMβ, PETL); YPO4 (YLα, PETL); fluorite (FKα, LDE1); albite (NaKα, TAP and AlKα, TAP); LaPO4 (LaLα, LIFH); CePO4 (CeLα, LIFH); NdPO4 (NdLα, LIFH); PrPO4 (PrLβ, LIFH); EuPO4 (EuLα, LIFH); SmPO4 (SmLβ, LIFH); TbPO4 (TbLα, LIFH); GdPO4 (GdLβ, LIFH); ErPO4 (ErLα, LIFH); TmPO4 (TmLα, LIFH); DyPO4 (DyLβ, LIFH); YbPO4 (YbLα, LIFH); HoPO4 (HoLβ, LIFH); LuPO4 (LuLα, LIFH); hematite (FeKα, LIF); rhodonite (MnKα, LIF); Cr2O3 (CrKα, LIF); ScVO4 (VKα, LIF); and rutile (TiKα, LIF). The detection limit (1σ) of every element ranged from 61–881 ppm. REE interferences were solved by overlap corrections as well as F / Fe and F / Ce X-ray line coincidences.

As shown by back-scattered electron images (Fig. 3), the sample studied is chemically zoned. Four different chemical domains were identified. One of them corresponds to the chemical domain (‘domain A’) characterised through single-crystal X-ray diffraction and its chemical data are given in Table 1 (average of 4 spot analyses). The content of Fe2O3 and FeO and the amount of H2O were recalculated in order to achieve 8 (A+M+T) atoms per formula unit (pfu). If the full oxidation of Fe is not sufficient, then Mn is also partially oxidised.

Figure 3. Back-scattered electron (BSE) image showing the chemical inhomogeneity of the material studied. Different chemical domains are shown. Domains C/D are hardly distinguishable in BSE, owing to similar grey hues. BSE photo T. Mikuš.

Table 1. Electron microprobe data (in wt.%) of ferriandrosite-(Ce) (domains A–C) and associated vielleaureite-(Ce) (domain D).

Notes: e.s.d. = estimated standard deviation. The symbol ‘–’ indicates that the chemical constituent was below the detection limit.

1 Calculated to yield 8 (A+M+T) cations.

2 Calculated to yield 1 OH group per formula unit.

The empirical formula of ferriandrosite-(Ce), based on 13 anions pfu, distributing the chemical constituents among different structural sites, in agreement with the crystal structure refinement (see below), is (with rounding errors and REE* corresponding to minor REE) A 1(Mn2+0.63Ca0.35Ce0.02)Σ1.00 A 2(Ce0.53La0.27Nd0.14Pr0.05REE*0.01)Σ1.00 M 1(Fe3+0.41Al0.12V3+0.01Mg0.40Ti0.05)Σ0.99 M 2Al1.00 M 3(Mn2+0.75Fe2+0.22Mg0.03)Σ1.00 T 1–3Si3.00O11O4(O0.67F0.33)(OH). It corresponds to a ideal formula of ferriandrosite-(Ce) of MnCeFe3+AlMn2+(Si2O7)(SiO4)O(OH).

Table 1 also reports the average compositions of the other three domains (B, C, and D). The crystal chemical formulae for these three domains are the following:

$$\eqalign{& {\rm domain\ B}\colon \cr& ^{A1}( {{\rm M}{\rm n}^{2 + }_{0.55} {\rm Ca}_{0.45}^\phantom{0.00}} ) _{\Sigma 1.00}{}^{A2} ( {{\rm C}{\rm e}_{0.54}{\rm L}{\rm a}_{0.26}{\rm N}{\rm d}_{0.14}{\rm P}{\rm r}_{0.05}{\rm RE}{\rm E}^\ast_{0.01} } ) _{\Sigma 1.00}\cr& ^{M1} ( {{\rm F}{\rm e}^{3 + }_{0.73} {\rm A}{\rm l}_{0.02}{\rm V}^{3 + }_{0.01} {\rm M}{\rm g}_{0.10}{\rm F}{\rm e}^{2 + }_{0.07} {\rm T}{\rm i}_{0.06}} ) _{\Sigma 0.99}{}^{M2} {\rm A}{\rm l}_{1.00}\cr&^{M3} ( {{\rm M}{\rm n}^{2 + }_{0.63} {\rm F}{\rm e}^{2 + }_{0.31}} {{\rm M}{\rm g}_{0.06}} ) _{\Sigma 1.00}{}^{T1\ndash 3} {\rm S}{\rm i}_{3.00}{\rm O}_{11}{}^{{\rm O}4} ( {{\rm O}_{0.90}{\rm F}_{0.10}} ) _{\Sigma 1.00}( {{\rm OH}} ) ; \;}$$
$${\eqalign{& {\rm domain\ C}\colon\cr& ^{A1}( {{\rm M}{\rm n}^{2 + }_{0.69} {\rm C}{\rm a}_{0.31}} ) _{\Sigma 1.00}{}^{A2} ( {{\rm C}{\rm e}_{0.50}{\rm L}{\rm a}_{0.24}{\rm N}{\rm d}_{0.16}{\rm P}{\rm r}_{0.05}{\rm RE}{\rm E}^\ast_{0.01} {\rm C}{\rm a}_{0.04}} ) _{\Sigma 1.00}\cr&^{M1} \hskip -1pt( {{\rm F}{\rm e}^{3 + }_{0.37} {\rm A}{\rm l}_{0.25}{\rm M}{\rm n}^{3 + }_{0.13} {\rm M}{\rm g}_{0.24}\hskip -1.5pt{\rm T}{\rm i}_{0.01}} \hskip -.8pt) \hskip -.2pt_{\Sigma 1.00}{}^{M2} {\rm A}{\rm l}_{1.00}{}^{M3} \hskip -.6pt( {{\rm M}{\rm n}^{2 + }_{0.82} {\rm M}{\rm g}_{0.18}} \hskip -.5pt) _{\Sigma 1.00}\cr&^{T1\ndash 3} ( {{\rm S}{\rm i}_{2.97}{\rm A}{\rm l}_{0.03}} ) _{\Sigma 3.00}{\rm O}_{11}{}^{{\rm O}4} ( {{\rm O}_{0.70}{\rm F}_{0.30}} ) _{\Sigma 1.00}( {{\rm OH}} ) ; \;}}$$
$${\eqalign{& {\rm domain\ D}\colon\cr& ^{A1}( {{\rm M}{\rm n}^{2 + }_{0.78} {\rm C}{\rm a}_{0.21}{\rm C}{\rm e}_{0.01}} ) _{\Sigma 1.00}{}^{A2} ( {{\rm C}{\rm e}_{0.52}{\rm L}{\rm a}_{0.26}{\rm N}{\rm d}_{0.15}{\rm P}{\rm r}_{0.06}{\rm RE}{\rm E}^\ast_{0.01} } ) _{\Sigma 1.00}\cr&^{M1} \hskip -.8pt ( \hskip -.2pt{{\rm M}{\rm g}_{0.51}{\rm M}{\rm n}^{3 + }_{0.27} {\rm F}{\rm e}^{3 + }_{0.17} {\rm A}{\rm l}_{0.03}\hskip -1.2pt{\rm T}{\rm i}_{0.01}} \hskip -.5pt) _{\Sigma 1.00}{}\hskip -.7pt^{M2}\hskip -1.8pt {\rm A}{\rm l}_{1.00}{}\hskip -.7pt^{M3} \hskip -1pt( \hskip -.2pt{{\rm M}{\rm n}^{2 + }_{0.70} {\rm M}{\rm g}_{0.30}} \hskip -.5pt) _{\Sigma 1.00}\cr&^{T1\ndash 3} ( {{\rm S}{\rm i}_{2.96}{\rm A}{\rm l}_{0.04}} ) _{\Sigma 3.00}{\rm O}_{11}{}^{{\rm O}4} ( {{\rm F}_{0.53}{\rm O}_{0.47}} ) _{\Sigma 1.00}( {{\rm OH}} ).}}$$

Domains B and C correspond to the end-member formula MnCeFe3+AlMn2+(Si2O7)(SiO4)O(OH), i.e. to ferriandrosite-(Ce), however domain D leads to end-member formula MnCeMgAlMn2+(Si2O7)(SiO4)F(OH) and corresponds to a recently approved new member of the dollaseite group, vielleaureite-(Ce) (Ragu et al., Reference Ragu, Bindi, Bonazzi and Chopin2023). The domains A and B can be easily distinguished in back-scattered electron images, whereas C and D domains have similar grey hues. In the material studied, there is a clear positive correlation between Mg and F contents, in agreement with the substitution M 1Fe3+ + O4O2– = M 1Mg2+ + O4F (Fig. 4).

Figure 4. Relationship between F and Mg atoms per formula unit (apfu). Different colours correspond to the chemical homogenous domains: A = orange, B = green, C = red and D = violet.

X-ray crystallography

X-ray diffraction study and structural refinements

Single-crystal X-ray intensity data were collected using a Bruker D8 Venture four-circle diffractometer equipped with an air-cooled Photon III detector, and microfocus MoKα radiation. The detector-to-crystal distance was set to 38 mm. Data were collected using φ scan modes, in 0.5° slices, with an exposure time of 30 s per frame. A total of 1644 frames were collected and they were integrated with the Bruker SAINT software package using a narrow-frame algorithm. Data were corrected for Lorentz-polarisation, absorption, and background using the package of software Apex4 (Bruker AXS, 2022). Unit-cell parameters were refined on the basis of the XYZ centroids of 9895 reflections above 20 σI with 7.898° < 2θ < 66.315°. Ferriandrosite-(Ce) is monoclinic, space group P21/m (#11), with the following unit-cell parameters: a = 8.8483(4) Å, b = 5.7307(3) Å c = 10.0314(5) Å, β = 113.3659(15)°, V = 10.0314(5) Å3 and Z = 2.

The crystal structure of ferriandrosite-(Ce) was refined using Shelxl-2018 (Sheldrick, Reference Sheldrick2015) starting from the atomic coordinates of ferriakasakaite-(Ce) (Biagioni et al., Reference Biagioni, Bonazzi, Pasero, Zaccarini, Balestra, Bracco and Ciriotti2019). The position of the H atoms was located through the difference-Fourier map. The following scattering curves for neutral atoms, taken from the International Tables for Crystallography (Wilson, Reference Wilson1992), were used: Mn vs. Ca at A1, Ce vs. Ca at A2, Fe vs. Mg at M1, Al vs. Fe at M2, Mn vs. Mg at M3, Si at T1–T3 sites, H at the H10 site, and O at all the O sites but the O4 position, whose site occupancy was modelled using the scattering curves of O vs. F. Refinement of the O/F atomic ratio pointed to O0.65(6)F0.35(5), in agreement with the electron microprobe data. The anisotropic structural model (only H was refined isotropically) converged to R = 0.0210 for 1910 reflections with F o > 4σ(F o) and 127 refined parameters.

Details of data collection and refinement are given in Table 2. Fractional atom coordinates and equivalent isotropic or isotropic displacement parameters are reported in Table 3. Table 4 reports selected bond distances, whereas Table 5 compares observed and calculated mean atomic numbers at the A1, A2 and M1–M3 sites of ferriandrosite-(Ce). Finally, Table 6 gives the weighted bond-valence calculations obtained using the bond-valence parameters of Brese and O'Keeffe (Reference Brese and O'Keeffe1991). The crystallographic information files have been deposited with the Principal Editor of Mineralogical Magazine and are available as Supplementary material (see below).

Table 2. Crystal and experimental data for ferriandrosite-(Ce).

*w = 1/[σ2(F o2)+(1.6608P)2].

Table 3. Sites, fractional coordinates and isotropic (*) or equivalent isotropic displacement parameters (in Å2) for ferriandrosite-(Ce).

Table 4. Selected bond distances (in Å) in ferriandrosite-(Ce).

Table 5. Refined site scattering vs. calculated site scattering (in electrons) and site population (in apfu) at the A and M sites in ferriandrosite-(Ce).

Table 6. Weighted bond-valence balance (in vu) in ferriandrosite-(Ce).

Notes: left and right superscripts indicate the number of equivalent bonds involving anions and cations, respectively. For sites with mixed occupancy, the bond valences have been weighted according to site populations given in Table 5.

*Anion positions involved in the hydrogen bond O10–H⋅⋅⋅O4 [d (O10⋅⋅⋅O4) = 2.843(4) Å]. According to Ferraris and Ivaldi (Reference Ferraris and Ivaldi1988), the bond-valence sum (BVS) of such a bond is 0.17 vu. Consequently, the corrected BVS at O4 and O10 sites are 1.62 and 1.16 vu, respectively.

Owing to the inhomogeneous nature of the material studied, powder X-ray diffraction data were collected only on the same crystal used for single-crystal X-ray diffraction, simulating a Gandolfi-like pattern using a Bruker D8 Venture single-crystal diffractometer equipped with a Photon III area detector and microfocused CuKα radiation. The observed X-ray diffraction pattern is reported in Table 7. Unit-cell parameters refined for the monoclinic space group P21/m from the powder data using the method of Holland and Redfern (Reference Holland and Redfern1997) on the basis of 19 unequivocally indexed reflections are as follows: a = 8.889(2) Å, b = 5.7356(16) Å, c = 10.0262(17) Å, β = 113.504(17)°, V = 468.75(12) Å3 and Z = 2.

Table 7. Powder X-ray diffraction data (d in Å) for ferriandrosite-(Ce).

Notes: Intensity and d hkl were calculated using the software PowderCell 2.3 (Kraus and Nolze, Reference Kraus and Nolze1996) on the basis of the structural model given in Table 3. Only the reflections with I calc > 10 are given, if not observed. Intensities were visually estimated: vs = strong; s = strong; ms = medium–strong; m = medium; mw = medium–weak; w = weak; vw = very weak.

1These reflections correspond to more than two calculated reflections with I < 10.

Crystal structure description

Ferriandrosite-(Ce) is isotypic with the other members of the epidote supergroup (e.g. Dollase, Reference Dollase1968). The crystal structure can be described as formed by single chains of edge-sharing M2-centered octahedra and zig-zag chains of M1-centered octahedra with M3-centered octahedra attached on alternate sides along b. Octahedral chains are bonded to Si2O7 and SiO4 groups. In this octahedral–tetrahedral framework, two kinds of structural cavities occur, i.e. the smaller cavity hosting the nine-fold coordinated A1 site and the larger one where the ten-fold coordinated A2 site is located.

Cation sites

In ferriandrosite-(Ce), the A1 site has a mixed (Mn,Ca) site occupancy, with a Mn/(Mn+Ca) atomic ratio of 0.65. Mean atomic number (Table 5) and weighted bond-valence sum (Table 6) agree with this occupancy. As discussed by previous authors (e.g. Bonazzi et al., Reference Bonazzi, Menchetti and Reinecke1996; Nagashima et al., Reference Nagashima, Armbruster, Akasaka and Minakawa2010, Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2013, Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015), the replacement of Ca2+ by Mn2+ promotes a decrease in the coordination number of the A1 site that can be described as six-fold coordinated. This coordination number decrease is related to the shift of the O6 and O9 atoms away from the A1 site, whereas O3 and O1 get closer to the cation site; O5 and O7 are unaffected by these changes. If one defines the difference between the bond distances A1–O6 and A1–O5 (with O6 and O5 being the seventh and sixth ligands of A1, respectively) as δ7–6, a direct relation between the Mn2+ content at A1 and the δ7–6 value can be observed (e.g. Bonazzi et al., Reference Bonazzi, Menchetti and Reinecke1996; Nagashima et al., Reference Nagashima, Armbruster, Akasaka and Minakawa2010, Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2013, Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015; Biagioni et al., Reference Biagioni, Bonazzi, Pasero, Zaccarini, Balestra, Bracco and Ciriotti2019). In ferriandrosite-(Ce), this difference is 0.508 Å, to be compared with the theoretical value of 0.502 Å calculated on the basis of the regression line given by Bonazzi et al. (Reference Bonazzi, Menchetti and Reinecke1996). Moreover, a decrease in the [6]<A1–O> can be observed and indeed, ferriandrosite-(Ce) shows the low value of 2.314 Å, to be compared with 2.320, 2.322, 2.310 and 2.335 Å reported for manganiandrosite-(La), manganiandrosite-(Ce), vanadoandrosite-(Ce) and ferriandrosite-(La), respectively (Bonazzi et al., Reference Bonazzi, Menchetti and Reinecke1996; Cenki-Tok et al., Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006; Nagashima et al., Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015).

The shift of the O9 site away from the cation located at the A1 position is associated with a reduction of the Si1–O9–Si2 bond angle that is related not only to the occupancy at the A1 site but also to the average bond length at M3 (e.g. Cenki-Tok et al., Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006). In ferriandrosite-(Ce), O9 is at 3.148 and 3.227 Å from A1, with a Si1–O9–Si2 bond angle of 137.6(2)°; this angle is similar to those reported by Cenki-Tok et al. (Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006) for manganiandrosite-(Ce) and vanadoandrosite-(Ce), i.e. 137.4(6) and 137.4(8)°, respectively, and slightly smaller than that given by Nagashima et al. (Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015) for ferriandrosite-(La), i.e. 139.4(2)°; however, it is smaller than those reported by Biagioni et al. (Reference Biagioni, Bonazzi, Pasero, Zaccarini, Balestra, Bracco and Ciriotti2019) in manganiakasakaite-(La) and ferriakasakaite-(Ce), i.e. 140.88(19) and 144.47(14)°, respectively.

The A2 site of ferriandrosite-(Ce) is a REE-bearing site, with Ce as the dominant element. Bonazzi et al. (Reference Bonazzi, Menchetti and Reinecke1996) observed a relationship between the ratio of the cell parameters c/V and the REE content in species belonging to the series piemontite – ‘androsite’, as well as between the amount of REE and the β angle. In ferriandrosite-(Ce), the c/V ratio is 0.0215, and the β angle is 113.37°. On the basis of the relationship given by Bonazzi et al. (Reference Bonazzi, Menchetti and Reinecke1996), considering ΣREE = 1 atom per formula unit (apfu), one should have c/V = 0.0213 and β = 113.4°, in agreement with observed values. Actually, as stressed by Nagashima et al. (Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015), the value of the β angle is also affected by the Mn2+ at the A1 site. The value observed in ferriandrosite-(Ce) can be compared with those reported for other ‘androsites’, i.e. 113.88° in manganiandrosite-(La) (Bonazzi et al., Reference Bonazzi, Menchetti and Reinecke1996), 113.84° in ferriandrosite-(La) (Nagashima et al., Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015), 113.42° in manganiandrosite-(Ce), and 113.09° in vanadoandrosite-(Ce) (Cenki-Toc et al., Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006).

Among the three independent M sites, cation assignments were based on refined mean atomic numbers (Table 5) and bond-valence sums (Table 6). M2 hosts Al only; a possible minor replacement of Al by Fe3+ (~2 at.%) or Ti4+ (~3 at.%) cannot be discarded, as observed in other samples, e.g. in manganiandrosite-(La) by Bonazzi et al. (Reference Bonazzi, Menchetti and Reinecke1996) who reported 4% Fe3+ in M2 and in ferriandrosite-(La) by Nagashima et al. (Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015) who hypothesised the presence of 0.04 Ti at the M2 site. The <M2–O> distance is 1.894 Å, to be compared with those observed in other androsites, i.e. 1.892 Å in manganiandrosite-(La) (Bonazzi et al., Reference Bonazzi, Menchetti and Reinecke1996), 1.902 and 1.894 Å in manganiandrosite-(Ce) and vanadoandrosite-(Ce), respectively (Cenki-Tok et al., Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006), and 1.902 Å in ferriandrosite-(La) (Nagashima et al., Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015). The M3 site is occupied mainly by Mn2+, with minor Fe2+ and trace Mg2+. This occupancy is in keeping with the large observed <M3–O>, 2.185 Å. This average distance is larger than that observed in manganiandrosite-(La) (i.e. 2.159 Å, where 0.28 M3+ apfu occur), and similar to those reported by Cenki-Tok et al. (Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006) for M3 sites in manganiandrosite-(Ce) (2.197 Å) and vanadoandrosite-(Ce) (2.195 Å), where only Mn2+ occurs. The slight contraction observed in ferriandrosite-(Ce) is probable due to the minor Fe2+–Mn2+ replacement, in agreement with the smaller ionic radius of VIFe2+ (0.78 Å) with respect to that of VIMn2+ (0.83 Å), according to Shannon (Reference Shannon1976). The M1 site has a complex chemistry, with mainly Fe3+ and Mg2+, minor Al and Ti, and trace amounts of V. Whereas Fe3+ and Mg2+ occur in the same amount (0.41 apfu), the sum of trivalent cations (Fe3+ + Al + V) is larger than those of divalent ones (Mg), i.e. 0.54 vs. 0.41, with Fe3+ as the dominant constituent of the dominant valence (e.g. Bosi et al., Reference Bosi, Biagioni and Oberti2019a, Reference Bosi, Hatert, Hålenius, Pasero, Miyawaki and Mills2019b). The average bond distance is 2.012 Å, similar to those observed in other androsites, i.e. 2.010, 2.019, 2.011 and 2.002 Å for manganiandrosite-(La), manganiandrosite-(Ce), vanadoandrosite(Ce) and ferriandrosite-(La), respectively (Bonazzi et al., Reference Bonazzi, Menchetti and Reinecke1996; Cenki-Tok et al., Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006; Nagashima et al., Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015). As discussed by Cenki-Tok et al. (Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006), the occurrence of Mn3+ at the M1 site increases the difference between the distances M1–O1 and M1–O4, owing to the Jahn–Teller effect shown by this cation; this difference is 0.231 and 0.227 Å in manganiandrosite-(La) and manganiandrosite-(Ce), respectively (Bonazzi et al., Reference Bonazzi, Menchetti and Reinecke1996; Cenki-Tok et al., Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006), whereas in vanadoandrosite-(Ce) and ferriandrosite-(La) this difference is smaller, i.e. 0.161 and 0.137 Å, respectively (Cenki-Tok et al., Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006; Nagashima et al., Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015). In ferriandrosite-(Ce), a difference of 0.145 Å was observed, in agreement with the absence of Mn3+ at the M1 site.

Three independent Si-centred tetrahedra occur in ferriandrosite-(Ce). Si1 and Si2 are bonded, sharing the O9 atom and forming Si2O7 groups; as discussed above, the Si1–O9–Si2 bond angle is sensitive to the occupancy at the A1 and the bond length at the M3 site. Si3 forms an isolated SiO4 group. Average bond distances range between 1.624 and 1.637 Å, with bond-valence sums between 3.86 and 4.00 valence units (vu).

Anion sites

Ten independent anion sites occur in ferriandrosite-(Ce). Among them, eight are four-fold coordinated, with bond-valence sums ranging between 1.78 and 2.01 vu; these sites are occupied by O2–. Two sites, namely O4 and O10, are three-fold coordinated and are underbonded, with bond-valence sums of 1.45 and 1.33 vu, respectively. The crystal-structure refinement allowed us to locate a residual maximum that was interpreted as a H atom located at 0.95(9) Å from O10. Moreover, the O10⋅⋅⋅O4 distance, 2.841(4) Å, agrees with a H bond, corresponding to a bond strength, calculated according to Ferraris and Ivaldi (Reference Ferraris and Ivaldi1988), of 0.17 vu. In this way, the correct bond-valence sums at O4 and O10 are 1.62 and 1.16 vu, respectively. These values agree with the mixed nature (O,F) of the O4 site, with an O/(O+F) atomic ratio close to 0.65 (in accord with both chemical and structural data) and the occurrence of (OH) at O10. Moreover, it is worth noting that the O10⋅⋅⋅O4 is similar to those observed in manganiandrosite-(Ce), ferriandrosite-(La), and vanadoandrosite-(Ce) (ca. 2.87 Å – Cenki-Tok et al., Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006; Nagashima et al., Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015), and shorter than that reported by Bonazzi et al. (Reference Bonazzi, Menchetti and Reinecke1996) in manganiandrosite-(La), i.e. 2.94 Å. This difference may be explained considering that the O4 site is bonded to two M1 and one M3 sites; in all the known species belonging to the androsite series but manganiandrosite-(La), the M3 site has a virtually pure M2+ occupancy. On the contrary, the latter species has 0.28 M3+ apfu, and consequently the bond strength on O4 is higher than in the other cases, thus favouring an elongation (= a weakening) of the H bond.

Relations with other epidote supergroup minerals

Ferriandrosite-(Ce) is a new member of the allanite group, composed, after the addition of this new species, of 18 minerals (Table 8). Five root-names, based on different combinations of cations at the A1 and M3 sites, are currently known: allanite (A1 = Ca2+, M3 = Fe2+), akasakaite (A1 = Ca2+, M3 = Mn2+), androsite (A1 = Mn2+, M3 = Mn2+), dissakisite (A1 = Ca2+, M3 = Mg2+), and uedaite (A1 = Mn2+, M3 = Fe2+).

Table 8. Valid mineral species belonging to the allanite group.

The name ‘androsite’ was first introduced by Bonazzi et al. (Reference Bonazzi, Menchetti and Reinecke1996) on the basis of a sample from a small metamorphic Mn ore body in the Andros Island, Cyclades, Greece. The new species was named ‘androsite-(La)’, later renamed manganiandrosite-(La) by Armbruster et al. (Reference Armbruster, Bonazzi, Akasaka, Bermanec, Chopin, Gieré, Heuss-Assbichler, Liebscher, Menchetti, Pan and Pasero2006), owing to the dominance of Mn3+ at the M1 site. Later, Cenki-Tok et al. (Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006) described the Ce-analogue manganiandrosite-(Ce) from the Praborna mine (Aosta Valley, Italy), along with the new species vanadoandrosite-(Ce), that has V3+ as the dominant cation at the M1 position, from a Mn ore deposit located near Vielle-Aure, Hautes-Pyrénées, France. Finally, Nagashima et al. (Reference Nagashima, Nishio-Hamane, Tomita, Minakawa and Inaba2015) identified the Fe3+-analogue of manganiandrosite-(La), i.e. ferriandrosite-(La) from a Mn deposit in the Mie Prefecture, Japan. The Ce-analogue of this latter species, ferriandrosite-(Ce), is now added to this series. As reported in the Introduction, samples corresponding to this species were reported by Girtler et al. (Reference Girtler, Tropper and Hauzenberger2013) and Kolitsch et al. (Reference Kolitsch, Schachinger, Auer and Walter2021). In particular, the former authors first used the name ‘ferriandrosite-(Ce)’ without approval by the IMA-CNMNC.

The partial replacement of Mn2+ by Ca2+ at the A1 site in androsite-series minerals is in keeping with the possible solid solution between members of this series and those of the akasakaite series, as suggested, for instance, by Biagioni et al. (Reference Biagioni, Bonazzi, Pasero, Zaccarini, Balestra, Bracco and Ciriotti2019), who described strongly zoned grains containing ferriakasakaite-(Ce), manganiandrosite-(Ce), and the potential ‘androsite-(Ce)’ end-member in samples from the Mn mineralisation of Monte Maniglia, Piedmont, Italy.

As shown in Fig. 4, the occurrence of the ‘dollaseite-type’ substitution is effective in the androsite series, as previously reported by Cenki-Tok et al. (Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006). This lead to the appearance of a species having end-member composition MnCeMgAlMn2+(Si2O7)(SiO4)F(OH). Such a composition would correspond to the Mn-analogue of khristovite-(Ce), a species defined by Pautov et al. (Reference Pautov, Khorov, Ignatenko, Sokolova and Nadezhina1993) as MnCeMgAlMn2+(Si2O7)(SiO4)F(OH) and recently described by Ragu et al. (Reference Ragu, Bindi, Bonazzi and Chopin2023) as a new mineral, vielleaureite-(Ce). Actually, as discussed by Cenki-Tok et al. (Reference Cenki-Tok, Ragu, Armbruster, Chopin and Medenbach2006), there are some doubts about the actual definition of khristovite-(Ce) as the structural data, reported by Sokolova et al. (Reference Sokolova, Nadezhina and Pautov1991), suggests the dominance of Mn2+ at the A1 site.

Conclusions

Ferriandrosite-(Ce), Mn2+CeFe3+AlMn2+(Si2O7)(SiO4)O(OH), is a new member of the epidote supergroup and another REE-bearing epidote first identified from Mn ore deposits. In this kind of occurrence, epidote-supergroup minerals can host REE and variable amounts of Fe and Mn, showing different oxidation states reflecting the $ f_{{\rm O}_2}\! $ conditions occurring during the geological evolution of the Mn-rich assemblages.

The possibility of extracting high-quality crystal-chemical information from micrometre-sized volumes of matter, as exemplified by the results obtained on the strongly zoned grains of ferriandrosite-(Ce), opens interesting future scenarios for unravelling the evolution of Mn ore deposits and the ability to shed further light on the complex geological processes shaping our planet.

Acknowledgements

The helpful comments of Principal Editor Stuart Mills and Structural Editor Peter Leverett, as well as two other anonymous reviewers are greatly appreciated. The study was financially supported by VEGA project 2/0029/23 and by Slovak Research and Development Agency project APVV-22-0041. CB acknowledges funding by the Italian Ministero dell'Istruzione, dell'Università e della Ricerca through the project PRIN 2020 “HYDROX – HYDRous-vs-Oxo-components in minerals: adding new pieces to the Earth's H2O cycle puzzle”, prot. 2020WYL4NY.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1180/mgm.2023.62.

Competing interests

The authors declare none.

Footnotes

Associate Editor: Daniel Atencio

References

Armbruster, T., Bonazzi, P., Akasaka, M., Bermanec, V., Chopin, C., Gieré, R., Heuss-Assbichler, S., Liebscher, A., Menchetti, S., Pan, Y. and Pasero, M. (2006) Recommended nomenclature of epidote-group minerals. European Journal of Mineralogy, 18, 551567.CrossRefGoogle Scholar
Bajaník, Š., Ivanička, J., Mello, J., Pristaš, J., Reichwalder, P., Snopko, L., Vozár, J. and Vozárová, A. (1983) Geological map of the Slovenské Rudohorie Mts. – eastern part 1: 50.000. Dionýz Štúr Institute of Geology, Bratislava.Google Scholar
Bayliss, P. and Levinson, A.A. (1988) A system of nomenclature for rare-earth mineral species: revision and extension. American Mineralogist, 73, 422423.Google Scholar
Biagioni, C., Bonazzi, P., Pasero, M., Zaccarini, F., Balestra, C., Bracco, R. and Ciriotti, M.E. (2019) Manganiakasakaite-(La) and ferriakasakaite-(Ce), two new epidote supergroup minerals from Piedmont, Italy. Minerals, 9, 353.CrossRefGoogle Scholar
Bonazzi, P., Menchetti, S. and Reinecke, T. (1996) Solid solution between piemontite and androsite-(La), a new mineral of the epidote group from Andros Island, Greece. American Mineralogist, 81, 735742.CrossRefGoogle Scholar
Bosi, F., Biagioni, C. and Oberti, R. (2019a) On the chemical identification and classification of minerals. Minerals, 9, 591.CrossRefGoogle Scholar
Bosi, F., Hatert, F., Hålenius, U., Pasero, M., Miyawaki, R. and Mills, S. (2019b) On the application of the IMA-CNMNC dominant-valency rule to complex mineral compositions. Mineralogical Magazine, 83, 627632.CrossRefGoogle Scholar
Brese, N.E. and O'Keeffe, M. (1991) Bond-valence parameters for solids. Acta Crystallographica, B47, 192197.CrossRefGoogle Scholar
Bruker AXS Inc. (2022) APEX 4. Bruker Advanced X-ray Solutions, Madison, Wisconsin, USA.Google Scholar
Cenki-Tok, B., Ragu, A., Armbruster, T., Chopin, C. and Medenbach, O. (2006) New Mn- and rare-earth-rich epidote group minerals in metacherts: manganiandrosite-(Ce) and vanadoandrosite-(Ce). European Journal of Mineralogy, 18, 569582.CrossRefGoogle Scholar
Dollase, W.A. (1968) Refinement and comparison of the structure of zoisite and clinozoisite. American Mineralogist, 53, 18821888.Google Scholar
Faryad, S.W. (1994) Mineralogy of Mn-rich rocks from greenschist facies sequences of the Gemericum, West Carpathians, Slovakia. Neues Jahrbuch für Mineralogie, Monatshefte, 10, 464480.Google Scholar
Ferraris, G. and Ivaldi, G. (1988) Bond valence vs bond length in O⋅⋅⋅O hydrogen bonds. Acta Crystallographica, B44, 341344.CrossRefGoogle Scholar
Girtler, D., Tropper, P. and Hauzenberger, C. (2013) Androsite-(Ce) and ferriandrosite-(Ce) as indicator for low-grade REE mobility in the Veitsch Mn Deposit (Styria). Mitteilungen der Österreichischen Mineralogischen Gesellschaft, 159, 56.Google Scholar
Grecula, P., Abonyi, A., Abonyiová, M., Antaš, J., Bartalský, B., Bartalský, J., Dianiška, I., Ďuďa, R., Gargulák, M., Gazdačko, Ľ., Hudáček, J., Kobulský, J., Lörincz, L., Macko, J., Návesňák, D., Németh, Z., Novotný, L., Radvanec, M., Rojkovič, I., Rozložník, L., Varček, C. and Zlocha, Z. (1995 ) Mineral Deposits of the Slovak Ore Mountains, Vol 1. Geocomplex, Bratislava, 834 pp.Google Scholar
Grew, E.S., Essene, E.J., Peacor, D.R., Su, S-C. and Asami, M. (1991) Dissakisite-(Ce), a new member of the epidote group and the Mg analogue of allanite-(Ce) from Antarctica. American Mineralogist, 76, 19901997.Google Scholar
Holland, T.J.B. and Redfern, S.A.T. (1997) Unit cell refinement from powder diffraction data: the use of regression diagnostics. Mineralogical Magazine, 61, 6577.CrossRefGoogle Scholar
Kantor, J. (1954) On the genesis of manganese ores in the Spišsko-gemerské rudohorie Mts. Geologické Práce, Zprávy, 1, 7071 [in Slovak].Google Scholar
Kartashov, P.M., Ferraris, G., Ivaldi, G., Sokolova, E. and McCammon, C.A. (2002) Ferriallanite-(Ce), CaCeFe3+AlFe2+(SiO4)(Si2O7)O(OH), a new member of the epidote group: description, X-ray and Mössbauer study. The Canadian Mineralogist, 40, 16411648.CrossRefGoogle Scholar
Kolitsch, U., Mills, S.K., Miyawaki, R. and Blass, G. (2012) Ferriallanite-(La), a new member of the epidote supergroup from the Eifel, Germany. European Journal of Mineralogy, 24, 741747.CrossRefGoogle Scholar
Kolitsch, U., Schachinger, T. and Auer, C. (2021) Erste mineralogische Untersuchungen an den metamorphen mesozoischen Manganvorkommen der Brandlscharte südlich des Imbachhorns, Kapruner und Fuscher Tal, Salzburg: Nachweise von Alabandin, Cobaltit, Fluorit, Gersdorffit, Gold, Hübnerit, Molybdänit, Percleveit-(La)(?), Sonolith, Thorit, Tsumoit, Ullmannit, Uraninit, Vittinkiit, Zinkgruvanit und weiteren Mineralien. Pp. 209220 in Neue Mineralfunde aus Österreich LXX (Walter, F. et al., editor). Carinthia II, 211/131.Google Scholar
Kraus, W. and Nolze, G. (1996) POWDER CELL – a program for the representation and manipulation of crystal structures and calculation of the resulting X-ray powder patterns. Journal of Applied Crystallography, 29, 301303.CrossRefGoogle Scholar
Maderspach, L. (1875) Die Manganerze von Csucsom und Betlér bei Rosenau im Gömörer Comitate. Österreichische Zeitschrift für Berg- und Hüttenwesen, 23, 548550.Google Scholar
Mandarino, J.A. (1979) The Gladstone-Dale relationship. Part III. Some general applications. The Canadian Mineralogist, 17, 7176.Google Scholar
Mandarino, J.A. (1981) The Gladstone-Dale relationship. Part IV. The compatibility concept and its application. The Canadian Mineralogist, 19, 441450.Google Scholar
Miyawaki, R., Yokoyama, K., Matsubara, S., Tsutsumi, Y. and Goto, A. (2008) Uedaite-(Ce), a new member of the epidote group with Mn at the A site, from Shodoshima, Kagawa Prefecture, Japan. European Journal of Mineralogy, 20, 261269.CrossRefGoogle Scholar
Nagashima, M., Armbruster, T., Akasaka, M. and Minakawa, T. (2010) Crystal chemistry of Mn2+-, Sr-rich and REE-bearing piemontite from the Kamisugai mine in the Sambagawa metamorphic belt, Shikoku, Japan. Journal of Mineralogical and Petrological Sciences, 105, 142150.CrossRefGoogle Scholar
Nagashima, M., Nishio-Hamane, D., Tomita, N., Minakawa, T. and Inaba, S. (2013) Vanadoallanite-(La): a new epidote-supergroup mineral from Ise, Mie Prefecture, Japan. Mineralogical Magazine, 77, 27392752.CrossRefGoogle Scholar
Nagashima, M., Nishio-Hamane, D., Tomita, N., Minakawa, T. and Inaba, S. (2015) Ferriakasakaite-(La) and ferriandrosite-(La): new epidote-supergroup minerals from Ise, Mie Prefecture, Japan. Mineralogical Magazine, 79, 735753.CrossRefGoogle Scholar
Nel, H.J., Strauss, C.A. and Wickman, F.E. (1949) Lombaardite, a new mineral from the Zaaiplaats Tin Mine, Central Transvaal. Union of South Africa Department of Mines, Geological Survey Memoir, 43, 4557.Google Scholar
Orlandi, P. and Pasero, M. (2006) Allanite-(La) from Buca della Vena mine, Apuan Alps, Italy, an epidote group mineral. The Canadian Mineralogist, 44, 6368.CrossRefGoogle Scholar
Pautov, L.A., Khorov, P.V., Ignatenko, K.I., Sokolova, E.V. and Nadezhina, T.N. (1993) Khristovite-(Ce), (Ca,REE)REE(Mg,Fe)AlMnSi3O11(OH)(F,O) – a new mineral of the epidote group. Zapiski Vserossijskogo Mineralogicheskogo Obshchestva, 122, 103111.Google Scholar
Peterec, D. and Ďuďa, R. (2003) Rare minerals of Mn deposit near Čučma. Natura Carpathica, 44, 229233 [in Slovak].Google Scholar
Ragu, A., Bindi, L., Bonazzi, P. and Chopin, C. (2023) Vielleaureite-(Ce), IMA 2022-134. CNMNC Newsletter 72. Mineralogical Magazine, 87, https://doi.org/10.1180/mgm.2023.21Google Scholar
Rojkovič, I. (2001) Early Paleozoic Manganese ores in the Gemericum Superunit, Western Carpathians, Slovakia. Geolines, 13, 3441.Google Scholar
Shannon, R.D. (1976) Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallographica, A32, 751767.CrossRefGoogle Scholar
Sheldrick, G.M. (2015) Crystal structure refinement with SHELXL. Acta Crystallographica, C71, 38.Google Scholar
Škoda, R., Cempírek, J., Filip, J., Novák, M., Veselovský, F. and Čtvrtlík, R. (2012) Allanite-(Nd), CaNdAl2Fe2+(SiO4)(Si2O7)O(OH), a new mineral from Åskagen, Sweden. American Mineralogist, 97, 983988.CrossRefGoogle Scholar
Sokolova, E.V., Nadezhina, T.N. and Pautov, L.A. (1991) Crystal structure of a new natural silicate of manganese from the epidote group. Soviet Physics Crystallography, 36, 172174.Google Scholar
Števko, M., Myšľan, P., Biagioni, C., Mauro, D. and Mikuš, T. (2023) Ferriandrosite-(Ce), IMA 2023-022. CNMNC Newsletter 74. Mineralogical Magazine, 87, https://doi.org/10.1180/mgm.2023.54Google Scholar
Thomson, T. (1811) Experiments on allanite, a new mineral from Greenland. A Journal of Natural Phylosophy, Chemistry, and the Arts, 29, 4759.Google Scholar
Tumiati, S., Godard, G., Martin, S., Nimis, P., Mair, V. and Boyer, B. (2005) Dissakisite-(La) from the Ulten zone peridotite (Italian Eastern Alps): a new end-member of the epidote group. American Mineralogist, 90, 11771185.CrossRefGoogle Scholar
Wilson, A.J.C. (editor) (1992) International Tables for Crystallography, Volume C: Mathematical, Physical and Chemical Tables. Kluwer Academic, Dordrecht, The Netherlands.Google Scholar
Figure 0

Figure 1. Dark brown aggregate of ferriandrosite-(Ce) enclosed in pyroxmangite (light pink)–spessartine (light yellow) matrix. Field of view is 2.06 mm. Photo L. Hrdlovič. Holotype material (P1P 2/2023, polished section Be-5).

Figure 1

Figure 2. Aggregates and subhedral grains of ferriandrosite-(Ce) (white) enclosed in pyroxmangite (light grey) associated with euhedral crystals of spessartine (dark grey). The red box indicates the area where the grain used for single-crystal X-ray diffraction was picked up. Back-scattered electron (BSE) photo T. Mikuš. Holotype material (P1P 2/2023, polished section Be-5).

Figure 2

Figure 3. Back-scattered electron (BSE) image showing the chemical inhomogeneity of the material studied. Different chemical domains are shown. Domains C/D are hardly distinguishable in BSE, owing to similar grey hues. BSE photo T. Mikuš.

Figure 3

Table 1. Electron microprobe data (in wt.%) of ferriandrosite-(Ce) (domains A–C) and associated vielleaureite-(Ce) (domain D).

Figure 4

Figure 4. Relationship between F and Mg atoms per formula unit (apfu). Different colours correspond to the chemical homogenous domains: A = orange, B = green, C = red and D = violet.

Figure 5

Table 2. Crystal and experimental data for ferriandrosite-(Ce).

Figure 6

Table 3. Sites, fractional coordinates and isotropic (*) or equivalent isotropic displacement parameters (in Å2) for ferriandrosite-(Ce).

Figure 7

Table 4. Selected bond distances (in Å) in ferriandrosite-(Ce).

Figure 8

Table 5. Refined site scattering vs. calculated site scattering (in electrons) and site population (in apfu) at the A and M sites in ferriandrosite-(Ce).

Figure 9

Table 6. Weighted bond-valence balance (in vu) in ferriandrosite-(Ce).

Figure 10

Table 7. Powder X-ray diffraction data (d in Å) for ferriandrosite-(Ce).

Figure 11

Table 8. Valid mineral species belonging to the allanite group.

Supplementary material: File

Števko et al. supplementary material 1

Števko et al. supplementary material
Download Števko et al. supplementary material 1(File)
File 492.8 KB
Supplementary material: File

Števko et al. supplementary material 2

Števko et al. supplementary material
Download Števko et al. supplementary material 2(File)
File 310.7 KB