Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-26T12:14:33.476Z Has data issue: false hasContentIssue false

Fluorbritholite-(Nd), Ca2Nd3(SiO4)3F, a new and key mineral for neodymium sequestration in REE skarns

Published online by Cambridge University Press:  08 June 2023

Dan Holtstam*
Affiliation:
Department of Geosciences, Swedish Museum of Natural History, Box 50007, SE-104 05 Stockholm, Sweden
Patrick Casey
Affiliation:
Geological Survey of Sweden, Villavägen 18, SE-752 36, Uppsala, Sweden
Luca Bindi
Affiliation:
Dipartimento di Scienze della Terra, Università degli Studi di Firenze, Via G. La Pira 4, 50121, Firenze, Italy
Hans-Jürgen Förster
Affiliation:
Helmholtz Centre Potsdam GFZ German Research Centre for Geosciences, Telegrafenberg, 14473 Potsdam, Germany
Andreas Karlsson
Affiliation:
Department of Geosciences, Swedish Museum of Natural History, Box 50007, SE-104 05 Stockholm, Sweden
Oona Appelt
Affiliation:
Helmholtz Centre Potsdam GFZ German Research Centre for Geosciences, Telegrafenberg, 14473 Potsdam, Germany
*
Corresponding author: Dan Holtstam; Email: dan.holtstam@nrm.se
Rights & Permissions [Opens in a new window]

Abstract

Fluorbritholite-(Nd), ideally Ca2Nd3(SiO4)3F, has been approved by the International Mineralogical Association (IMA2023–001) and constitutes a new member of the britholite group of the apatite supergroup. It occurs in skarn from the Malmkärra iron mine, Norberg, Västmanland (one of the Bastnäs-type deposits in Sweden), associated with calcite, dolomite, magnetite, lizardite, talc, fluorite, baryte, scheelite, gadolinite-(Nd) and other REE minerals. Fluorbritholite-(Nd) forms anhedral and small grains, rarely up to 250 μm across. They are brownish pink and transparent with a vitreous to greasy lustre. The mineral is brittle, with an uneven or subconchoidal fracture and lacks a cleavage. In thin section, the mineral is nonpleochroic, uniaxial (–). Dcalc = 4.92(1) g⋅cm−3 and ncalc = 1.795. The empirical chemical formula from electron microprobe (WDS) point analyses is (Ca1.62Nd0.97Ce0.83Y0.52Sm0.30Gd0.23Pr0.17La0.16Dy0.11Er0.03Tb0.03Ho0.01Yb0.01)Σ4.99(Si2.92P0.08As0.01)Σ3.01O12.00[O0.48F0.26(OH)0.14Cl0.10Br0.02]Σ1.00. The crystal structure of fluorbritholite-(Nd) was refined from single-crystal X-ray diffraction data to R1= 0.043 for 704 unique reflections. It belongs to the hexagonal system, space group P63/m, with unit cell parameters a = 9.5994(3), c = 6.9892(4) Å and V = 557.76(5) Å3 for Z = 2. Fluorbritholite-(Nd) and other britholite-group minerals are a major sink for neodymium in REE-bearing skarns of Bastnäs type.

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 global quest for economic sources of rare earth elements (REE), prompted by the increased demand for neodymium (Nd) and related critical metals as components in e.g. permanent magnetic materials and lasers (Chakhmouradian and Wall, Reference Chakhmouradian and Wall2012; Goodenough et al., Reference Goodenough, Wall and Merriman2018), has revived interest in REE deposits of the Bergslagen ore region in Sweden (Sadeghi et al., Reference Sadeghi, Arvanitidis, Ripa, Jonsson, Nysten, Bergman, Söderhielm and Claeson2019). In particular, the occurrences of skarn-hosted britholite-group minerals — referred to as Bastnäs-type deposits, subtype 2, by Holtstam and Andersson (Reference Holtstam and Andersson2007) — in the Norberg District are of interest because of their unusual enrichment in heavier REE. Previous work (Holtstam and Andersson, Reference Holtstam and Andersson2007) included a few chemical analyses of britholite-group minerals suggesting the existence of a Nd-dominant species, corresponding to UM2007–044 in the list of valid unnamed mineral species (Smith and Nickel, Reference Smith and Nickel2007). It has now been approved by the Commission on New Minerals, Nomenclature and Classification of the International Mineralogical Association (IMA2023–001, Holtstam et al., Reference Holtstam, Casey, Bindi, Förster and Karlsson2023) with the name fluorbritholite-(Nd), ideally Ca2Nd3(SiO4)3F. The mineral was discovered at one of the Bastnäs-type deposits, Malmkärra, during a recent survey and sampling campaign. Type material is preserved in the mineral collection of the Department of Geosciences, Swedish Museum of Natural History, Box 50007, SE-10405 Stockholm, Sweden, under collection number GEO-NRM #20220221 (specimen and polished section). The accepted mineral symbol for fluorbritholite-(Nd) is Fbri-Nd.

Occurrence

The Malmkärra iron mine is situated close to the tarn Stora Malmtjärnen, Norberg, Västmanland County, Sweden (60°3’34’’N, 15°50’45’’E, 200 m a.s.l.). The deposit is hosted in a synclinal marble layer (dolomite), intercalated with felsic metavolcanic beds, and is divided into separate ore bodies due to local faulting (Geijer, Reference Geijer1936). The underlying bedrock consists of plastically deformed mica-schist, with major quartz, muscovite and cordierite, along with accessory tourmaline, magnetite and allanite-(Ce). It grades into a Na-rich, less altered metavolcanic rock in certain sections. The magnetite-skarn ore has locally replaced the carbonate along its contact with the country rock (Geijer, Reference Geijer1936; Holtstam et al., Reference Holtstam, Andersson, Broman and Mansfeld2014).

The oldest record of the Malmkärra mine is from 1664. Magnetite ore was mined (~100,000 t of Fe) until production stopped in 1936 (Geijer and Magnusson, Reference Geijer and Magnusson1944). Geijer (Reference Geijer1927) first noted the REE mineralisation, later referred to as Bastnäs-type (Geijer, Reference Geijer1961). The mineralisation occurs as bands of REE-bearing silicates deposited between magnetite-rich, tremolite-bearing skarns and dolomitic marble, with patches of Mg silicates (humite and serpentine) in recrystallised carbonate rock (‘ophicalcite’). Malmkärra is the type locality for three rare mineral species so far: västmanlandite-(Ce) (Holtstam et al., Reference Holtstam, Kolitsch and Andersson2005), ulfanderssonite-(Ce) (Holtstam et al., Reference Holtstam, Bindi, Hålenius, Kolitsch and Mansfeld2017) and gadolinite-(Nd) (Škoda et al., Reference Škoda, Plášil, Čopjaková, Novák, Jonsson, Galiová and Holtstam2018). According to a recent estimation (Rönnåsen, Reference Rönnåsen2023), there is still 83,000 t of waste rock present in the mine area, with an average of 0.55% REE.

Fluorbritholite-(Nd) was discovered in a boulder from the old dumps (Fig. 1). All observed occurrences of the new mineral relate to a thin skarn zone situated at the contact between the ‘ophicalcite’ and magnetite ore. Associated minerals in this skarn band include calcite, dolomite, magnetite, lizardite, talc, fluorite, baryte, scheelite and gadolinite-(Nd), accompanied by an allanite-group mineral and a gatelite-group mineral.

Figure 1. Colour image of cut slab of boulder from Malmkärra, GEO-NRM #20220221. Dark spots are mostly serpentine, greyish areas carbonate. Fluorbritholite-(Nd) was obtained from the magnetite-rich part encircled.

The Palaeoproterozoic (Orosirian) Bastnäs-type REE−Fe (± Cu, Bi, Mo, Co and Au) skarn mineralisations in the area originated from reactions involving hot (T > 400°C) magmatic–hydrothermal fluids containing REE−Si−chloro−fluoro ion complexes and primary carbonate (Holtstam and Broman, Reference Holtstam and Broman2002; Holtstam et al., Reference Holtstam, Andersson, Broman and Mansfeld2014; Sahlström et al., Reference Sahlström, Jonsson, Högdahl, Troll, Harris, Jolis and Weis2019).

Appearance and physical properties

Fluorbritholite-(Nd) has an irregular morphology, forming anhedral grains of small size (rarely up 250 μm across, Fig. 2). The colour is brownish pink, with a white streak. The crystals are transparent with a vitreous to greasy lustre. The hardness (Mohs) is estimated as 5, in analogy with fluorbritholite-(Ce). Fluorbritholite-(Nd) is brittle with an uneven or subconchoidal fracture. No parting or cleavage have been observed. Density was not measured because of the minute sample size; the calculated value is 4.92(1) g⋅cm−3. In a petrographic thin section, the mineral is nonpleochroic and uniaxial (–). The refractive indices were not measured due to limited amount of type material (only tiny fragments could be extracted from the one large grain in the type sample). Overall n calc = 1.795 from Gladstone–Dale coefficients (Mandarino, Reference Mandarino1981), very close to the highest available index liquid (1.80).

Figure 2. Back-scattered electron image of fluorbritholite-(Nd) (Fbri-Nd), an allanite-like mineral (Alnl), magnetite (Mag), fluorite (Flr) and dolomite (Dol). The white arrow points to a gatelite-like mineral. Sample GEO-NRM #20220221.

Chemical data

The chemical composition (Table 1) was determined using a JEOL JXA-8230 electron-microprobe operated in wavelength-dispersive mode (20 kV acceleration voltage, 10 nA sample current and 5 μm beam size). The number of spot analyses was 12. Natural and synthetic reference materials were used (Table 1). Measurements involved the following X-ray spectral lines and analysing crystals: Si and Al (Kα, TAP); Y and As (Lα, TAP); P, Ca and Cl (Kα, PETH); La, Ce and Lu (Lα, LIF); Gd and Tb (Lβ, LIF); Pr, Nd, Sm, Dy, Ho and Er (Lβ, LIFL); Yb (Lα, LIFL); F (Kα, LD1); Br (Kα, LIFL). Peak overlaps between various REE and F−Ce were corrected empirically. Thulium was not measured but calculated from chondrite normalisation. Iron, Ti, Al, Mg, U, Pb and Th were sought but found to be below the detection limit. The H2O content was not measured directly because of the small sample volume; it was inferred from structural, chemical and spectroscopic data.

Table 1. Chemical data (in wt.%) for fluorbritholite-(Nd).

*Calculated value from chondrite normalisation ** H2O calculated on the basis of (OH + F + Cl + Br + O) = 13 apfu.

The empirical formula calculated on the basis of 8 cations can be written as: (Ca1.62Nd0.97Ce0.83Y0.52Sm0.30Gd0.23Pr0.17La0.16Dy0.11Er0.03Tb0.03Ho0.01Yb0.01)Σ4.99(Si2.92P0.08As0.01)Σ3.01O12[O0.48F0.26(OH)0.14Cl0.10Br0.02]Σ1.00 The ideal mineral formula is Ca2Nd3(SiO4)3F, which requires (in wt.%) CaO 13.88, Nd2O3 62.45, SiO2 22.31, F 2.35, F ≡ O − 0.99, sum 100.00.

Raman spectroscopy

Micro-Raman measurements were conducted using a Horiba (Jobin Yvon) LabRam HR Evolution instrument on a randomly oriented crystal. The sample was excited with frequency-doubled 532 nm and 785 nm Nd−YAG lasers utilising an Olympus 100 × objective (numerical aperture = 0.9). The lateral resolution of the unpolarised confocal laser beam was in the order of 1 μm. Raman spectra of the sample were collected through two acquisition cycles with single counting times of 45 s. Spectra were generated in the range of 100 to 4000 cm−1 utilising a 600 grooves/cm grating and a thermoelectric-cooled charge-coupled device (CCD). The wavenumber calibration was performed using the 520.7 cm−1 Raman band on a polished silicon wafer with a wavenumber accuracy usually better than 0.5 cm−1.

The spectra are largely overwhelmed by fluorescence. There is a distinct peak at 852 cm−1 in the 532 nm spectrum (Fig. 3), resulting from symmetric stretching of SiO4 groups, and a weak one at 950 cm−1 corresponding to a contribution from the minor PO4 group (ν1 symmetric stretching; cf. Boughzala et al., Reference Boughzala, Salem, Kooli, Gravereau and Bouzouita2008). A sample of P-rich fluorbritholite-(Ce) from Norra Kärr, Sweden, has bands close to these positions (RRUFF no. R070412). A weak signal at 3300–3400 cm−1 in the 785 nm spectrum (Fig. 4) can be related to O−H stretching vibration modes.

Figure 3. Raman spectrum of fluorbritholite-(Nd), obtained with a 532 nm laser.

Figure 4. Raman spectrum of fluorbritholite-(Nd), obtained with a 785 nm laser.

X-ray crystallography

Single-crystal X-ray diffraction

Single-crystal X-ray diffraction data were collected with a Bruker D8 Advance from a 0.060 × 0.055 × 0.045 mm fragment, extracted from the largest available grain previously analysed with the electron microprobe (Fig. 1). The crystal structure of fluorbritholite-(Nd) was refined using SHELXL-2013 (Sheldrick, Reference Sheldrick2008) starting from the atom coordinates of fluorbritholite-(Y) (Pekov et al., Reference Pekov, Zubkova, Chukanov, Husdal, Zadov and Pushcharovsky2011) to R 1 = 0.043 for 704 unique reflections. Crystal parameters and refinement conditions are given in Table 2. Refined atom coordinates, site occupancies and equivalent isotropic displacement parameters are reported in Table 3. Selected interatomic distances are provided in Table 4. The bond valence sums (site populations given in the notes to the table), computed with the parameters given by Brese and O'Keeffe (Reference Brese and O'Keeffe1991), are listed in Table 5. Considering the correspondence in atomic number (‘mean’ atomic number of [REE + Y] = 56.8 in fluorbritholite-(Nd)), we used La (atomic number = 57) as the average REE in the refinement. The crystallographic information file has been deposited with the Principal Editor of Mineralogical Magazine and is available as Supplementary material (see below).

Table 2. Single-crystal data and experimental details.

Note: wR 2 = [Σ w (|F o|2 – |F c|2)2 / Σ w (|F o|4)]½.

Table 3. Refined fractional atom coordinates, equivalent displacement parameters (Å2) and site occupancies in fluorbritholite-(Nd).

Table 4. Selected interatomic distances (Å) in fluorbritholite-(Nd).*

* The O1/Cl1 positions, giving rise to the 7-fold coordinated M2 polyhedron, are mutually present.

Table 5. Bond valence sums (BVS) for fluorbritholite-(Nd) calculated with the parameters of Brese and O'Keeffe (Reference Brese and O'Keeffe1991).*

* M1 = REE0.66Ca0.34; M2 = REE0.70Ca0.30, with REE = Nd0.31Ce0.26Y0.16Sm0.10Gd0.07Pr0.05La0.05

Powder X-ray diffraction

Powder X-ray diffraction data (Table 6) were collected with an Oxford Diffraction Excalibur PX Ultra diffractometer fitted with a 165 mm diagonal Onyx CCD detector and using copper radiation (CuKα, λ = 1.54138 Å) from the same crystal used for the single-crystal experiment. The hexagonal P63/m unit-cell parameters refined from powder data are a = 9.5966(7) Å, c = 6.9816(8) Å and V = 556.83(7) Å3.

Table 6. Observed and calculated X-ray powder diffraction data (d in Å) for fluorbritholite-(Nd).*

* The calculated diffraction pattern is obtained with the atom coordinates reported and the site occupancies given in Table 3 (only reflections with I rel ≥ 3 are listed); the strongest observed Bragg reflections are given in bold.

Results and discussion

Crystal structure

Fluorbritholite-(Nd) has the hexagonal P63/m apatite-type structure, with the general formula M12M23(TO4)3X, in this case based on T = Si plus minor P at the tetrahedrally coordinated sites and polyhedra of M1O9 and M2O6X (M = Ca and REE; X = O, F, OH and Cl). Calcium and the REE are essentially disordered over the M sites. The slight preponderance of REE at M2 agrees with the pattern seen in previously studied LREE-rich fluorbritholite-(Ce) (Oberti et al., Reference Oberti, Ottolini, Ventura and Parodi2001; Zubkova et al., Reference Zubkova, Chukanov, Pekov, Schäfer, Yapaskurt and Pushcharovsky2015), where this preference is even more pronounced. The X anions are in channels parallel to the 6-fold axis. The X site is split; whereas the smaller anions lie on the mirror plane at z = ¼ and Cl is found at distance of 1.1 Å below or above. The site occupancy of the two split sites was fixed during the refinement according to the compositional data obtained from microprobe analyses. The bond valence sums are in good agreement with the site populations and conformable with all lanthanides being trivalent (the presence of tetravalent Ce is highly unlikely as the mineral was deposited in a magnetite-bearing skarn). No symmetry lowering was observed for this sample (cf. Oberti et al., Reference Oberti, Ottolini, Ventura and Parodi2001).

Remarks on nomenclature

The type specimen of fluorbritholite-(Nd) has a surplus of ΣREE atoms (> 3 atoms per formula unit) relative the ideal formula, which must be charge-compensated by introducing a significant REE oxysilicate apatite (CaLn 4[SiO4]3O) component (e.g. Kobayashi and Sakka, Reference Kobayashi and Sakka2014). A similar component was in part inferred for fluorbritholite-(Y) (Pekov et al., Reference Pekov, Zubkova, Chukanov, Husdal, Zadov and Pushcharovsky2011). However, as the sum of monovalent X atoms exceeds the amount of O2−, the name prefix is decided by the dominating F (see Pasero et al., Reference Pasero, Kampf, Ferraris, Pekov, Rakovan and White2010), in accord with the dominant-valency rule (Hatert and Burke, Reference Hatert and Burke2008). The small Br concentration varies to some extent within the grain and is essentially antipathetic to Cl (linear correlation coefficient R 2 = −0.33), but is not correlated with any other element. It is thus ascertained that the sum of monovalent ions is > 0.5 apfu in this crystal. Fluorbritholite-(Nd) is a member of the britholite group of the apatite supergroup (Pasero et al., Reference Pasero, Kampf, Ferraris, Pekov, Rakovan and White2010). It represents the Nd-analogue of fluorbritholite-(Ce) and fluorbritholite-(Y) (Table 7), and belongs to the Strunz group 9.AH.25 (Strunz and Nickel, Reference Strunz and Nickel2001).

Table 7. Comparison of fluorbritholite species.

Geochemical considerations

In the present mineral assemblage, the allanite-like mineral, with large Ce, La, Mg and Fe3+ contents but no measurable F, represents the dominant REE mineral. REE-fluorocarbonates (e.g. bastnäsite) are notably absent, but quite common elsewhere in the REE skarn. As it appears from the chondrite-normalised patterns of the two major REE minerals and a bulk sample from Malmkärra, fluorbritholite-(Nd) is a major sink for Nd in the skarn system (Fig. 5). Similar, nearly flat REE profiles also prevail for gadolinite-subgroup minerals (Škoda et al., Reference Škoda, Plášil, Čopjaková, Novák, Jonsson, Galiová and Holtstam2018), arrheniusite-(Ce) (Holtstam et al., Reference Holtstam, Bindi, Bonazzi, Förster and Andersson2021) and magnesiorowlandite-(Y) (Holtstam and Andersson, Reference Holtstam and Andersson2007), but these minerals occur in much less quantities in the deposits of the Norberg District. In fact, all fluorbritholite samples here (Fig. 6) contain significant Nd (11.5–19.8 wt.% Nd2O3; Holtstam and Andersson, Reference Holtstam and Andersson2007, and this work), with a generally antipathetic relation between Ce and Y (12.8–31.7% Ce2O3 vs. 13.9–0.9% Y2O3). In the Nya Bastnäs deposit (belonging to sub-type 1 in the terminology of Holtstam and Andersson, Reference Holtstam and Andersson2007), the cerite-group minerals are similarly enriched in Nd, however no Nd-dominant member has been found to date.

Figure 5. Chondrite-normalised REE patterns for fluorbritholite-(Nd) (Fbri-Nd), the coexisting allanite-like mineral (Alnl) and the bulk system (obtained with lithium borate fusion and ICP-MS) at Malmkärra. Chondrite values were taken from Boynton (Reference Boynton and Henderson1984).

Figure 6. Triangular plot of Ce:Nd:Y atomic ratios in fluorbritholite samples from the Norberg area, Västmanland, Sweden. Filled circles correspond to the analyses of fluorbritholite-(Nd) from this study. Open symbols refer to data reported by Holtstam and Andersson (Reference Holtstam and Andersson2007).

Compared to the chemical fingerprint of britholite-group minerals, which are predominantly of magmatic–volcanic origin (Della Ventura et al., Reference Della Ventura, Williams, Cabella, Oberti, Caprilli and Bellatreccia1999; Zozulya et al., Reference Zozulya, Lyalina and Savchenko2015; Zubkova et al., Reference Zubkova, Chukanov, Pekov, Schäfer, Yapaskurt and Pushcharovsky2015; Zozulya et al., Reference Zozulya, Lyalina and Savchenko2017; Lorentz et al., Reference Lorenz, Altenberger, Trumbull, Lira, de Luchi, Günter and Eidner2019), the Norberg samples are poor in Ca, despite crystallisation in a carbonate-rich environment (dolomite). The apatite-type substitution, Ca2+ + P5+ ↔ REE3+ + Si4+, is accordingly modest. Interestingly, the Ca-poor britholite samples of the skarn-related Anadol REE ores of the eastern Azov region, Ukraine (Khomenko et al., Reference Khomenko, Rhede, Kosorukov and Strekozov2013), are Nd-rich as well (up to 14 wt.% Nd2O3). The Norberg britholite-group minerals are also practically devoid of the actinides (<0.05 wt.%), which is exceptional. Exploitation of skarn-hosted REE deposits of this type is thus more environmentally friendly (owing to the toxicity and radiation protection aspects related to U and Th) and economically attractive, provided that larger tonnages can be secured.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1180/mgm.2023.45.

Acknowledgements

We thank Matthias Konrad-Schmolke at the Univ. of Gothenburg for providing access to the Raman spectroscopy facility. Sample collection and preliminary analyses were conducted in the course of the governmental directive “Increasing the opportunities for sustainable extraction and recycling of minerals from secondary resources” by the Geological Survey of Sweden. We appreciate the constructive comments and suggestions by two anonymous reviewers and Structures Editor Peter Leverett.

Competing interest

The authors declare none.

Footnotes

Associate Editor: Daniel Atencio

References

Boughzala, K., Salem, E.B., Kooli, F., Gravereau, P. and Bouzouita, K. (2008) Spectroscopic studies and Rietveld refinement of strontium-britholites. Journal of Rare Earths, 26, 483489.CrossRefGoogle Scholar
Boynton, W.V. (1984) Cosmochemistry of the rare earth elements: meteorite studies. Pp. 63114 in: Developments in Geochemistry (Henderson, P., editor). Elsevier, Amsterdam.Google Scholar
Brese, N.E. and O'Keeffe, M. (1991) Bond-valence parameters for solids. Acta Crystallographica, B47, 192197.CrossRefGoogle Scholar
Chakhmouradian, A.R. and Wall, F. (2012) Rare earth elements: minerals, mines, magnets (and more). Elements, 8, 333340.CrossRefGoogle Scholar
Della Ventura, G., Williams, C.T., Cabella, R., Oberti, R., Caprilli, E. and Bellatreccia, F. (1999) Britholite−hellandite intergrowths and associated REE-minerals from the alkali-syenitic ejecta of the Vico volcanic complex (Latium, Italy); petrological implications bearing on REE mobility in volcanic systems. European Journal of Mineralogy, 11, 843854.CrossRefGoogle Scholar
Geijer, P. (1927) Some mineral associations from the Norberg district. Sveriges Geologiska Undersökning, C343, 132.Google Scholar
Geijer, P. (1936) Norbergs berggrund och malmfyndigheter. Sveriges Geologiska Undersökning, Ca24, 1162.Google Scholar
Geijer, P. (1961) The geological significance of the cerium mineral occurrences of the Bastnäs type in central Sweden. Arkiv för Mineralogi och Geologi, 3, 99105.Google Scholar
Geijer, P. and Magnusson, N.H. (1944) De mellansvenska järnmalmernas geologi. Sveriges Geologiska Undersökning, Ca35, 1654.Google Scholar
Goodenough, K.M., Wall, F. and Merriman, D. (2018) The rare earth elements: demand, global resources, and challenges for resourcing future generations. Natural Resources Research, 27, 201216.CrossRefGoogle Scholar
Gu, J., Tang, S. and Chao, G.Y. (1994) A new mineral – fluorbritholite-(Ce). Journal of Wuhan University of Technology, 9, 914.Google Scholar
Hatert, F. and Burke, E.A.J. (2008) The IMA–CNMNC dominant-constituent rule revisited and extended. The Canadian Mineralogist, 46, 717728.CrossRefGoogle Scholar
Holtstam, D. and Andersson, U.B. (2007) The REE minerals of the Bastnäs-type deposits, South-Central Sweden. The Canadian Mineralogist, 45, 10731114.CrossRefGoogle Scholar
Holtstam, D. and Broman, C. (2002) Lanthanide mineralizations of Bastnäs type: overview and new data. GFF, 124, 230231.CrossRefGoogle Scholar
Holtstam, D., Kolitsch, U. and Andersson, U.B. (2005) Västmanlandite-(Ce)-a new lanthanide-and F-bearing sorosilicate mineral from Västmanland, Sweden description, crystal structure, and relation to gatelite-(Ce). European Journal of Mineralogy, 17, 129141.CrossRefGoogle Scholar
Holtstam, D., Andersson, U.B., Broman, C. and Mansfeld, J. (2014) Origin of REE mineralization in the Bastnäs-type Fe−REE−(Cu−Mo−Bi−Au) deposits, Bergslagen, Sweden. Mineralium Deposita, 49, 933966.CrossRefGoogle Scholar
Holtstam, D., Bindi, L., Hålenius, U., Kolitsch, U. and Mansfeld, J. (2017) Ulfanderssonite-(Ce), a new Cl-bearing REE silicate mineral species from the Malmkärra mine, Norberg, Sweden. European Journal of Mineralogy, 29, 10151026.CrossRefGoogle Scholar
Holtstam, D., Bindi, L., Bonazzi, P., Förster, H.J. and Andersson, U.B. (2021) Arrheniusite-(Ce), CaMg[(Ce7Y3)Ca5](SiO4)3(Si3B3O18)(AsO4)(BO3)F11, a new member of the vicanite group, from the Östanmossa Mine, Norberg, Sweden. The Canadian Mineralogist, 59, 177189.CrossRefGoogle Scholar
Holtstam, D., Casey, P., Bindi, L., Förster, H.-J. and Karlsson, A. (2023) Fluorbritholite-(Nd), IMA 2023-001. CNMNC Newsletter 73. Mineralogical Magazine, 87, https://doi.org/10.1180/mgm.2023.44.Google Scholar
Khomenko, V.M., Rhede, D., Kosorukov, O.O. and Strekozov, S.M. (2013) Britholite, cerite and bastnäsite in the Anadol ore occurrence (Eastern Azov area). Mineralogical Journal (Ukraine), 35, 1126.Google Scholar
Kobayashi, K. and Sakka, Y. (2014) Rudimental research progress of rare-earth silicate oxyapatites: their identification as a new compound until discovery of their oxygen ion conductivity. Journal of the Ceramic Society of Japan, 122, 649663.CrossRefGoogle Scholar
Lorenz, M., Altenberger, U., Trumbull, R.B., Lira, R., de Luchi, M.L., Günter, C. and Eidner, S. (2019) Chemical and textural relations of britholite- and apatite-group minerals from hydrothermal REE mineralization at the Rodeo de los Molles deposit, Central Argentina. American Mineralogist, 104, 18401850.CrossRefGoogle Scholar
Mandarino, J.A. (1981) The Gladstone-Dale relationship; Part IV, The compatibility concept and its application. The Canadian Mineralogist, 19, 441450.Google Scholar
Oberti, R., Ottolini, L., Ventura, G.D. and Parodi, G.C. (2001) On the symmetry and crystal chemistry of britholite: New structural and microanalytical data. American Mineralogist, 86, 10661075.CrossRefGoogle Scholar
Pasero, M., Kampf, A.R., Ferraris, C., Pekov, I.V., Rakovan, J. and White, T.J. (2010) Nomenclature of the apatite supergroup minerals. European Journal of Mineralogy, 22, 163179.CrossRefGoogle Scholar
Pekov, I.V., Zubkova, N.V., Chukanov, N.V., Husdal, T.A., Zadov, A.E. and Pushcharovsky, D.Y. (2011) Fluorbritholite-(Y), (Y,Ca,Ln)5[(Si,P)O4]3F, a new mineral of the britholite group. Neues Jahrbuch für Mineralogie Abhandlungen, 2011, 191197.CrossRefGoogle Scholar
Rönnåsen, L. (editor) (2023) Hållbar utvinning och återvinning av metaller och mineral från sekundära resurser. Sveriges Geologiska Undersökning, rapport 2023:1, Uppsala, Sweden, 137 pp.Google Scholar
Sadeghi, M., Arvanitidis, N., Ripa, M., Jonsson, E., Nysten, P., Bergman, T., Söderhielm, J. and Claeson, D. (2019) Rare earth elements distribution, mineralisation and exploration potential in Sweden. Sveriges Geologiska Undersökning, Rapporter och meddelanden, 146, 1184.Google Scholar
Sahlström, F., Jonsson, E., Högdahl, K., Troll, V.R., Harris, C., Jolis, E.M. and Weis, F. (2019) Interaction between high-temperature magmatic fluids and limestone explains ‘Bastnäs-type’ REE deposits in central Sweden. Scientific Reports, 9, 19.CrossRefGoogle ScholarPubMed
Sheldrick, G.M. (2008) A short history of SHELX. Acta Crystallographica, A64, 112122.CrossRefGoogle Scholar
Škoda, R., Plášil, J., Čopjaková, R., Novák, M., Jonsson, E., Galiová, M.V. and Holtstam, D. (2018) Gadolinite-(Nd), a new member of the gadolinite supergroup from Fe−REE deposits of Bastnäs-type, Sweden. Mineralogical Magazine, 82(S1), S133S145.CrossRefGoogle Scholar
Smith, D.G.W. and Nickel, E.H. (2007) A system of codification for unnamed minerals: Report of the subcommittee for unnamed minerals of the IMA Commission on New Minerals, Nomenclature and Classification. The Canadian Mineralogist, 45, 9831055.CrossRefGoogle Scholar
Strunz, H. and Nickel, E.H. (2001) Strunz Mineralogical Tables, Ninth Edition. Schweizerbart'sche Verlagsbuchhandlung, Stuttgart, Germany, 870 pp.Google Scholar
Zozulya, D.R., Lyalina, L.M. and Savchenko, Y.E. (2015) Britholite ores of the Sakharjok Zr–Y–REE deposit, Kola Peninsula: Geochemistry, mineralogy, and formation stages. Geochemistry International, 53, 892902.CrossRefGoogle Scholar
Zozulya, D.R., Lyalina, L.M. and Savchenko, Y.E. (2017) Britholite-group minerals as sensitive indicators of changing fluid composition during pegmatite formation: evidence from the Keivy alkaline province, Kola Peninsula, NW Russia. Mineralogy and Petrology, 111, 511522.CrossRefGoogle Scholar
Zubkova, N.V., Chukanov, N.V., Pekov, I.V., Schäfer, C., Yapaskurt, V.O. and Pushcharovsky, D.Y. (2015) Lanthanum-rich fluorbritholite-(Ce) from young alkaline volcanic rock of Eifel (Germany) and its crystal structure. Cation ordering in britholites. Doklady Earth Sciences, 464, 936939.CrossRefGoogle Scholar
Figure 0

Figure 1. Colour image of cut slab of boulder from Malmkärra, GEO-NRM #20220221. Dark spots are mostly serpentine, greyish areas carbonate. Fluorbritholite-(Nd) was obtained from the magnetite-rich part encircled.

Figure 1

Figure 2. Back-scattered electron image of fluorbritholite-(Nd) (Fbri-Nd), an allanite-like mineral (Alnl), magnetite (Mag), fluorite (Flr) and dolomite (Dol). The white arrow points to a gatelite-like mineral. Sample GEO-NRM #20220221.

Figure 2

Table 1. Chemical data (in wt.%) for fluorbritholite-(Nd).

Figure 3

Figure 3. Raman spectrum of fluorbritholite-(Nd), obtained with a 532 nm laser.

Figure 4

Figure 4. Raman spectrum of fluorbritholite-(Nd), obtained with a 785 nm laser.

Figure 5

Table 2. Single-crystal data and experimental details.

Figure 6

Table 3. Refined fractional atom coordinates, equivalent displacement parameters (Å2) and site occupancies in fluorbritholite-(Nd).

Figure 7

Table 4. Selected interatomic distances (Å) in fluorbritholite-(Nd).*

Figure 8

Table 5. Bond valence sums (BVS) for fluorbritholite-(Nd) calculated with the parameters of Brese and O'Keeffe (1991).*

Figure 9

Table 6. Observed and calculated X-ray powder diffraction data (d in Å) for fluorbritholite-(Nd).*

Figure 10

Table 7. Comparison of fluorbritholite species.

Figure 11

Figure 5. Chondrite-normalised REE patterns for fluorbritholite-(Nd) (Fbri-Nd), the coexisting allanite-like mineral (Alnl) and the bulk system (obtained with lithium borate fusion and ICP-MS) at Malmkärra. Chondrite values were taken from Boynton (1984).

Figure 12

Figure 6. Triangular plot of Ce:Nd:Y atomic ratios in fluorbritholite samples from the Norberg area, Västmanland, Sweden. Filled circles correspond to the analyses of fluorbritholite-(Nd) from this study. Open symbols refer to data reported by Holtstam and Andersson (2007).

Supplementary material: File

Holtstam et al. supplementary material

Holtstam et al. supplementary material 1

Download Holtstam et al. supplementary material(File)
File 447.2 KB
Supplementary material: PDF

Holtstam et al. supplementary material

Holtstam et al. supplementary material 2

Download Holtstam et al. supplementary material(PDF)
PDF 122.2 KB