Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-05-16T09:39:17.197Z Has data issue: false hasContentIssue false

Sulfur-bearing monazite-(Ce) from the Eureka carbonatite, Namibia: oxidation state, substitution mechanism, and formation conditions

Published online by Cambridge University Press:  11 December 2019

Sam Broom-Fendley*
Affiliation:
Camborne School of Mines and the Environment and Sustainability Institute, University of Exeter, Penryn Campus, CornwallTR10 9FE, UK
Martin P Smith
Affiliation:
School of Environment and Technology, University of Brighton, Cockcroft Building, Lewes Road, BrightonBN4 2GJ, UK
Marcelo B Andrade
Affiliation:
São Carlos Institute of Physics, University of São Paulo, PO Box 369, 13560-970, São Carlos, SP, Brazil
Santanu Ray
Affiliation:
School of Environment and Technology, University of Brighton, Cockcroft Building, Lewes Road, BrightonBN4 2GJ, UK
David A Banks
Affiliation:
School of Earth and Environment, University of Leeds, LeedsLS2 9JT, UK
Edward Loye
Affiliation:
Camborne School of Mines and the Environment and Sustainability Institute, University of Exeter, Penryn Campus, CornwallTR10 9FE, UK E-Tech Metals Ltd., Woodlands Grange, Bradley Stoke, BristolBS32 4JY, UK
Daniel Atencio
Affiliation:
Departamento de Mineralogia e Geotectônica, Instituto de Geociências, Universidade de São Paulo, Rua do Lago 562, 05508-080 São Paulo, SP, Brazil
Jonathan R Pickles
Affiliation:
Camborne School of Mines and the Environment and Sustainability Institute, University of Exeter, Penryn Campus, CornwallTR10 9FE, UK
Frances Wall
Affiliation:
Camborne School of Mines and the Environment and Sustainability Institute, University of Exeter, Penryn Campus, CornwallTR10 9FE, UK
*
*Author for correspondence: Sam Broom-Fendley, Email: s.l.broom-fendley@exeter.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

Sulfur-bearing monazite-(Ce) occurs in silicified carbonatite at Eureka, Namibia, forming rims up to ~0.5 mm thick on earlier-formed monazite-(Ce) megacrysts. We present X-ray photoelectron spectroscopy data demonstrating that sulfur is accommodated predominantly in monazite-(Ce) as sulfate, via a clino-anhydrite-type coupled substitution mechanism. Minor sulfide and sulfite peaks in the X-ray photoelectron spectra, however, also indicate that more complex substitution mechanisms incorporating S2– and S4+ are possible. Incorporation of S6+ through clino-anhydrite-type substitution results in an excess of M2+ cations, which previous workers have suggested is accommodated by auxiliary substitution of OH for O2–. However, Raman data show no indication of OH, and instead we suggest charge imbalance is accommodated through F substituting for O2–. The accommodation of S in the monazite-(Ce) results in considerable structural distortion that may account for relatively high contents of ions with radii beyond those normally found in monazite-(Ce), such as the heavy rare earth elements, Mo, Zr and V. In contrast to S-bearing monazite-(Ce) in other carbonatites, S-bearing monazite-(Ce) at Eureka formed via a dissolution–precipitation mechanism during prolonged weathering, with S derived from an aeolian source. While large S-bearing monazite-(Ce) grains are likely to be rare in the geological record, formation of secondary S-bearing monazite-(Ce) in these conditions may be a feasible mineral for dating palaeo-weathering horizons.

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 in any medium, provided the original work is properly cited.
Copyright
Copyright © Mineralogical Society of Great Britain and Ireland 2019

Introduction

The Ce-dominant rare earth element (REE) phosphate mineral monazite-(Ce) is the most commonly occurring REE-bearing mineral and most abundant species of the monoclinic monazite group. It occurs typically as a small accessory phase in granitic and metamorphic rocks, but reaches larger sizes and higher modal abundancies in some carbonatites, pegmatites, alkaline rocks and metamorphic veins (e.g. Bermanec et al., Reference Bermanec, Tibljaš, Gessner and Kniewald1988; Andreoli et al., Reference Andreoli, Smith, Watkeys, Moore, Ashwal and Hart1994; Lehmann et al., Reference Lehmann, Nakai, Höhndorf, Brinckmann, Dulski, Hein and Masuda1994; Wall and Mariano, Reference Wall, Mariano, Jones, Wall and Williams1996; Gonçalves et al., Reference Gonçalves, Lana, Scholz, Buick, Gerdes, Kamo, Corfu, Marinho, Chaves, Valeriano and Nalini2016; Broom-Fendley et al., Reference Broom-Fendley, Wall, Spiro and Ullmann2017; Montel et al., Reference Montel, Razafimahatratra, de Parseval, Poitrasson, Moine, Seydoux-Geuillaume, Pik, Arnaud and Gibert2018; Ntiharirizwa et al., Reference Ntiharirizwa, Boulvais, Poujol, Branquet, Morelli, Ntungwanayo and Midende2018; Slezak and Spandler, Reference Slezak and Spandler2019). Monazite-(Ce) is of considerable importance for both practical and research reasons. It is a major ore mineral for the light (L)REE (and Th) (Chakhmouradian and Wall, Reference Chakhmouradian and Wall2012; Wall, Reference Wall and Gunn2014; Cuney and Kyser, Reference Cuney, Kyser, Cuney and Kyser2015), forming high grade ‘hard rock’ deposits such as Steenkampskraal, South Africa (Andreoli et al., Reference Andreoli, Smith, Watkeys, Moore, Ashwal and Hart1994) and Kangankunde, Malawi (Wall and Mariano, Reference Wall, Mariano, Jones, Wall and Williams1996), as well as forming in secondary environments such as laterites (e.g. Mt Weld, Western Australia; Lottermoser, Reference Lottermoser1990) and heavy mineral sands, both of which are mined. It is also a potential host for long-term storage of radioactive waste (Ewing and Wang, Reference Ewing, Wang, Kohn, Rakovan and Hughes2002; Oelkers and Montel, Reference Oelkers and Montel2008), due to its ability to structurally incorporate a variety of actinide elements and resistance to metamictisation (Seydoux-Guillaume et al., Reference Seydoux-Guillaume, Deschanels, Baumier, Neumeier, Weber and Peuget2018).

Monazite-group minerals are used widely as petrochronometers (Engi, Reference Engi, Kohn, Engi and Lanari2017). Both Th and U are incorporated readily into the monazite structure, while common Pb concentrations are typically low, enabling Th–Pb and, when U contents are sufficiently high, U–Pb age determination using isotopic (e.g. Parrish, Reference Parrish1990) and non-isotopic techniques such as by electron-probe microanalysis (EPMA; Montel et al., Reference Montel, Foret, Veschambre, Nicollet and Provost1996; Williams et al., Reference Williams, Jercinovic and Hetherington2007). Moreover, diffusion of heavy elements is very slow (Cherniak et al., Reference Cherniak, Watson, Grove and Harrison2004; Gardés et al., Reference Gardés, Jaoul, Montel, Seydoux-Guillaume and Wirth2006, Reference Gardés, Montel, Seydoux-Guillaume and Wirth2007), meaning that inherited cores and individual zones in texturally complex monazite may preserve conditions from multiple, temporally distinct, growth events (Hetherington et al., Reference Hetherington, Backus, McFarlane, Fisher, Pearson, Moser, Corfu, Darling, Reddy and Tait2018). Combining geochronological data with other geochemical, isotopic and textural details means that the history of rocks formed over multiple episodes can be unravelled (Engi et al., Reference Engi, Lanari, Kohn, Kohn, Engi and Lanari2017). For example, monazite-(Ce) in equilibrium with xenotime-(Y) (Spear and Pyle, Reference Spear, Pyle, Kohn, Rakovan and Hughes2002) may be used as a thermometer, constraining PTt conditions. Combining δ18O (e.g. Ayers et al., Reference Ayers, Loflin, Miller, Barton and Coath2006) or 143Nd/144Nd (e.g. McFarlane and McCulloch, Reference McFarlane and McCulloch2007) with Th–Pb age determination can trace hydrothermal fluid infiltration and mixing between different melt or fluid sources. Textural features, such as oscillatory and sector zoning (e.g. Cressey et al., Reference Cressey, Wall and Cressey1999) may be attributed to magmatic crystallisation, resorption may indicate the presence of a melt with which monazite-(Ce) is not in equilibrium, while development of mineral inclusions, porosity, or mottled textures may be indicative of fluid-mediated dissolution–reprecipitation and therefore ingress of external fluids (Putnis, Reference Putnis2002, Reference Putnis, Oelkers and Schott2009; Hetherington et al., Reference Hetherington, Backus, McFarlane, Fisher, Pearson, Moser, Corfu, Darling, Reddy and Tait2018).

While rare, the substitution of sulfur into monazite-(Ce) has been described in several localities, predominantly in carbonatites (e.g. Kukharenko et al., Reference Kukharenko, Bulakh and Baklanova1961; Cressey et al., Reference Cressey, Wall and Cressey1999; Bulakh et al., Reference Bulakh, Nesterov, Zaitsev, Pilipiuk, Wall and Kirillov2000; Wall, Reference Wall and Vladykin2004; Enkhbayar et al., Reference Enkhbayar, Sea, Choi, Lee and Batmunkh2016; Prokopyev et al., Reference Prokopyev, Doroshkevich, Ponomarchuk and Sergeev2017; Nikolenko et al., Reference Nikolenko, Redina, Doroshkevich, Prokpyv, Rafozin and Vladykin2018), but also in kimberlites (Chakhmouradian and Mitchell, Reference Chakhmouradian and Mitchell1999), fluid-altered rhyolite (Ondrejka et al., Reference Ondrejka, Uher, Pršek and Ozdín2007) and in metamorphic rocks (Suzuki and Kato, Reference Suzuki and Kato2008; Pršek et al., Reference Pršek, Ondrejka, Bačík, Budzyń and Uher2010; Krenn et al., Reference Krenn, Putz, Finger and Paar2011; Ondrejka et al., Reference Ondrejka, Putiš, Uher, Schmiedt, Pukančík and Konečný2016; Laurent et al., Reference Laurent, Seydoux-Guilaume, Duchene, Bingen, Bosse and Datas2016). Small amounts of sulfur in monazite-(Ce) and monazite-(La) are also sometimes noted during EPMA radiometric age determination, owing to the close proximity of the SKα line to the PbMα lines (Jercinovic and Williams, Reference Jercinovic and Williams2005; Suzuki and Kato, Reference Suzuki and Kato2008), as well as the association of high S contents with increased common Pb (Krenn et al., Reference Krenn, Putz, Finger and Paar2011). Early work predominantly explored substitution mechanisms for incorporating sulfur in monazite-(Ce) (e.g. Chakhmouradian and Mitchell, Reference Chakhmouradian and Mitchell1999; Bulakh et al., Reference Bulakh, Nesterov, Zaitsev, Pilipiuk, Wall and Kirillov2000; Ondrejka et al., Reference Ondrejka, Uher, Pršek and Ozdín2007), but the presence of sulfur is being used increasingly to infer formation environments. Sulfur incorporation is assumed generally to reflect the introduction of S-bearing fluids in relatively acidic and high $f_{O_2}$ conditions (Pršek et al., Reference Pršek, Ondrejka, Bačík, Budzyń and Uher2010; Krenn et al., Reference Krenn, Putz, Finger and Paar2011; Ondrejka et al., Reference Ondrejka, Putiš, Uher, Schmiedt, Pukančík and Konečný2016; Laurent et al., Reference Laurent, Seydoux-Guilaume, Duchene, Bingen, Bosse and Datas2016), possibly at temperatures lower than 400°C based on monazite-(Ce)–xenotime-(Y) thermometry (Krenn et al., Reference Krenn, Putz, Finger and Paar2011). Recently, Laurent et al. (Reference Laurent, Seydoux-Guilaume, Duchene, Bingen, Bosse and Datas2016) showed that it is possible to precisely date S incorporation in monazite-(Ce) and therefore understand the timing of S mobilisation in metamorphic rocks. Their study highlighted the potential to date S-rich domains in monazite-(Ce) from other rock types, such as ore deposits, to further understand the timing of mineralisation. On the basis of co-crystallisation of baryte and the absence of sulfides (e.g. Ondrejka et al., Reference Ondrejka, Uher, Pršek and Ozdín2007, Reference Ondrejka, Putiš, Uher, Schmiedt, Pukančík and Konečný2016), it is generally assumed that monazite-(Ce) only incorporates S as the S6+ ion, and therefore the presence of S-bearing monazite-(Ce) reflects S mobility in oxidising conditions only. However, experimental and ab initio studies have shown that apatite, which has been assumed typically to accommodate sulfur as S6+ only (e.g. Pan and Fleet, Reference Pan, Fleet, Kohn, Rakovan and Hughes2002), is also capable of accommodating more reduced varieties of S such as S4+ and S2– (Konecke et al., Reference Konecke, Fiege, Simon, Parat and Stechern2017a; Kim et al., Reference Kim, Konecke, Fiege, Simon and Becker2017). There has been no direct investigation into the redox state of S in monazite-(Ce) to investigate its capability to accommodate S in different oxidation states. Moreover, the substitution mechanism incorporating S in monazite-(Ce) remains incompletely understood. These two less-constrained factors reduce our confidence for inferring the formation environment from the presence of S in LREE phosphate species of the monazite group.

This study presents the first description of S-bearing monazite-(Ce) from the Eureka carbonatite, Namibia. Eureka exhibits the largest grains of S-bearing monazite-(Ce) yet documented (>0.5 mm) with an SO3 concentration within uncertainty of the highest published SO3 contents [11.23 wt.% SO3 ≡ 0.294 S6+ atoms per formula unit (apfu); cf. Pršek et al., Reference Pršek, Ondrejka, Bačík, Budzyń and Uher2010]. The large size of the S-bearing monazite-(Ce) has enabled analysis by X-ray photoelectron spectroscopy (XPS), and we use this technique to present the first direct data on the oxidation state of S in monazite-(Ce). We also expand on previous work investigating S substitution in monazite-(Ce), moving towards a complete, charge-balanced, understanding of the S substitution mechanism. Lastly, on the basis of the crystallisation environment, we show that sulfur can be incorporated into monazite-(Ce) during a protracted weathering process, opening up the possibility of directly dating palaeo-weathering horizons.

Geology of the Eureka carbonatite

The Eureka carbonatite is located at 22.044°S, 15.254°E on the property of Eureka Farm 99, near Usakos, Namibia (Fig. 1a). Quaternary gravels and calcrete obscure much of the exposure, but remnant mineral exploration trenches from the late 1980s, and modern trenching by E-Tech Metals, facilitates easy sampling. The carbonatites comprise a series of at least four steeply-dipping (>75°) monazite-(Ce)-bearing dolomite–carbonatite dykes and at least three monazite-(Ce)-poor dolomite carbonatites and sövites, all striking roughly parallel, (Fig. 1b; von Knorring and Clifford, Reference von Knorring and Clifford1960; Dunai, Reference Dunai1989). The width of the dykes is typically 1–2 m, but may locally reach up to 7 m. The dykes intrude parallel to the foliation of intensely folded Damaran schists and quartzites of the Etusis formation (Miller, Reference Miller and Miller1983, Reference Miller2008). An 87Sr/86Sr isotope study supports a mantle origin, with minor crustal input (Dunai et al., Reference Dunai, Stoessel and Ziegler1989).

Fig. 1. Location map (a) and geological sketch map (b) of the Eureka carbonatite dykes, with pit and sample locations. Host rocks are quartzite and schists of the Etusis formation. Geological map grid is UTM 33S, WGS 1984 datum, redrawn and georeferenced from Dunai (Reference Dunai1989).

A distinct feature of the dykes at Eureka is the abundance and size of monazite-(Ce) in the carbonatite. Grains reach up to 30 cm in size and have been divided by Dunai (Reference Dunai1989) by morphology, where (more abundant) larger grains are typically euhedral to subhedral while smaller grains are lenticular and display resorption features. Monazite-(Ce) from Eureka has been dated, by U–Pb, as 500 ± 20 Ma (Burger et al., Reference Burger, von Knorring and Clifford1965) and 548 ± 4 Ma (Ragettli et al., Reference Ragettli, Hebeda, Signer and Wieler1994; Gonçalves et al., Reference Gonçalves, Lana, Scholz, Buick, Gerdes, Kamo, Corfu, Rubatto, Wiedenbeck, Nalini and Oliveira2018).

Extensive trenching indicates that the area is weathered to at least 2 m and carbonatite dykes are locally replaced by an assemblage of secondary carbonate minerals, gypsum, serpentine(?), iron oxide/hydroxides, clay and minor fluorite. Locally, the carbonatite is intensely silicified; silicification is most prominent in the north-western area of the intrusion (Zone 2 of Dunai, Reference Dunai1989), with recent drilling indicating that silicification extends to a maximum probable depth of ~5 m. In these rocks, the carbonate matrix has been completely replaced by iron-bearing, rust-brown chalcedony (Fig. 2). Remnant monazite-(Ce) grains in the silicified rocks are typically, although not always, rimmed by a thin (up to 5 mm) alteration halo of S-bearing monazite-(Ce), which is the subject of this paper (Fig. 2). The local area is commonly capped by a layer of calcrete ~30 cm thick.

Fig. 2. Example of monazite-(Ce)-bearing (Mnz) silicified carbonatite from Eureka. Note the ~500 μm S-bearing monazite-(Ce) rims (S-Mnz) around the monazite-(Ce) grains.

Samples and methods

Representative silicified carbonatite samples were collected from Pit 2, in the north-western zone of the deposit (Fig. 1b), where silicification is most intense. Petrographic observations were undertaken using conventional polarising microscopy and back-scattered electron (BSE) imaging. BSE images and maps from energy-dispersive spectroscopy (EDS) were obtained using a FEI Quanta 650 FEG SEM hosted at the Environment and Sustainability Institute, University of Exeter.

Electron probe microanalyses were carried out at Camborne School of Mines on a JEOL JXA-8200 Superprobe, using a defocussed 10 μm, 15 kV, 50 nA beam. Peak counting times were 20 s for S, Fe, P, Ca and Sr; 30 s for F, Na, Ba and Ce; 40 s for Si, Nd, La, Pr and Sm; 60 s for Th; 100 s for Y; and 120 s for Gd. Background count times were half those of the peak value. X-ray counts were converted to wt.% oxide using the in-built JEOL CITZAF correction program. Empirical interference corrections were performed for the REE following Williams (Reference Williams, Jones, Wall and Williams1996). Correction factors were calculated by measuring the interferences observed on individual REE reference standards. Commercial (Astimex) natural minerals standards were used for all elements except for Th and the REE, where Astimex Th metal and REE–Si–Al–Ca–O glasses from the University of Edinburgh, were used. Representative detection limits are included in Supplementary Table S1.

Laser ablation inductively-coupled plasma mass spectrometry (LA-ICPMS) was carried out at the University of Leeds, using a Lambda Physik 193 nm ArF Excimer laser coupled to an Agilent 7500c ICPMS. This was run in reaction mode with 2.5 ml/min H2 in the reaction cell. The laser utilised a 10 Hz, 25 μm spot for ~30 s of ablation time. Typical fluence was 10 J cm–2. Isotopes analysed were: 23Na, 28Si, 31P, 32S, 40Ca, 47Ti, 51V, 56Fe, 75As, 88Sr, 89Y, 90Zr, 93Nb, 65Mo, 137Ba, 139La, 140Ce, 141Pr, 147Sm, 153Eu, 157Gd, 159Tb, 163Dy, 165Ho, 166Er, 169Tm, 172Yb, 175Lu, 178Hf, 208Pb, 232Th and 238U. Data were processed using Iolite software (version 2.5; Paton et al., Reference Paton, Hellstrom, Paul, Woodhead and Hergt2011). Neodymium contents from the EPMA were used as the internal standard value and NIST SRM 610 was used as the external standard. NIST 612 and 614 glasses were utilised as secondary standards, and the concentrations of all analysed elements were within 10% of the standard values, with most within a <5% window.

Raman spectra of S rich and regular monazite-(Ce) were recorded at the University of São Paulo, using a LabRAM HR Evolution micro-Raman system working in back-scattering geometry, using a solid-state laser with a frequency of 633 nm and, equipped with 1800 gr/mm gratings, 1 cm–1 spectral resolution and a liquid-nitrogen-cooled CCD detector.

X-ray photoelectron spectroscopy depth profile measurements were conducted using ESCALAB 250xi XPS (Thermo Scientific, UK) with a MAGCIS™ Dual Beam Ion Source. Analysis was carried out using the selected area analysis mode with a nominal width of analysis of 200 μm and monochromated AlKα X-rays at 1486.6 eV. The MAGCIS source can generate both monatomic ion beams for profiling inorganic materials and also cluster ions for organic layer profiling. In this present case on depth profiling monazite-(Ce), the gun was selected in monoatomic mode with 3 keV ion energy and a raster size of 0.5 mm. The sputtering cycle was kept at 120 s each time. The charge neutraliser and X-ray source were only used during the acquisition of spectra, both being turned off during the sputtering cycle. Survey (wide) scans (step size 1 eV, pass energy 150 eV, dwell time 50 ms, number of scans 15) and narrow scans (step size 0.1 eV, pass energy 40 eV, dwell time 100 ms, number of scans 25) of the S 2p (binding energy, BE ≈ 164 eV), C 1s (BE ≈ 285 eV), O 1s (BE ≈ 531 eV), Ca 2p (BE ≈ 342 eV), P 2p (BE ≈ 132 eV), Y 3d (BE ≈ 153 eV), Ce 3d (BE ≈ 884 eV), La 3d (BE ≈ 835 eV), Nd 4d (BE ≈ 120 eV) and Fe 2p (BE ≈ 717 eV) regions were acquired. Data analysis were carried out using Thermo Avantage software version 5.952.

There is a possibility that redox reactions could be induced in the sample as it is sputtered with a mono-atomic argon beam for depth profiling. In order to fully investigate variation in the redox state of elements it was necessary to demonstrate that there was no redox interaction with the ion beam during surface cleaning. A single crystal of apatite from the Slyudjanka mica deposit (Voskoboinikova, Reference Voskoboinikova1938), Siberia, Russia, sampled at a depth of 120 m to avoid weathering effects (provided by J. Kynický) was rastered with the Ar+ beam over 31 cycles with narrow scan XPS collected after each cycle. There was no discernible chemical shift in Ca or P 2p electron binding energies (Supplementary Fig. S1A–B). A similar process was carried out on reagent grade FeSO4nH2O over 10 raster cycles, again with no discernible chemical shift in the position of the S 2p electron binding energy peak (Supplementary Fig. 1C). From this we conclude that the Ar+ rastering process used for surface cleaning and depth profiling preserves robust redox state information in phosphates and for S species. Tests on homogenous areas of monazite-(Ce) suggest Ce3+ may oxidise to Ce4+ under the Ar+ beam, and so Ce narrow scans were analysed following polishing of the sample with no sputtering.

Results

Petrography

Silicified carbonatite samples are composed principally of amorphous silica phases, including opal and chalcedony, with minor hematite, calcite and quartz veins. The chalcedony contains abundant inclusions of iron-oxide minerals which and locally follow remnant cleavage planes, intersecting at 60° and 120°, after the original dolomite crystals (Fig. 3a). Small (25–50 μm) anhedral celestine grains are also distributed randomly within the amorphous silica. Both the iron and strontium from these minerals is likely to be sourced from the breakdown of the original dolomite, which is present in un-silicified samples and contains up to 10 wt.% FeO and 3 wt.% SrO (unpublished data, Broom-Fendley, 2019).

Fig. 3. BSE images (ad, f) and EDS maps (e, gj) of silicified carbonatite and monazite-(Ce) from Eureka. (a) Planes of hematite (Hem), intersecting at ~120°, cemented by chalcedony (Cdy) and local, anhdedral, celestite (Cls) grains. (b) Monazite-(Ce) (Mnz) with a large S-bearing monazite-(Ce) rim (S-Mnz), hosted in chalcedony. (c) Close-up of monazite-(Ce)/S-bearing monazite-(Ce) boundary, showing symplectic texture between calcite (Cal) and S-bearing monazite-(Ce). (d) Close-up of monazite and S-bearing monazite rods, showing pore formation at the monazite-(Ce)/S-bearing monazite-(Ce) boundary, with (e) showing changes in the S content of the S-bearing monazite over a small area. (f) Example of cross-cutting chalcedony and hematite veins, through monazite-(Ce) and S-bearing monazite-(Ce). (gj) EDS maps demonstrating heterogeneous distribution of Ca and S in the S-bearing monazite (gh), the presence of Ca (g) and Fe (i) inclusions in the S-bearing monazite assemblage, and the presence of chalcedony and hematite veins. Larger circles correspond to sites of XPS analysis, while smaller circles are areas of LA-ICPMS analysis. Numbering corresponds to analytical locations and data in Tables 2, S2, and Fig. 5.

Monazite-(Ce) in the silicified rocks is texturally similar to large monazite-(Ce) grains in un-silicified rocks. Common features include similar roundness, size (between ~0.1 and ~7 cm), orange colour, and the presence of hematite and rounded dolomite inclusions. However, monazite-(Ce) from the un-silicified samples has a distinct rim of S-bearing monazite-(Ce), up to 500 μm wide (e.g. Fig. 3bh), and is also commonly cross-cut by later silica and hematite veins (Fig. 3f,i,j). The S-bearing monazite-(Ce) rim is porous and heterogeneous, with iron-oxide minerals (probably hematite) and calcite making up the rest of the assemblage; these non-phosphate minerals are complexly intermixed (Fig. 3c, fj). At the margin of monazite-(Ce) grains, S-bearing monazite-(Ce) occurs as rounded, rod-like, globules which are interconnected by thin <5 μm veins, hosted in calcite, akin to a symplectic texture (Fig. 3ce). The abundance of calcite intermixed with the S-bearing monazite-(Ce) decreases with increasing distance away from the core monazite-(Ce) grain, reaching a point where S-bearing monazite-(Ce) forms a seemingly homogeneous phase (Fig. 3gh). Locally, S-bearing monazite-(Ce) is intermixed with Fe-oxide minerals (e.g. Fig. 3i).

Monazite-(Ce) composition

Representative EPMA and LA-ICPMS data for monazite-(Ce) and S-bearing monazite-(Ce) from Eureka are shown in Table 1, with the full dataset available in Supplementary Tables S1 and S2. Analysis locations are indicated on Fig. 3. The composition of monazite-(Ce) formed during carbonatite crystallisation is typical, in terms of its relatively low ThO2 and UO2 contents, for monazite-(Ce) from such rocks (cf. Chen et al., Reference Chen, Honghui, Bai and Jiang2017). However, chondrite-normalised REE contents exhibit a small negative Eu and slightly larger negative Y anomaly (Fig. 4), the former of which is atypical of carbonatite-derived minerals. S-bearing monazite-(Ce) has similar concentrations of ThO2 and SiO2 to the monazite-(Ce) grains, as well as the same Eu and Ce anomalies (Fig. 4). However, it has notably different SO3 ( ≈ 9.6 wt.%), CaO ( ≈ 6.6 wt.%), SrO ( ≈ 4.0 wt.%), FeO ( ≈ 1.0 wt.%), F ( ≈ 0.8 wt.%) and Na2O ( ≈ 0.2 wt.%) contents, well in excess of the EPMA detection limit (e.g. Fig. 3gh). Trace elements such as U ( ≈ 210 ppm), V ( ≈ 79 ppm), Zr ( ≈ 22 ppm), Nb ( ≈ 10 ppm), Mo ( ≈ 45 ppm) and Ba ( ≈ 159 ppm) are also enriched relative to the monazite-(Ce) grains (all below detection except U, ≈ 80 ppm, Ba, ≈ 7 ppm). Moreover, S-bearing monazite-(Ce) is relatively HREE-enriched, and depleted in La and Ce contents, compared to grains which grew during carbonatite crystallisation.

Fig. 4. Chondrite-normalised REE distribution of monazite-(Ce) and S-bearing monazite-(Ce) from Eureka. Chondrite values after McDonough and Sun (Reference McDonough and Sun1995).

Table 1. Representative compositions of monazite-(Ce) and S-bearing monazite-(Ce) from Eureka (Sample SoS_63c).*

* Blank cells denote analyses below the EPMA detection limit; Y, Dy and U are below detection in all samples, Ba below detection in all but 2 S-bearing samples. Full dataset in Supplementary Table S1.

Table 2. Representative trace-element data (in μg/g) of monazite-(Ce) and S-bearing monazite-(Ce) from Eureka (Sample SoS_63c).*

* Blank cells denote analyses below the LA-ICPMS detection limit. Full dataset in Supplementary Table S2.

** 56Fe contents >15,000 μg/g may be contaminated with inclusions of an Fe-oxide mineral.

Oxidation state of sulfur

The locations of XPS analyses are shown in Fig. 3. Survey scans from binding energies of 0–1350 eV are shown in Fig. 5. The presence of S is identified clearly in the rim monazite-(Ce), as well as elevated HREE contents, Ca, and F (Fig. 5a,b). Narrow scans were obtained from the Y 3d peak as a proxy for the heavy rare earth elements (HREE; Fig. 5b), and sulfur peaks which vary between 160–170 eV, depending on their bonding environment. The peak position at 154 eV is characteristic of Y–O bonding in compounds (Mesarwi and Ignatiev, Reference Mesarwi and Ignatiev1993). Deconvolution of the S 2p spectra shows that S is present dominantly as SO42– (BE 168.14 eV; Yu et al., Reference Yu, Liu, Wang and Chen1990), but with a component of S2– (BE 160.25 eV; Vasquez, 1991) and SO32– (BE 166.34 eV; Abraham and Chaudhri, Reference Abraham and Chaudhri1986) (Fig. 5c).

Fig. 5. (a) X-ray photoelectron binding energy spectra of monazite-(Ce) from sites shown in Fig. 3. Note the HREE, S and Ca peaks present in the S-bearing monazite-(Ce). (b) Narrow scan of the Dy, Si, Y and S peaks. (c) Narrow scan of the S 2p binding energy peak indicating sulfur present as sulfite, sulfate and sulfide structurally bound in monazite-(Ce). Binding-energy peak positions from ThermoScientific, xpssimplified.com (accessed 2017), NIST XPS database (accessed 2017), Yu et al., (Reference Yu, Liu, Wang and Chen1990), Vasquez (Reference Vasquez1991) and Abraham and Chaudhri (Reference Abraham and Chaudhri1986).

Raman spectroscopy

The Raman spectra of S-rich and regular monazite-(Ce) is shown in Fig. 6. Band assignments of monazite-(Ce) have been proposed previously by Begun et al. (Reference Begun, Beall, Boatner and Gregor1981), Silva et al. (Reference Silva, Ayala, Guedes, Paschoal, Moreira, Loong and Boatner2006) and Lenz et al. (Reference Lenz, Nasdala, Talla, Hauzenberger, Seitz and Kolitsch2015). Tentative assignments of major Raman bands can be grouped into three different regions: 900–1100 cm–1 are attributable to the stretching of the PO4 tetrahedron; 450–700 cm–1 originate from bending vibrations within the PO4 group; and <450 cm–1 are related to lattice modes (Supplementary Table S3). No OH stretches were detected above 2900 cm–1 in the Raman spectra of the samples. However, the 633 nm laser excites an emission of Nd3+ (4f5/2 → 4l9/2 transition), which corresponds to bands in the range 3000–3600 cm–1.

Fig. 6. (a) Raman spectra of monazite-(Ce) and S-bearing monazite-(Ce) from sample SoS_63c. (b) Enlarged version of the above, demonstrating peak broadening and shifting.

The Raman spectra of S-rich and regular monazite-(Ce) are broadly analogous, but it is possible to observe some differences. In particular, between 900 and 1100 cm–1, there is an intense, sharp band at 968 cm–1, related to symmetric stretching, which has a shoulder at 991 cm–1 and a small band at 1056 cm–1, related to antisymmetric stretching for regular monazite-(Ce). In contrast the bands at 968, 991 and 1056 cm–1 are broadened significantly and shifted to 973, 1009 and 1074 cm–1 for S-rich monazite-(Ce) (Fig. 6b).

Discussion

Substitution of minor elements into monazite-(Ce)

Monazite-(Ce) is monoclinic (space group P21/n), comprising equal ratios of PO4 tetrahedra and REEO9 polyhedra (Ni et al., Reference Ni, Hughes and Mariano1995). The monazite-(Ce) structure differs from the tetragonal REE phosphate, xenotime-(Y), by a 2.2 Å offset along the [010] plane, allowing space for the larger LREE atoms, and leading to an additional REE–O bond compared with the 8-fold REEO site in xenotime (Ni et al., Reference Ni, Hughes and Mariano1995). In contrast to xenotime-(Y), the composition of monazite-(Ce) varies considerably in natural samples (e.g. Förster, Reference Förster1998), which might be a consequence of the irregular REE–O bond distances from the 9-fold coordination (Beall et al., Reference Beall, Boatner, Mullica and Milligan1981). In addition to isomorphous REE 3+ substitution, the 9 fold site also commonly accommodates other large cations, such as Th4+, U4+ and M2+ (e.g. Ca2+, Sr2+, Pb2+, Ba2+) with charge-balance achieved typically by the cheralite or huttonite substitution:

(1)$${\rm Cheralite\ substitution\colon \ 2C}{\rm e}^{{3+}} \leftrightarrow {\rm T}{\rm h}^{{4 + }}{\rm}+{\rm C}{\rm a}^{{2+}}$$
(2)$${\rm Huttonite\ substitution\colon \ }{\rm P}^{{5+}}+ REE^{{3 + }} \leftrightarrow {\rm S}{\rm i}^{{4 + }}{\rm}+{\rm T}{\rm h}^{{4 + }}$$

The above two mechanisms account for the incorporation of minor elements in monazite-(Ce) grains (without S) at Eureka. The similar number of Th4+ and Ca2+ + Sr2+ + Fe2+ apfu suggests that cheralite substitution predominantly accommodates the Th in this phase, with a minor huttonite component. A minor huttonite component can also predominantly account for Th4+ in the S-bearing monazite-(Ce), as the number of Si4+ apfu is equal, or greater than the Th4+ apfu in this phase.

Sulfur substitution in monazite-(Ce) is relatively uncommon and, consequently, relatively understudied. Early work by Kukharenko et al. (Reference Kukharenko, Bulakh and Baklanova1961) suggested that S substitutes into the tetragonal PO4 site via a coupled substitution with anhydrite (also termed ‘clino-anhydrite’ and ‘anhydrite-celestite’ exchange by subsequent workers), based on a positive correlation between S and Ca + Sr cations:

(3)$$\lpar {{\rm Sr\comma \;Ca}} \rpar ^{{2 + }}+ {\rm S}^{{6 + }} \leftrightarrow REE^{3 + } + {\rm P}^{{5 + }}$$

An isomorphous substitution with Si has also been suggested (Williams et al., Reference Williams, Jercinovic and Hetherington2007):

(4)$${\rm S}^{{6 + }}{\rm}+{\rm S}{\rm i}^{{4 + }} \leftrightarrow {\rm \;2}{\rm P}^{{5 + }}$$

Narrow XPS scans of the S peaks clearly demonstrate that sulfur is predominantly incorporated in monazite-(Ce) as sulfate (Fig. 5bc). A positive correlation between Sr + Ca and S in the S-bearing monazite-(Ce) at Eureka is a strong indication that clino-anhydrite exchange (Fig. 7) accounts for the S content in this example, while low Si and Th contents suggest substitutions involving these elements do not play a role (Chakhmouradian and Mitchell, Reference Chakhmouradian and Mitchell1999; Bulakh et al., Reference Bulakh, Nesterov, Zaitsev, Pilipiuk, Wall and Kirillov2000; Wall, Reference Wall and Vladykin2004; Ondrejka et al., Reference Ondrejka, Uher, Pršek and Ozdín2007, Reference Ondrejka, Putiš, Uher, Schmiedt, Pukančík and Konečný2016; Pršek et al., Reference Pršek, Ondrejka, Bačík, Budzyń and Uher2010; Krenn et al., Reference Krenn, Putz, Finger and Paar2011; Laurent et al., Reference Laurent, Seydoux-Guilaume, Duchene, Bingen, Bosse and Datas2016). The capability of anhydrite to exist in the monoclinic crystal system, albeit at high pressures, gives credence to the possibility of clino-anhydrite substitution (Kahn, Reference Kahn1975; Borg and Smith, Reference Borg and Smith1975; Crichton et al., Reference Crichton, Parise, Antao and Grzechnik2005; Bradbury and Williams, Reference Bradbury and Williams2009). However, clino-anhydrite exchange cannot be the sole mechanism for the incorporation of sulfate, as it results invariably in an excess of M 2+ cations (Fig. 7a). This excess is not accommodated by cheralite substitution as there is insufficient Th4+ to balance the M 2+ excess (Fig. 7b). To accommodate this imbalance, Chakhmouradian and Mitchell (Reference Chakhmouradian and Mitchell1999) suggested the presence of an auxiliary substitution mechanism:

(5)$$REE^{3 + } + {\rm O}^{{\rm 2\ndash }} \leftrightarrow \lpar {{\rm Ca\comma \;Sr}} \rpar ^{{\rm 2+}} +{\rm O}{\rm H}^{\rm \ndash }$$

The absence of an OH peak in Raman data for the S-bearing monazite-(Ce) from Eureka (Fig. 6), however, means the above charge-balancing substitution is not a viable mechanism to accommodate the M 2+ excess. Moreover, our analytical totals from EPMA are within uncertainty of 100% (Table 1), rather than <100% as might be expected if OH substitution is taking place. We, therefore, propose that OH does not accommodate the M 2+ excess and, instead, we suggest an alternative auxiliary substitution where excess M 2+ cations are countered by F substituting for oxygen:

(6)$$REE^{3 + } + {\rm O}^{{\rm 2\ndash }} \leftrightarrow \lpar {{\rm Ca\comma \;Sr}} \rpar ^{{2+ }}+ {\rm F}^{\rm \ndash }$$

The additional charge-balancing of F considerably, although not entirely, reduces the M 2+ excess (Fig. 7c). The F content of monazite-(Ce) is not commonly analysed, but F concentrations up to 0.8 wt.% have been reported from Steenkampskraal monazite-(Ce) (Andreoli et al., Reference Andreoli, Smith, Watkeys, Moore, Ashwal and Hart1994). Moreover, F has been suggested to account for small excesses in Si after cheralite- and huttonite-type substitutions have been accounted for (Williams et al., Reference Williams, Jercinovic and Hetherington2007). Few data are available where both SO3 and F contents are reported and above the EPMA limit of detection, although monazite-(Ce) from the Tomtor, Khaluta and Huanglongpu carbonatites, and the Tisovec–Rejkovo rhyolite, are notable exceptions (Doroshkevich et al., Reference Doroshkevich, Ripp and Moore2001; Ondrejka et al., Reference Ondrejka, Uher, Pršek and Ozdín2007; Lazareva et al., Reference Lazareva, Zhmodik, Dobretsov, Tolstov, Shcherbov, Karmanov, Gerasimov and Bryanskaya2015; Chen et al., Reference Chen, Honghui, Bai and Jiang2017). In these instances, however, the proportion of Ca + Sr and S + F are not balanced, indicating that other mechanisms, such as co-substitution with SiO32– may be accommodating F (Williams et al., Reference Williams, Jercinovic and Hetherington2007). Nonetheless, in the absence of a better explanation to accommodate the M 2+ excess, we place considerable weight onto an auxiliary F substitution mechanism to accommodate charge imbalance.

Structural distortion in S-bearing monazite-(Ce)

The XPS binding energy spectra show that sulfur is present predominantly in S-bearing monazite-(Ce) as sulfate, as expected from the correlation between S and Ca + Sr contents. However, the X-ray photoelectron spectra also exhibits a minor additional peak for sulfide (S2–), and peak deconvolution indicates a small overlap with sulfite (S4+; Fig. 5c). Tests on iron sulfate show that sulfur does not undergo redox reaction under the Ar+ beam, and these peaks are significantly above background, so do not represent artefacts. The possibility of contamination of sulfite and sulfide from other phases is also unlikely as no other S-bearing phases are present in association with the S-bearing monazite-(Ce) (Fig. 3e, h). We cannot discount entirely the possibility that such phases may be present as nano-inclusions (e.g. Laurent et al., Reference Laurent, Seydoux-Guilaume, Duchene, Bingen, Bosse and Datas2016). However, 32S and 56Fe contents from LA-ICPMS analyses shows no clear correlation, indicating that pyrite is not present in the analyses (Supplementary Fig. S2; Fig. 3ij).

Accommodating sulfur in monazite-(Ce) in oxidation states other than S6+ is somewhat challenging. In terms of charge-balance, S4+ could replace the Si4+ component in a pseudo-huttonite substitution, or as a coupled substitution with P5+:

(7)$${\rm S}^{{4 + }}{\rm}+{\rm T}{\rm h}^{{4 + }} \leftrightarrow {\rm P}^{{5 + }} + REE^{3 + }$$
(8)$${\rm S}^{{6 + }}{\rm}+ {\rm S}^{{4 + }} \leftrightarrow {\rm \;2}{\rm P}^{{5 + }}$$

The former option can be ruled out in this instance as Th4+ contents are very low, and can be accommodated by huttonite substitution. For sulfide, direct S2– for O2– may accommodate S2–:

(9)$${\rm S}^{{\rm 2\ndash }} \leftrightarrow {\rm O}^{{\rm 2\ndash }}$$

Recent micro X-ray absorption near-edge structure analyses of both laboratory-synthesised and natural samples have indicated that sulfide and sulfite can occur in apatite (Konecke et al., Reference Konecke, Fiege, Simon, Parat and Stechern2017a, Reference Konecke, Fiege, Simon and Holtz2017b; Brounce et al., Reference Brounce, Boyce, McCubbin, Humphreys, Reppart, Stolper and Eiler2019; Sadove et al., Reference Sadove, Konecke, Fiege and Simon2019), which helps to support the concept that these species may occur in monazite-(Ce) as well. Ab initio studies of apatite support the mechanism for coupled substitution of S6+ and S4+ for P5+ (Kim et al., Reference Kim, Konecke, Fiege, Simon and Becker2017). Sulfite occurs as a trigonal pyramid with a lone electron pair replacing one of the O atoms in the substituted PO4 (or SO4). The stability of sulfite substitution in apatite is affected by the presence of the anion column, which repels the lone pair of electrons in the sulfite trigonal pyramid. The highest stability occurs in chlorapatite owing to increased distance between the Cl anion and the electron pair, relative to OH and F (Kim et al., Reference Kim, Konecke, Fiege, Simon and Becker2017). The absence of the anion column in monazite-group minerals therefore suggests that sulfite substitution may be more stable than in apatite. Conversely, however, S2– is accommodated in the anion site in apatite (Kim et al., Reference Kim, Konecke, Fiege, Simon and Becker2017), and no such site is available in monazite-(Ce). Substituting S2– for O2– is difficult to reconcile with the monazite-(Ce) structure owing to the considerable size difference between these two anions. All three of the above substitutional mechanisms would result in mis-matched geometry, and would probably result in local defects and structural distortion. Moreover, these mechanisms also exacerbate the problem of M 2+ excess, as the charge-balancing contribution of S is reduced.

While we lack evidence to support conclusively the three substitution mechanisms suggested above, stretches in the Raman bands at 968, 991 and 1056 cm–1 do indicate changes in the P–O bond distance caused by structural distortion (Fig. 6). The sulfate ion is ~70% of the size of the phosphate ion (Shannon, Reference Shannon1976), suggesting the unit cell of monazite-(Ce) contracts as a result of increasing degrees of exchange with clino-anhydrite, which could potentially accommodate larger ions to compensate. Such a contraction is supported by the data of Bulakh et al. (Reference Bulakh, Nesterov, Zaitsev, Pilipiuk, Wall and Kirillov2000) who demonstrated that the unit cell volume of monazite-(Ce) with ~3% SO3 is ~2% lower than that of monazite-(Ce) with no S substitution. The smaller size of the sulfate ion, in comparison with the substituted phosphate, may also help explain the slightly elevated Mo and V contents, as both molybdate and vanadate ions form substantially larger tetrahedra than PO4. Ondrejka et al. (Reference Ondrejka, Uher, Pršek and Ozdín2007) noted elevated SO3 contents in As-bearing monazite-(Ce)/gasparite-(Ce) which, owing to the similar size of the arsenate, molybdate and vanadate ions, may be analogous. Similarly, we suggest that the slightly elevated contents of HREE and Zr are a consequence of this structural distortion, rather than reflecting any significant change in the formation environment or REE fractionation during transport.

Crystallisation environment of S-bearing monazite-(Ce) at Eureka

Sulfur-bearing monazite-(Ce) is relatively rare in carbonatites although this may, in-part, result from S not being included routinely in a monazite-(Ce) EPMA routine. A compilation of monazite-(Ce) compositions from different carbonatites was presented by Chen et al. (Reference Chen, Honghui, Bai and Jiang2017). The majority of these data have <1% SO3, in addition, published occurrences with monazite-(Ce) containing >1% SO3 are limited to Nkombwa, Zambia (Wall, Reference Wall and Vladykin2004); Vuoriyarvi and Seligdar, Russia (Kukharenko et al., Reference Kukharenko, Bulakh and Baklanova1961; Bulakh, Reference Bulakh, Nesterov, Zaitsev, Pilipiuk, Wall and Kirillov2000; Prokopyev et al., Reference Prokopyev, Doroshkevich, Ponomarchuk and Sergeev2017); and Mushgai Khudag, Mongolia (Enkhbayar et al., Reference Enkhbayar, Sea, Choi, Lee and Batmunkh2016; Nikolenko et al., Reference Nikolenko, Redina, Doroshkevich, Prokpyv, Rafozin and Vladykin2018). In all of these occurrences, monazite-(Ce) forms at a late stage in the crystallisation history of the carbonatites. Commonly, formation is considered to be related to a hydrothermal process, with temperature estimates from fluid-inclusion data between 385–315°C at Seligdar and 150–250°C at Mushgai Khudag (Prokopyev et al., Reference Prokopyev, Doroshkevich, Ponomarchuk and Sergeev2017; Nikolenko et al., Reference Nikolenko, Redina, Doroshkevich, Prokpyv, Rafozin and Vladykin2018).

There is little evidence for substantial hydrothermal alteration at Eureka. Local alteration includes small-scale quartz veins and carbonate recrystallisation. However, there is no evidence linking these events with the formation of the S-bearing monazite-(Ce). Instead, S-bearing monazite-(Ce) is considered here to be related to carbonatite weathering and duricrust formation. S-bearing monazite-(Ce) is found exclusively in silicified rocks (Fig. 3), which are restricted to ~5 m depth, and limited commonly to much shallower levels. Duricrusts are prevalent in the Namib Desert, with calcrete common in the area around Eureka (Viles and Goudie, Reference Viles and Goudie2013). Localised silcrete can form in calcrete owing to the pH-controlled inverse solubility relationship between silica and calcite, with precipitation of silica favoured when pH is <9, and vice-versa (Watts, Reference Watts1980). Mixing of Si-rich porefluids with alkaline water leads to Si precipitation, while Si-rich solutions in the presence of carbonate minerals become unstable in a saline or high CO2 environment (Watts, Reference Watts1980; Nash and Shaw, Reference Nash and Shaw1998). A change in salinity is possible in a continental environment by mixing meteoric water with saline lake fluids, while CO2 can be introduced into groundwater through biological activity or the decomposition of organic matter (Bustillo, Reference Bustillo2010). Alternatively, reduction of pH at oxic/anoxic boundaries, such as the through oxidation of H2S, can lead to silica precipitation (Bustillo, Reference Bustillo2010). Of these three possibilities, salinity change seems the most probable, owing to the close proximity of the Eureka carbonatites to ephemeral rivers and the salinity of local groundwater (encountered at ~60 m during drilling). However, when considering the S source, a contribution from H2S cannot be entirely excluded.

Formation mechanism for S-bearing monazite-(Ce) at Eureka

While the sole association of the S-bearing monazite-(Ce) with silicified rocks demonstrates that weathering and silicification play a key role in its formation, the exact growth mechanism of the S-bearing monazite-(Ce) is unclear. Solid-state diffusion is unlikely given the lack of a local source of S, and negligible diffusion rates in monazite below 800°C (Cherniak et al., Reference Cherniak, Watson, Grove and Harrison2004; Gardés et al., Reference Gardés, Jaoul, Montel, Seydoux-Guillaume and Wirth2006, Reference Gardés, Montel, Seydoux-Guillaume and Wirth2007). Overgrowths can also be excluded owing to the corroded and pitted texture of the core monazite-(Ce) grains (Fig. 3b,c). Instead, we propose that the S-bearing monazite-(Ce) probably formed via a dissolution–precipitation reaction (Putnis, Reference Putnis2002, Reference Putnis, Oelkers and Schott2009). Such a reaction involves fluid-mediated replacement of a phase to reduce the free energy of a system. The product phase(s) must be a lower molar volume or higher solubility in order for the reaction to propagate (Putnis and Ruiz-Agudo, Reference Putnis and Ruiz-Agudo2013). Although monazite is a relatively insoluble mineral (e.g. Poitrasson et al., Reference Poitrasson, Oelkers, Schott and Montel2004; Cetiner et al., Reference Cetiner, Wood and Gammons2005), dissolution–precipitation has been demonstrated experimentally across a wide range of pressures and temperatures (e.g. Hetherington et al., Reference Hetherington, Harlov and Budzyń2010; Harlov and Hetherington, Reference Harlov and Hetherington2010; Harlov et al., Reference Harlov, Wirth and Hetherington2011; Budzyń et al., Reference Budzyń, Harlov, Kozub-Budzyń and Majka2016). Texturally, the S-bearing monazite-(Ce) fits several characteristics of a dissolution–precipitation reaction (Putnis, Reference Putnis2002, Reference Putnis, Oelkers and Schott2009), such as: (1) the close relationship between the parent and product phases; (2) a sharp reaction front between these two phases; and (3) development of porosity and permeability in the product phase (Fig. 3d). Fractures also occur perpendicular to the reaction front (e.g. Fig. 3f), although these are localised and cross-cut the product phase, suggesting formation unrelated to the growth of S-bearing monazite-(Ce).

Although the textural evidence for a dissolution–precipitation formation mechanism is convincing, we note that no experimental work has yet demonstrated that such a process occurs under ambient conditions similar to those suggested to form the S-bearing monazite-(Ce) at Eureka. Indeed, the lowest-temperature dissolution–precipitation experiments on monazite-(Ce) demonstrate quite contrasting findings. Grand'Homme et al. (Reference Grand'Homme, Janots, Seydoux-Guillaume, Guillaume, Magnin, Hövelmann, Höschen and Boiron2018) showed that a relatively simple assemblage of monazite, quartz, H2O and NaOH dissolved at 300°C, but secondary monazite did not form. In contrast, Budzyń et al. (Reference Budzyń, Konečný and Kozub-Budzyń2015) found monazite does show dissolution and precipitation textures in an experiment at 250°C with a similar, but more complex, assemblage of monazite, K-feldspar, albite, muscovite, biotite, H2O and Na2Si2O5. At crystallisation temperatures below those of the above experiments, the more common LREE phosphate occurring in nature is the hydrated equivalent to monazite-(Ce), rhabdophane-(Ce) (Kolitsch and Holtstam, Reference Kolitsch and Holtstam2004). Nonetheless, rare occurrences of monazite-(Ce) in diagenetic, weathering and low-grade metamorphic environments have been described (e.g. Cooper et al., Reference Cooper, Basham and Smith1983; Oliveira and Imbernon, Reference Oliveira and Imbernon1998; Cabella et al., Reference Cabella, Lucchetti and Marescotti2001; Rasmussen and Muhling, Reference Rasmussen and Muhling2007; Čopjaková et al., Reference Čopjaková, Novák and Franců2011), indicating that monazite can form in such environments at low temperatures. It is not clear, however, why S-bearing monazite at Eureka is a more stable assemblage under such conditions, rather than a mixture of gypsum and monazite-(Ce).

Cross-cutting relationships indicate that chalcedony continued to form after the cessation of S-bearing monazite-(Ce) growth. Such growth would probably limit the permeability of the rock, reducing further ingress of S-bearing solutions. Complete silicification may, therefore, be a limiting factor in the alteration of the monazite-(Ce) to S-bearing monazite-(Ce).

Source of elements in the S-bearing monazite

Most of the elements in the S-bearing monazite are probably derived from the breakdown of local minerals. For instance, dissolution–precipitation textures of the core monazite-(Ce) grains indicate that these are the source for P and REE (Fig. 3be). Strontium may also be derived, in part, from the core monazite-(Ce) grains, but is probably derived predominantly from the release of Sr from the breakdown and silicification of dolomite. Dolomite breakdown is also the probable source of Ca. Owing to a lack of sulfide minerals in the carbonatite and surrounding Etusis schists and quartzites, breakdown of local sulfide minerals is a negligible source of sulfur for forming the S-bearing monazite-(Ce). An alternative source is the extensive gypcrete formations occurring along the coast of the Namib Desert, up to 70 km inland (Ekardt et al., Reference Ekardt, Drake, Goudie, White and Viles2001). The exact source of the S in the gypcrete is enigmatic, but possibilities include marine-derived, fog-borne H2S (Brüchert et al., Reference Brüchert, Currie and Peard2009) or dimethylsulfide (Ekardt and Spiro, Reference Ekardt and Spiro1999). While Eureka is located slightly further inland (~90 km from the coast) than the limit of gypcrete formation, wind-blown transport of gypsum has also been suggested as a source of sulfate for inland weathering (Ekardt et al., Reference Ekardt, Drake, Goudie, White and Viles2001). Sulfur isotope analyses would help further address this question, but are beyond the scope of the present study.

Conclusions and implications

Sulfur-bearing monazite-(Ce) occurs in silicified carbonatite at Eureka, Namibia, forming rims up to ~0.5 mm thick on earlier-formed monazite-(Ce) megacrysts. We present new XPS data which proves for the first time that the sulfur is predominantly accommodated as sulfate, via a clino-anhydrite-type coupled substitution mechanism. This largely confirms the assumption that S uptake into monazite-(Ce) is limited to oxidising conditions only, controlled by the oxidation state of the host rock (e.g. Laurent et al., Reference Laurent, Seydoux-Guilaume, Duchene, Bingen, Bosse and Datas2016). Nonetheless, minor sulfide and sulfite peaks in the X-ray photoelectron spectra suggest that more complex substitution mechanisms incorporating S2– and S4+ are possible and that a more cautious approach for interpreting crystallisation environment from the presence of S in monazite is warranted. Incorporation of S6+ through clino-anhydrite-type substitution invariably results in an excess of M 2+ cations, which previous workers have suggested is accommodated by auxiliary substitution of OH for O2–. However, our new Raman data show no indication of OH, and instead we suggest charge imbalance is accommodated through F substituting for O2–. The accommodation of S in the monazite-(Ce) results in considerable structural distortion which can account for relatively high contents of ions with radii beyond those normally found in monazite-(Ce), such as the HREE, Mo, Zr and V.

In contrast to S-bearing monazite-(Ce) in other carbonatites, S-bearing monazite-(Ce) at Eureka formed through a prolonged weathering process. This is indicated predominantly by the localisation of the S-bearing monazite-(Ce) to silicified rocks which are constrained to the upper-most 5 m of exposure. Ca, REE and P were derived from the local dissolution monazite-(Ce) and dolomite, and S-bearing monazite formed via a coupled dissolution–precipitation mechanism. We suggest that S is sourced via protracted aeolian input, as is the case for local gypcrete formations. While large S-bearing monazite-(Ce) grains are likely to be rare in the geological record, formation of secondary S-bearing monazite-(Ce) in these conditions may be a feasible mineral for dating palaeo-weathering, especially weathered carbonatites. Co-substitution of PbSO4 during clino-anhydrite exchange is a notable challenge for dating S-bearing monazite-(Ce) (Krenn et al., Reference Krenn, Putz, Finger and Paar2011). Nonetheless, at Eureka the 208Pb concentration in the S-bearing monazite-(Ce) is approximately half that of the original grain, while U contents are higher.

Acknowledgements

Pete Siegfried (GeoAfrica) and Tim Smalley (E-Tech Metals) are both thanked for suggesting the importance of weathering at Eureka. Jindrich Kynický is thanked for supplying an apatite sample from Slyudjanka. The original sampling and fieldwork was funded by a Society of Economic Geologists Hugh McKinstry Fund award to SBF, with additional funding from the NERC SoS RARE consortium (NE/M011429/1; SBF, MS, MBA, DAB, DA and FW) and a NERC Industrial Innovation Fellowship (NE/R013403/1) to SBF. DA and MBA acknowledge the Brazilian research foundation FAPESP (process 2014/50819-9) for financial support. The authors are grateful for the thoughtful reviews of Martin Ondrejka, Callum J Hetherington and Jean-Marc Montel, as well as the editorial handling of Éimear Deady, Stuart Mills and Helen Kerbey.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1180/mgm.2019.79

Footnotes

Guest Editor: Eimear Deady

This paper is part of a thematic set arising from the 3rd International Critical Metals Conference (Edinburgh, May 2019).

References

Abraham, K.M. and Chaudhri, S.M. (1986) The lithium surface film in the Li/SO2 cell. Journal of the Electrochemical Society, 133, 13071311.10.1149/1.2108858CrossRefGoogle Scholar
Andreoli, M.A.G., Smith, C.B., Watkeys, M., Moore, J.M., Ashwal, L.D. and Hart, R.J. (1994) The Geology of the Steenkampskraal monazite deposit, South Africa: Implications for REE–Th–Cu mineralisation in charnockite-granulite terranes. Economic Geology, 89, 9941016.10.2113/gsecongeo.89.5.994CrossRefGoogle Scholar
Ayers, J.C., Loflin, M., Miller, C.F., Barton, M.D. and Coath, C.D. (2006) In situ oxygen isotope analysis of monazite as a monitor of fluid infiltration during contact metamorphism: Birch Creek Pluton aureole, White Mountains, eastern California. Geology, 34, 653656.10.1130/G22185.1CrossRefGoogle Scholar
Beall, G.W., Boatner, L.A., Mullica, D.F. and Milligan, W.O. (1981) The structure of cerium orthophosphate, a synthetic analogue of monazite. Journal of Inorganic and Nuclear Chemistry, 43, 101105.10.1016/0022-1902(81)80443-5CrossRefGoogle Scholar
Begun, G.M., Beall, G. W., Boatner, L.A. and Gregor, W.J. (1981) Raman spectra of the rare earth orthophosphates. Journal of Raman Spectroscopy, 11, 273278.10.1002/jrs.1250110411CrossRefGoogle Scholar
Bermanec, V., Tibljaš, D., Gessner, M. and Kniewald, G. (1988) Monazite in hydrothermal veins from Alinci, Yugoslavia. Mineralogy and Petrology, 38, 139150.10.1007/BF01164318CrossRefGoogle Scholar
Borg, I.Y. and Smith, D.K. (1975) A high pressure polymorph of CaSO4. Contributions to Mineralogy and Petrology, 50, 127133.10.1007/BF00373332CrossRefGoogle Scholar
Bradbury, S.E. and Williams, Q. (2009) X-ray diffraction and infrared spectroscopy of monazite–structured CaSO4 at high pressures: Implications for shocked anhydrite. Journal of Physics and Chemistry of Solids, 70, 134141.CrossRefGoogle Scholar
Broom-Fendley, S., Wall, F., Spiro, B. and Ullmann, C.V. (2017) Deducing the source and composition of rare earth mineralising fluids in carbonatites: insights from isotopic (C, O, 87Sr/86Sr) data from Kangankunde, Malawi. Contributions to Mineralogy and Petrology, 172, 96.CrossRefGoogle ScholarPubMed
Brounce, M., Boyce, J., McCubbin, F.M., Humphreys, J., Reppart, J., Stolper, E. and Eiler, J. (2019) The oxidation state of sulfur in lunar apatite. American Mineralogist, 104, 307312.10.2138/am-2019-6804CrossRefGoogle Scholar
Brüchert, V., Currie, B. and Peard, K.R. (2009) Hydrogen sulphide and methane emissions on the central Namibian shelf. Progress in Oceanography, 83, 169179.10.1016/j.pocean.2009.07.017CrossRefGoogle Scholar
Budzyń, B., Konečný, P. and Kozub-Budzyń, G.A. (2015) Stability of monazite and disturbance of the Th–U–Pb system under experimental conditions of 250–350°C and 200–400 MPa. Annales Societatis Geologorum Poloniae, 85, 405424.CrossRefGoogle Scholar
Budzyń, B., Harlov, D.E., Kozub-Budzyń, G.A. and Majka, J. (2016) Experimental constraints on the relative stabilities of the two systems monazite-(Ce) – allanite-(Ce) – fluorapatite and xenotime-(Y) – (Y,HREE)-rich epidote – (Y,HREE)-rich fluorapatite, in high Ca and Na–Ca environments under P–T conditions of 200–1000 MPa and 450–750°C. Mineralogy and Petrology, 111, 183217.CrossRefGoogle Scholar
Bulakh, A.G., Nesterov, A.R., Zaitsev, A.N., Pilipiuk, A.N., Wall, F. and Kirillov, A.S. (2000) Sulfur–containing monazite-(Ce) from late-stage mineral assemblages at the Kandaguba and Vuoriyarvi carbonatite complexes, Kola peninsula, Russia. Neues Jahrbuch für Mineralogie – Monatshefte, 5, 217233.Google Scholar
Burger, A.J., von Knorring, O. and Clifford, T.N. (1965) Mineralogical and radiometric studies of monazite and sphene occurrences in the Namib Desert, South-West Africa. Mineralogical Magazine, 35, 519528.CrossRefGoogle Scholar
Bustillo, M.A. (2010) Silicification of Continental Carbonates. Developments in Sedimentology, 62, 153178.10.1016/S0070-4571(09)06203-7CrossRefGoogle Scholar
Cabella, R., Lucchetti, G. and Marescotti, P. (2001) Authigenic monazite and xenotime from peletic metacherts in pumpellyite–actinolite facies conditions, Sestri–Voltaggio Zone, Central Liguria, Italy. The Canadian Mineralogist, 39, 717727.10.2113/gscanmin.39.3.717CrossRefGoogle Scholar
Cetiner, Z.S., Wood, S.A., and Gammons, C.H. (2005) The aqueous geochemistry of the rare earth elements. Part XIV. The solubility of rare earth element phosphates from 23 to 150°C. Chemical Geology, 217, 147169.CrossRefGoogle Scholar
Chakhmouradian, A.R. and Mitchell, R.H. (1999) Niobian ilmenite, hydroxylapatite and sulfatian monazite; alternative hosts for incompatible elements in calcite kimberlite from Internatsional'naya, Yakutia. The Canadian Mineralogist, 37, 11771189.Google Scholar
Chakhmouradian, A.R. and Wall, F. (2012) Rare earth elements: minerals, mines, magnets (and more). Elements, 8, 333340.10.2113/gselements.8.5.333CrossRefGoogle Scholar
Chen, W., Honghui, H., Bai, T. and Jiang, S. (2017) Geochemistry of monazite within carbonatite related REE deposits. Resources, 6, doi:10.3390/resources6040051CrossRefGoogle Scholar
Cherniak, D.J., Watson, E.B., Grove, M., and Harrison, T.M. (2004) Pb diffusion in monazite: a combined RBS/SIMS study. Geochimica et Cosmochimica Acta, 68, 829840.10.1016/j.gca.2003.07.012CrossRefGoogle Scholar
Cooper, D.C., Basham, I.R. and Smith, T.K. (1983) On the occurrence of an unusual form of monazite in panned stream sediments in Wales. Geological Journal, 18, 121127.CrossRefGoogle Scholar
Čopjaková, R., Novák, M. and Franců, E. (2011) Formation of authigenic monazite-(Ce) to monazite-(Nd) from Upper Carboniferous graywackes of the Drahany Upland: Roles of the chemical composition of host rock and burial temperature. Lithos, 127, 373385.CrossRefGoogle Scholar
Cressey, G., Wall, F. and Cressey, B.A. (1999) Differential REE uptake by sector growth of monazite. Mineralogical Magazine, 63, 813828.CrossRefGoogle Scholar
Crichton, W.A., Parise, J.B., Antao, S.M. and Grzechnik, A. (2005) Evidence for monazite-, barite-, and AgMnO4 (distorted barite)-type structures of CaSO4 at high pressure and temperature. American Mineralogist, 90, 2227.10.2138/am.2005.1654CrossRefGoogle Scholar
Cuney, M. and Kyser, K. (2015) Thorium deposits. Pp. 319334 in: Geology and Geochemistry of Uranium and Thorium Deposits (Cuney, M. and Kyser, K., editors). Mineralogical Association of Canada Short Course Series, 46.Google Scholar
Doroshkevich, A.G., Ripp, G.S. and Moore, K.R. (2001) Genesis of the Khaluta alkaline-basic Ba–Sr carbonatite complex (West Transbaikala, Russia). Mineralogy and Petrology, 98, 245268.CrossRefGoogle Scholar
Dunai, T. (1989) Petrographische, Geochemische und Lagerstättenkundliche Untersuchungen an Karbonatitgängen auf der Farm Eureka Nr 99, Damaraland, Namibia. Diploma dissertation [in German], Heidelberg University, Germany.Google Scholar
Dunai, T., Stoessel, G.F.U. and Ziegler, U.R.F. (1989) A Sr isotope study of the Eureka Carbonatite, Damaraland, Namibia. Communications of the Geological Survey of Namibia, 5, 9192.Google Scholar
Ekardt, F.D. and Spiro, B. (1999) The origin of sulphur in gypsum and dissolved sulphate in the Central Namib Desert, Namibia. Sedimentary Geology, 123, 255273.CrossRefGoogle Scholar
Ekardt, F.D., Drake, N., Goudie, A.S., White, K. and Viles, H. (2001) The role of playas in pedogenic gypsum crust formation in the Central Namib Desert: a theoretical model. Earth Surface Processes and Landforms, 26, 11771193.CrossRefGoogle Scholar
Engi, M. (2017) Petrochronology based on REE-minerals: monazite, allanite, xenotime, apatite. Pp. 365418 in: Petrochronology: Methods and Applications (Kohn, M.J., Engi, M., and Lanari, P., editors). Reviews in Mineralogy and Geochemistry, 83. Mineralogical Society of America and the Geochemical Society, Chantilly, Virginia, USA.Google Scholar
Engi, M., Lanari, P. and Kohn, M.J. (2017) Significant ages – an introduction to petrochronology. Pp. 112 in: Petrochronology: Methods and Applications (Kohn, M.J., Engi, M., and Lanari, P., editors). Reviews in Mineralogy and Geochemistry, 83. Mineralogical Society of America and the Geochemical Society, Chantilly, Virginia, USA.Google Scholar
Enkhbayar, D., Sea, J., Choi, S-G., Lee, Y.J. and Batmunkh, E. (2016) Mineral Chemistry of REE-rich apatite and sulfur-rich monazite from the Mushgai Khudag, Alkaline Volcanic-Plutonic Complex, South Mongolia. International Journal of Geosciences, 7, 2031.CrossRefGoogle Scholar
Ewing, R.C. and Wang, L. (2002) Phosphates as Nuclear Waste Forms. Pp. 673699 in: Phosphates (Kohn, M.L., Rakovan, J. and Hughes, J.M., editors). Reviews in Mineralogy and Geochemistry, 48. Mineralogical Society of America and the Geochemical Society, Washington, DC.CrossRefGoogle Scholar
Förster, H.J. (1998) The chemical composition of REE–Y–Th–U–rich accessory minerals in peraluminous granites of the Erzgebirge–Fichtelgebirge region, Germany; Part I, The monazite-(Ce)–brabantite solid solution series. American Mineralogist, 83, 259272.CrossRefGoogle Scholar
Gardés, E., Jaoul, O., Montel, J-M., Seydoux-Guillaume, A-M. and Wirth, R. (2006) Pb diffusion in monazite: An experimental study of Pb2+ + Th4+ ⇔ 2Nd3+ interdiffusion. Geochimica et Cosmochimica Acta, 70, 23252336.CrossRefGoogle Scholar
Gardés, E., Montel, J-M., Seydoux-Guillaume, A-M. and Wirth, R. (2007) Pb diffusion in monazite: New constraints from the experimental study of Pb2+ <=> Ca2+ interdiffusion. Geochimica et Cosmochimica Acta, 71, 40364043.CrossRefGoogle Scholar
Grand'Homme, A., Janots, E., Seydoux-Guillaume, A.M., Guillaume, D., Magnin, V., Hövelmann, J., Höschen, C. and Boiron, M.C. (2018) Mass transport and fractionation during monazite alteration by anisotropic replacement. Chemical Geology, 484, 5168.CrossRefGoogle Scholar
Gonçalves, G.O., Lana, C., Scholz, R., Buick, I.S., Gerdes, A., Kamo, S.L., Corfu, F., Marinho, M.M., Chaves, A.O., Valeriano, C. and Nalini, H.A.N. Jr (2016) An assessment of monazite from the Itambé pegmatite district for use as U-Pb isotope reference material for microanalysis and implications for the origin of the “Moacyr” monazite. Chemical Geology, 424, 3050.CrossRefGoogle Scholar
Gonçalves, G.O., Lana, C., Scholz, R., Buick, I.S., Gerdes, A., Kamo, S.L., Corfu, F., Rubatto, D., Wiedenbeck, M., Nalini, H.A.M. Jr. and Oliveira, L.C.A. (2018) The Diamantina monazite: A new low-Th reference material for microanalysis. Geostandards and Geoanalytical Research, 42, 2547.CrossRefGoogle Scholar
Harlov, D.E., and Hetherington, C.J. (2010) Partial high-grade alteration of monazite using alkali-bearing fluids: experiment and nature. American Mineralogist, 95, 11051108.CrossRefGoogle Scholar
Harlov, D.E., Wirth, R., and Hetherington, C.J. (2011) Fluid-mediated partial alteration in monazite: the role of coupled dissolution–reprecipitation in element redistribution and mass transfer. Contributions to Mineralogy and Petrology, 162, 329348.CrossRefGoogle Scholar
Hetherington, C.J., Harlov, D.E. and Budzyń, B. (2010) Experimental metasomatism of monazite and xenotime: mineral stability, REE mobility and fluid composition. Mineralogy and Petrology, 99, 165184.CrossRefGoogle Scholar
Hetherington, C.J., Backus, E.L., McFarlane, C.R.M., Fisher, C.M. and Pearson, D.G. (2018) Origins of textural, compositional, and isotopic complexity in monazite and its petrochronological analysis. Pp. 6390 in: Microstructural Geochronology: Planetary Records Down to Atom Scale, Geophysical Monograph 232, First Edition (Moser, D.E., Corfu, F., Darling, J.R., Reddy, S.M., and Tait, K., editors). John Wiley & Sons.Google Scholar
Jercinovic, M.J., and Williams, M.L. (2005) Analytical perils (and progress) in electron microprobe trace element analysis applied to geochronology: Background acquisition, interferences, and beam irradiation effects. American Mineralogist, 90, 526546.CrossRefGoogle Scholar
Kahn, J.S. (1975) High-pressure phase transformation of CaSO4 (Anhydrite) during a nuclear explosion. Science, 189, 454455.CrossRefGoogle Scholar
Kim, Y.J., Konecke, B., Fiege, A., Simon, A. and Becker, U. (2017) An ab initio study of the energetics and geometry of sulfide, sulfite, and sulfate incorporation into apatite: The thermodynamic basis for using this system as an oxybarometer. American Mineralogist, 102, 16461656.CrossRefGoogle Scholar
Kolitsch, U., and Holtstam, D. (2004) Crystal chemistry of REEXO4 compounds (X = P, As,V). II. Review of REEXO4 compounds and their stability fields. European Journal of Mineralogy, 16, 117126.CrossRefGoogle Scholar
Konecke, B.A., Fiege, A., Simon, A.C., Parat, F. and Stechern, A. (2017 a) Co-variability of S6+, S4+, and S2– in apatite as a function of oxidation state: Implications for a new oxybarometer. American Mineralogist, 102, 548557.CrossRefGoogle Scholar
Konecke, B.A., Fiege, A., Simon, A.C. and Holtz, F. (2017 b) Cryptic metasomatism during late-stage lunar magmatism implicated by sulfur in apatite. Geology, 45, 739742.Google Scholar
Krenn, E., Putz, H., Finger, F. and Paar, W.H. (2011) Sulfur-rich monazite with high common Pb in ore-bearing schists from the Schellgaden mining district (Tauern Window, Eastern Alps), Mineralogy and Petrology, 102, 5162.CrossRefGoogle Scholar
Kukharenko, A.A., Bulakh, A.G. and Baklanova, (1961) Sulphate-monazite from the carbonatites of the Kola Peninsula. Zapiski Vsesoyuznogo Mineralogicheskogo Obshchestva, 90, 373381 [in Russian].Google Scholar
Laurent, A.T., Seydoux-Guilaume, A-M., Duchene, S., Bingen, B., Bosse, V. and Datas, L. (2016) Sulphate incorporation in monazite lattice and dating the cycle of sulphur in metamorphic belts. Contributions to Mineralogy and Petrology, 171, 94.CrossRefGoogle Scholar
Lazareva, E.V., Zhmodik, S.M., Dobretsov, N.L., Tolstov, A.V., Shcherbov, B.L., Karmanov, N.S., Gerasimov, E.Yu. and Bryanskaya, A.V. (2015) Main minerals of abnormally high-grade ores of the Tomtor deposit (Arctic Siberia). Russian Geology and Geophysics, 56, 844873.CrossRefGoogle Scholar
Lehmann, B., Nakai, S, Höhndorf, A., Brinckmann, J., Dulski, P., Hein, U.F. and Masuda, A. (1994) REE mineralization at Gakara, Burundi: Evidence for anomalous upper mantle in the western Rift Valley. Geochimica et Cosmochimica Acta, 58, 985992.CrossRefGoogle Scholar
Lenz, C., Nasdala, L., Talla, D., Hauzenberger, C., Seitz, R. and Kolitsch, U. (2015) Laser-induced REE 3+ photoluminescence of selected accessory minerals – An “advantageous artefact” in Raman spectroscopy. Chemical Geology, 415, 116.CrossRefGoogle Scholar
Lottermoser, B.G. (1990) Rare-earth element mineralisation within the Mt. Weld carbonatite laterite, Western Australia. Lithos, 24, 151167.CrossRefGoogle Scholar
McDonough, W.F. and Sun, S. (1995) The composition of the earth. Chemical Geology, 120, 223253.CrossRefGoogle Scholar
McFarlane, C.R.M. and McCulloch, M.T. (2007) Coupling of in situ Sm–Nd systematics and U–Pb dating of monazite and allanite with applications to crustal evolution studies. Chemical Geology, 245, 4560.CrossRefGoogle Scholar
Mesarwi, A. and Ignatiev, A. (1993) Interaction of Y overlayers with the GaAs(100) surface and oxidation of the Y/GaAs interface. Surface Science, 282, 371379.CrossRefGoogle Scholar
Miller, R. McG. (1983) The Pan-African Damara Orogen of South West Africa/Namibia. Pp. 431515 in: Evolution of the Damara Orogen of South West Africa/Namibia (Miller, R. McG, editor). Special Publications of the Geological Society of South Africa, 11.Google Scholar
Miller, R. McG (2008) The Geology of Namibia. Geological Survey of Namibia, Windhoek.Google Scholar
Montel, J-M., Foret, S., Veschambre, M., Nicollet, C. and Provost, A. (1996) Electron microprobe dating of monazite. Chemical Geology, 131, 3753.CrossRefGoogle Scholar
Montel, J-M., Razafimahatratra, D., de Parseval, P., Poitrasson, F., Moine, B., Seydoux-Geuillaume, A-M., Pik, R., Arnaud, N. and Gibert, F. (2018) The giant monazite crystals from Manangotry (Madagascar). Chemical Geology, 484, 3650.CrossRefGoogle Scholar
Nash, D.J. and Shaw, P.A. (1998) Silica and carbonate relationships in silcrete-calcrete intergrade duricrusts from the Kalahari of Botswana and Namibia. Journal of African Earth Sciences, 27, 1125.CrossRefGoogle Scholar
Ni, Y., Hughes, J.M. and Mariano, A.N. (1995) Crystal chemistry of the monazite and xenotime structures. American Mineralogist, 80, 2126.CrossRefGoogle Scholar
Nikolenko, A.M., Redina, A.A., Doroshkevich, A.G., Prokpyv, I.R., Rafozin, A.L. and Vladykin, N.V. (2018) The origin of magnetite–apatite rocks of Mushgai–Khudag Complex, South Mongolia: mineral chemistry and studies of melt and fluid inclusions. Lithos, 320–321, 567582.CrossRefGoogle Scholar
Ntiharirizwa, S., Boulvais, P., Poujol, M., Branquet, Y., Morelli, C., Ntungwanayo, J. and Midende, G. (2018) Geology and U–Th–Pb Dating of the Gakara REE Deposit, Burundi. Minerals, 8, doi:10.3390/min8090394CrossRefGoogle Scholar
Oelkers, E.H. and Montel, J-M. (2008) Phosphates and nuclear waste storage. Elements, 4, 113116.CrossRefGoogle Scholar
Oliveira, S.M.B. and Imbernon, R.A.L. (1998) Weathering alteration and related REE concentration in the Catalão I carbonatite complex, central Brazil. Journal of South American Earth Sciences, 11, 379388.CrossRefGoogle Scholar
Ondrejka, M., Uher, P., Pršek, J. and Ozdín, D. (2007) Arsenian monazite-(Ce) and xenotime–(Y), REE arsenates and carbonates from the Tisovec–Rejkovo rhyolite, Western Carpathians, Slovakia: Composition and substitutions in the (REE,Y)XO4 system (X = P, As, Si, Nb, S). Lithos, 95, 116129.CrossRefGoogle Scholar
Ondrejka, M., Putiš, M, Uher, P., Schmiedt, I., Pukančík, L. and Konečný, P. (2016) Fluid-driven destabilization of REE-bearing accessory minerals in the granitic orthogneisses of North Veporic basement (Western Carpathians, Slovakia). Mineralogy and Petrology, 110, 561580.CrossRefGoogle Scholar
Pan, Y.M. and Fleet, M.E. (2002) Compositions of the apatite-group minerals: substitution mechanisms and controlling factors. Pp. 1349 in: Phosphates (Kohn, M.L., Rakovan, J. and Hughes, J.M., editors). Reviews in Mineralogy and Geochemistry, 48. Mineralogical Society of America and the Geochemical Society, Washington, DC.CrossRefGoogle Scholar
Parrish, R.R. (1990) U–Pb dating of monazite and its application to geological problems. Canadian Journal of Earth Sciences, 27, 14311450.CrossRefGoogle Scholar
Paton, C., Hellstrom, J., Paul, B., Woodhead, J. and Hergt, J. (2011) Iolite: Freeware for the visualisation and processing of mass spectrometric data. Journal of Analytical Atomic Spectrometry, 26, 25082518.CrossRefGoogle Scholar
Poitrasson, F., Oelkers, E., Schott, J., and Montel, J-M (2004) Experimental determination of synthetic NdPO4 monazite end-member solubility in water from 21°C to 300°C: implications for rare earth element mobility in crustal fluids. Geochimica et Cosmochimica Acta, 68, 22072221.CrossRefGoogle Scholar
Prokopyev, I.R., Doroshkevich, A., Ponomarchuk, A.V. and Sergeev, S.A. (2017) Mineralogy, age and genesis of apatite–dolomite ores at the Seligdar apatite deposit (Central Aldan, Russia). Ore Geology Reviews, 81, 296308.CrossRefGoogle Scholar
Pršek, J., Ondrejka, M., Bačík, P., Budzyń, B. and Uher, P. (2010) Metamorphic-hydrothermal REE minerals in the Bacúch magnetite deposit, Western Carpathians, Slovakia: (Sr,S)-rich monazite-(Ce) and Nd-dominant hingganite. The Canadian Mineralogist, 48, 8194.CrossRefGoogle Scholar
Putnis, A. (2002) Mineral replacement reactions: from macroscopic observations to microscopic mechanisms. Mineralogical Magazine, 66, 689708.CrossRefGoogle Scholar
Putnis, A. (2009) Mineral replacement reactions. Pp. 87124 in: Thermodynamics and Kinetics of Water-Rock Interaction (Oelkers, E.H. and Schott, J., editors). Reviews in Mineralogy and Geochemistry, 70. Mineralogical Society of America and the Geochemical Society, Chantilly, Virginia, USA.CrossRefGoogle Scholar
Putnis, C.V. and Ruiz-Agudo, E. (2013) The mineral-water interface: where minerals react with the environment. Elements, 9, 177182.CrossRefGoogle Scholar
Ragettli, R.A., Hebeda, E.H., Signer, P. and Wieler, R. (1994) Uranium-xenon chronology: precise determination of λsf*136Ysf for spontaneous fission of 238U. Earth and Planetary Science Letters, 128, 653670.CrossRefGoogle Scholar
Rasmussen, B. and Muhling, J.R. (2007) Monazite begets monazite: evidence for dissolution of detrital monazite and reprecipitation of syntectonic monazite during low-grade regional metamorphism. Contributions to Mineralogy and Petrology, 154, 675689.CrossRefGoogle Scholar
Sadove, G., Konecke, B.A., Fiege, A. and Simon, A.C. (2019) Structurally bound S2−, S1−, S4+, S6+ in terrestrial apatite: The redox evolution of hydrothermal fluids at the Phillips mine, New York, USA. Ore Geology Reviews, 107, 10841096.CrossRefGoogle Scholar
Seydoux-Guillaume, A-M., Deschanels, X., Baumier, C., Neumeier, S., Weber, W.J. and Peuget, S. (2018) Why natural monazite never becomes amorphous: Experimental evidence for alpha self-healing. American Mineralogist, 103, 824827.CrossRefGoogle 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
Silva, E.N., Ayala, A.P., Guedes, I., Paschoal, C.W.A., Moreira, R.L., Loong, C.K. and Boatner, L.A. (2006) Vibrational spectra of monazite-type rare-earth orthophosphates. Optical Materials, 29(2–3), 224230.CrossRefGoogle Scholar
Slezak, P. and Spandler, C. (2019) Carbonatites as recorders of mantle-derived magmatism and subsequent tectonic events: An example of the Gifford Creek Carbonatite Complex, Western Australia. Lithos, 328–329, 212227.CrossRefGoogle Scholar
Spear, F.S. and Pyle, J.M. (2002) Apatite, monazite, and xenotime in metamorphic rocks. Pp. 293335 in: Phosphates (Kohn, M.L., Rakovan, J. and Hughes, J.M., editors). Reviews in Mineralogy and Geochemistry, 48. Mineralogical Society of America and the Geochemical Society, Washington, DC.CrossRefGoogle Scholar
Suzuki, K. and Kato, T. (2008) CHIME dating of monazite, xenotime, zircon and polycrase: Protocol, pitfalls and chemical criterion of possibly discordant age data. Gondwana Research, 14, 569586.CrossRefGoogle Scholar
Vasquez, R.P. (1991) X-ray photoelectron spectroscopy study of Sr and Ba compounds. Journal of Electron Spectroscopy and Related Phenomenon, 56, 217240.CrossRefGoogle Scholar
Viles, H.A. and Goudie, A.S. (2013) Weathering in the central Namib Desert, Namibia: Controls, processes and implications. Journal of Arid Environments, 93, 2029.CrossRefGoogle Scholar
von Knorring, O. and Clifford, T.N. (1960) On a skarn occurrence from the Namib desert near Usakos, South-West Africa. Mineralogical Magazine, 32, 650653.CrossRefGoogle Scholar
Voskoboinikova, N. (1938) The mineralogy of the Slyudyanka deposits of lazurite. Zapiski Vserossiyskogo Mineralogicheskogo Obshchestva, 67, 601622 [in Russian].Google Scholar
Wall, F. (2004) An illustration of the evolution and alteration of carbonatites using REE, Sr-rich carbonatites at Nkombwa, Zambia. Pp. 4968 in: Deep-Seated Magmatism, its Sources and Their Relation to Plume Processes (Vladykin, N.V., editor). Russian Academy of Sciences, Irkutsk.Google Scholar
Wall, F. (2014) Rare earth elements. Pp. 312339 in: Critical Metals Handbook (Gunn, A., editor). Wiley, New York.Google Scholar
Wall, F. and Mariano, A. (1996) Rare earth minerals in carbonatites: a discussion centred on the Kangankunde Carbonatite, Malawi. Pp. 193226 in: Rare Earth Minerals: Chemistry Origin and Ore Deposits (Jones, A.P., Wall, F. and Williams, C.T., editors). Chapman and Hall, London.Google Scholar
Watts, N.L. (1980) Quaternary pedogenic calcretes from the Kalahari (southern Africa): mineralogy, genesis and diagenesis. Sedimentology, 27, 661686.CrossRefGoogle Scholar
Williams, C.T. (1996) Analysis of rare earth minerals. Pp. 193226 in: Rare Earth Minerals: Chemistry Origin and Ore Deposits (Jones, A.P., Wall, F. and Williams, C.T., editors). Chapman and Hall, London.Google Scholar
Williams, M., Jercinovic, M.J. and Hetherington, C.J. (2007) Microprobe monazite geochronology: understanding geologic processes by integrating composition and chronology. Annual Review of Earth and Planetary Sciences, 35, 137175.CrossRefGoogle Scholar
Yu, X.R., Liu, F., Wang, Z.Y. and Chen, Y. (1990) Auger parameters for sulfur-containing compounds using a mixed aluminum–silver excitation source. Journal of Electron Spectroscopy and Related Phenomena, 50, 159166.CrossRefGoogle Scholar
Figure 0

Fig. 1. Location map (a) and geological sketch map (b) of the Eureka carbonatite dykes, with pit and sample locations. Host rocks are quartzite and schists of the Etusis formation. Geological map grid is UTM 33S, WGS 1984 datum, redrawn and georeferenced from Dunai (1989).

Figure 1

Fig. 2. Example of monazite-(Ce)-bearing (Mnz) silicified carbonatite from Eureka. Note the ~500 μm S-bearing monazite-(Ce) rims (S-Mnz) around the monazite-(Ce) grains.

Figure 2

Fig. 3. BSE images (ad, f) and EDS maps (e, gj) of silicified carbonatite and monazite-(Ce) from Eureka. (a) Planes of hematite (Hem), intersecting at ~120°, cemented by chalcedony (Cdy) and local, anhdedral, celestite (Cls) grains. (b) Monazite-(Ce) (Mnz) with a large S-bearing monazite-(Ce) rim (S-Mnz), hosted in chalcedony. (c) Close-up of monazite-(Ce)/S-bearing monazite-(Ce) boundary, showing symplectic texture between calcite (Cal) and S-bearing monazite-(Ce). (d) Close-up of monazite and S-bearing monazite rods, showing pore formation at the monazite-(Ce)/S-bearing monazite-(Ce) boundary, with (e) showing changes in the S content of the S-bearing monazite over a small area. (f) Example of cross-cutting chalcedony and hematite veins, through monazite-(Ce) and S-bearing monazite-(Ce). (gj) EDS maps demonstrating heterogeneous distribution of Ca and S in the S-bearing monazite (gh), the presence of Ca (g) and Fe (i) inclusions in the S-bearing monazite assemblage, and the presence of chalcedony and hematite veins. Larger circles correspond to sites of XPS analysis, while smaller circles are areas of LA-ICPMS analysis. Numbering corresponds to analytical locations and data in Tables 2, S2, and Fig. 5.

Figure 3

Fig. 4. Chondrite-normalised REE distribution of monazite-(Ce) and S-bearing monazite-(Ce) from Eureka. Chondrite values after McDonough and Sun (1995).

Figure 4

Table 1. Representative compositions of monazite-(Ce) and S-bearing monazite-(Ce) from Eureka (Sample SoS_63c).*

Figure 5

Table 2. Representative trace-element data (in μg/g) of monazite-(Ce) and S-bearing monazite-(Ce) from Eureka (Sample SoS_63c).*

Figure 6

Fig. 5. (a) X-ray photoelectron binding energy spectra of monazite-(Ce) from sites shown in Fig. 3. Note the HREE, S and Ca peaks present in the S-bearing monazite-(Ce). (b) Narrow scan of the Dy, Si, Y and S peaks. (c) Narrow scan of the S 2p binding energy peak indicating sulfur present as sulfite, sulfate and sulfide structurally bound in monazite-(Ce). Binding-energy peak positions from ThermoScientific, xpssimplified.com (accessed 2017), NIST XPS database (accessed 2017), Yu et al., (1990), Vasquez (1991) and Abraham and Chaudhri (1986).

Figure 7

Fig. 6. (a) Raman spectra of monazite-(Ce) and S-bearing monazite-(Ce) from sample SoS_63c. (b) Enlarged version of the above, demonstrating peak broadening and shifting.

Figure 8

Fig. 7. Composition of monazite-(Ce) from Eureka compared with data compiled from other carbonatite complexes (circles; Kukharenko et al., 1961; Cressey et al., 1999; Bulakh et al., 2000; Doroshkevich et al., 2001; Wall, 2004; Lazareva et al., 2015; Enkhbayar et al., 2016; Prokopyev et al., 2017; Nikolenko et al., 2018), the Internatsional'naya kimberlite (Chakhmouradian and Mitchell, 1999), and other published occurrences (squares) with SO3 >1% (Ondrejka et al., 2007; Pršek et al., 2010; Krenn et al., 2011; Ondrejka et al., 2016).

Supplementary material: File

Broom-Fendley et al. supplementary material

Broom-Fendley et al. supplementary material 1

Download Broom-Fendley et al. supplementary material(File)
File 1.6 MB
Supplementary material: File

Broom-Fendley et al. supplementary material

Broom-Fendley et al. supplementary material 2

Download Broom-Fendley et al. supplementary material(File)
File 48.4 KB
Supplementary material: File

Broom-Fendley et al. supplementary material

Supplementary table

Download Broom-Fendley et al. supplementary material(File)
File 22.2 KB