Hostname: page-component-6bf8c574d5-b4m5d Total loading time: 0 Render date: 2025-02-20T09:38:37.294Z Has data issue: false hasContentIssue false

Comparative thermal and compressional behaviour of natural xenotime-(Y), chernovite-(Y) and monazite-(Ce)

Mineralogy, petrology and geochemistry of pegmatites: Alessandro Guastoni memorial issue

Published online by Cambridge University Press:  12 November 2024

Francesco Pagliaro
Affiliation:
Dipartimento di Scienze della Terra, Università degli Studi di Milano, Via Botticelli 23, 20133 Milano, Italy
Davide Comboni
Affiliation:
Dipartimento di Scienze della Terra, Università degli Studi di Milano, Via Botticelli 23, 20133 Milano, Italy European Synchrotron Radiation Facility, 71 Avenue des Martyrs, CS 40220, 38043 Grenoble Cedex 9, France
Tommaso Battiston
Affiliation:
Dipartimento di Scienze della Terra, Università degli Studi di Milano, Via Botticelli 23, 20133 Milano, Italy
Hannes Krüger
Affiliation:
Institut für Mineralogie und Petrographie, Universität Innsbruck, Innrain 52, 6020 Innsbruck, Austria
Clivia Hejny
Affiliation:
Institut für Mineralogie und Petrographie, Universität Innsbruck, Innrain 52, 6020 Innsbruck, Austria
Volker Kahlenberg
Affiliation:
Institut für Mineralogie und Petrographie, Universität Innsbruck, Innrain 52, 6020 Innsbruck, Austria
Lara Gigli
Affiliation:
Elettra Sincrotrone Trieste S.c.P.A., Strada Statale 14 – km 163.5, 34149 Basovizza, Trieste, Italy
Konstantin Glazyrin
Affiliation:
Deutsches Elektronen-Synchrotron DESY, Notkestrasse 85, 22607 Hamburg, Germany
Hanns-Peter Liermann
Affiliation:
Deutsches Elektronen-Synchrotron DESY, Notkestrasse 85, 22607 Hamburg, Germany
Gaston Garbarino
Affiliation:
European Synchrotron Radiation Facility, 71 Avenue des Martyrs, CS 40220, 38043 Grenoble Cedex 9, France
G. Diego Gatta
Affiliation:
Dipartimento di Scienze della Terra, Università degli Studi di Milano, Via Botticelli 23, 20133 Milano, Italy
Paolo Lotti*
Affiliation:
Dipartimento di Scienze della Terra, Università degli Studi di Milano, Via Botticelli 23, 20133 Milano, Italy
*
Corresponding author: Paolo Lotti; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

ATO4 compounds are a class of oxides which includes the rare earth element (REE) bearing phosphates and arsenates, REEPO4 and REEAsO4. In this study, we have investigated the isothermal high-pressure and the isobaric high-temperature behaviour of natural samples of xenotime-(Y) (ideally YPO4), chernovite-(Y) (YAsO4) and monazite-(Ce) (CePO4) from the hydrothermal veins cropping out at Mt. Cervandone in the Western Italian Alps. Experimental data based on in situ X-ray diffraction (both single-crystal and powder techniques with conventional or synchrotron radiation) have allowed us to fit the unit-cell volumes and axial thermal and compressional evolution and provide a suite of refined thermo-elastic parameters. A comprehensive analysis of the role played by the crystal chemistry on the thermo-elastic response of these minerals is discussed, along with the description of the main crystal-structural deformation mechanisms for both the zircon (xenotime and chernovite) and monazite (monazite) structural types. Pressure-induced phase transitions of xenotime-(Y) and chernovite-(Y) are discussed and compared with previous literature data, whereas a change in the compressional behaviour of monazite-(Ce) at ∼18 GPa, involving an increase in the coordination number of the REE-hosting A site, is presented and discussed.

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
© The Author(s), 2024. Published by Cambridge University Press on behalf of The Mineralogical Society of the United Kingdom and Ireland.

Introduction

The general formula ATO4 is commonly used in the literature to define ternary inorganic oxides (Vorres, Reference Vorres1962), where A and T represent two cations that can be combined with oxygen (and occasionally with other anions) into several structural types, including, but not limited to, scheelite, zircon, monazite, fergusonite, baryte, quartz, cristobalite, wolframite and rutile (Fukunaga and Yamaoka, Reference Fukunaga and Yamaoka1979). In the context of the present study, the A site is occupied by a rare earth element (REE: lanthanides and Y), Ca, U and Th, whereas T stands for tetrahedrally-coordinated cations (As and P). In regard to the structural types, this manuscript focuses almost exclusively on the zircon and monazite structures, as demonstrated by the four minerals which are subject of this and two previous works (Pagliaro et al., Reference Pagliaro, Lotti, Comboni, Battiston, Guastoni, Fumagalli, Rotiroti and Gatta2022a, Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b): chernovite-(Y) [nominally YAsO4], xenotime-(Y) [nominally YPO4], gasparite-(Ce) [nominally CeAsO4] and monazite-(Ce) [nominally CePO4]. The crystal structure of these minerals has been the subject of a large number of studies and reviews (e.g. Mooney, Reference Mooney1948; Ni et al., Reference Ni, Hughes and Mariano1995; Boatner, Reference Boatner, Mottana, Paolo Sassi, Thompson and Guggenheim2002; Finch and Hanchar, Reference Finch, Hanchar, Hanchar and Hoskin2003 Kolitsch and Holtstam, Reference Kolitsch and Holtstam2004; Clavier et al., Reference Clavier, Podor and Dacheux2011) and an overview of the monoclinic monazite-type structure and of the tetragonal zircon-type (also known, but occasionally reported as ‘xenotime-type’) is discussed in the next section.

As reported by several authors (Fukunaga and Yamaoka, Reference Fukunaga and Yamaoka1979; Ushakov et al., Reference Ushakov, Helean, Navrotsky and Boatner2001; Boatner, Reference Boatner, Mottana, Paolo Sassi, Thompson and Guggenheim2002; Kolitsch and Holtsam, Reference Kolitsch and Holtstam2004), whether the monazite or the zircon-type structure is stable (within ATO4 phosphates and arsenates) depends on different factors. Among others, the atomic radii of either the A or T sites play a dominant role. In general, a large sized A cation promotes the crystallisation of the monazite structural type over the zircon one; on the other hand, the larger the T-site cation, the more stable is the zircon structure across the REE series. Within the REE-bearing phosphates, light REE ranging from La to Eu, with larger ionic radii than heavy REE, are hosted preferentially by the monazite-type structure, whereas heavy-REE, from Tb to Lu, including Y and Sc, fit best into the zircon-type structure (Mooney, Reference Mooney1948; Ni et al., Reference Ni, Hughes and Mariano1995; Boatner, Reference Boatner, Mottana, Paolo Sassi, Thompson and Guggenheim2002; Kolitsch and Holtstam, Reference Kolitsch and Holtstam2004; Clavier et al., Reference Clavier, Podor and Dacheux2011). A similar behaviour has been reported for the REEAsO4 series, although the threshold among the two structures is shifted to smaller Z numbers in the lanthanoid series: the monazite-type structure preferentially hosts REE from La to Nd, whereas the REE from Sm to Lu (as well as Y and Sc) are hosted by the tetragonal zircon-type crystal structure (e.g. Ushakov et al., Reference Ushakov, Helean, Navrotsky and Boatner2001; Boatner, Reference Boatner, Mottana, Paolo Sassi, Thompson and Guggenheim2002).

The REE-bearing phosphates are common accessory minerals in hydrothermal alteration of granitoid rocks that can control the partitioning of REE, as well as uranium and thorium, as they tend to incorporate these elements into their crystal structures (Rapp and Watson, Reference Rapp and Watson1986). In addition, due to their much lower tendency to incorporate Pb, they have found a significant use in geochronological applications (Harrison et al., Reference Harrison, Catlos, Montel, Mottana, Paolo Sassi, Thompson and Guggenheim2002). Due to their peculiar physical, chemical and optical properties (such as low solubility in water fluids), REE phosphates are used, or have been proposed, in several technological applications, e.g. to produce phosphors (de Sousa Filho and Serra, Reference Sousa Filho and Serra2009), ceramic coatings (Morgan et al., Reference Morgan, Marshall and Housley1995; Davis et al., Reference Davis, Marshall, Housley and Morgan1998), and materials for the safe storage of actinides originating from radioactive waste (Oelkers and Montel, Reference Oelkers and Montel2008; Orlova and Ojovan, Reference Orlova and Ojovan2019). Consequently, there has been significant interest in studying the behaviour of REE phosphate minerals and their synthetic counterparts under varying pressure and temperature conditions.

Recent reviews have been published by Errandonea (Reference Errandonea2017) and Strzelecki et al. (Reference Strzelecki, Zhao, Estevenon, Xu, Dacheux, Ewing and Guo2024) for monazite- and zircon-type structures, respectively. Tables S1 and S2 (supplementary material, see below) provide a comprehensive list of the thermo-elastic parameters published in the literature for several REETO4 compounds, along with the related bibliographic references. Concerning the response of REETO4 compounds to high pressure (hereafter HP), as a general rule it has been postulated that, for a given structural type, the bulk modulus shifts towards lower values as the atomic radius of the A and T sites increases, which has been highlighted by several authors and corroborated by both theoretical (Zhang et al., Reference Zhang, Zhou, Li and Li2008; Li et al., Reference Li, Zhang, Zhou and Cao2009) and experimental studies (Zhang et al., Reference Zhang, Zhou, Li and Li2008; Lacomba-Perales et al., Reference Lacomba-Perales, Errandonea, Meng and Bettinelli2010; Errandonea et al., Reference Errandonea, Kumar, López-Solano, Rodríguez-Hernández, Muñoz, Rabie and Sáez Puche2011). Regarding the thermal behaviour, there is general agreement on the thermal expansion coefficient of REETO4, which shows a clear compositional trend: the thermal expansivity increases along with the ionic radii of the A cation, while it reduces if the radius of the T site increases (Subbarao et al., Reference Subbarao, Agrawal, McKinstry, Sallese and Roy1990; Perrière et al., Reference Perrière, Bregiroux, Naitali, Audubert, Champion, Smith and Bernache-Assollant2007; Zhang et al., Reference Zhang, Zhou, Li and Li2008; Li et al., Reference Li, Zhang, Zhou and Cao2009). Unfortunately, the published values of both the compressibility and thermal expansivity of the REETO4 studied are not always internally consistent (see Tables S1 and S2), showing a certain degree of scattering, even for the same compound. In the case of thermal studies, the use of different thermal equations of state further complicates a comparative analysis. Therefore, in this study, we used the linear thermal expansion coefficient, which is commonly used in the literature, although it is not the most accurate model of the thermal elastic response (along with thermal equations of state commonly used in Earth Sciences).

The relative ratio of the A and T ionic radii not only affects the structure types adopted by a given compound at ambient conditions, but also its pressure stability field and the structural type of the high-pressure polymorphs. Such a relationship is well described by the so-called ‘Bastide diagram’ (Bastide, Reference Bastide1987), for which one of the most recent graphical representations is reported in Lopez-Solano et al. (Reference López-Solano, Rodríguez-Hernández, Muñoz, Gomis, Santamaría-Perez, Errandonea, Manjón, Kumar, Stavrou and Raptis2010) (Fig. 1, modified). The zircon-type compounds, xenotime-(Y) and chernovite-(Y) in this study, may transform at high-pressure into monazite-type or scheelite-type polymorphs. Whether a zircon → scheelite or a zircon → monazite → scheelite transformation occurs depends on the reciprocal relations among the ionic radii of the three atoms involved. A large T cation promotes a zircon-to-scheelite phase transition, whereas a small T cation favours an intermediate monazite-type polymorph. For the A-site cation, the larger it is, the more likely it is that a monazite polymorph will form over the scheelite one, ‘shifting’ the stability field to higher pressures. All the studied zircon-type ATO4 silicates show a zircon-to-scheelite phase transition, with high-pressure Raman and Density Functional Theory calculations suggesting the occurrence of a high-pressure lower symmetry polymorph preserving the zircon structural configuration before the reconstructive phase transition to reidite (Stangarone et al., Reference Stangarone, Angel, Prencipe, Mihailova and Alvaro2019). The zircon-to-scheelite phase transition has been described for YAsO4 (∼8 GPa) and YCrO4 (∼4.2 GPa) (Errandonea et al., Reference Errandonea, Kumar, López-Solano, Rodríguez-Hernández, Muñoz, Rabie and Sáez Puche2011), as well as YVO4 (above ∼7.5 GPa) (Jayaraman et al., Reference Jayaraman, Kourouklis, Espinosa, Cooper and van Uitert1987; Wang et al., Reference Wang, Loa, Syassen, Hanfland and Ferrand2004; Manjón et al., Reference Manjón, Rodríguez-Hernández, Muñoz, Romero, Errandonea and Syassen2010). A comprehensive description of the zircon-to-monazite phase transition in phosphates, including YPO4 (xenotime) is reported in Hay et al. (Reference Hay, Mogilevsky and Boakye2013). The relations between monazite and its HP-polymorphs again depend upon the reciprocal relations among the A, T and oxygen ionic radii. The monazite-to-post-baryte phase transition (space group P212121) has been described for REEPO4 and REEVO4 at increasing pressures with decreasing REE atomic radius (Lacomba-Perales et al., Reference Lacomba-Perales, Errandonea, Meng and Bettinelli2010; Ruiz-Fuertes et al., Reference Ruiz-Fuertes, Hirsch, Friedrich, Winkler, Bayarjargal, Morgenroth, Peters, Roth and Milman2016; Errandonea, Reference Errandonea2017). On the other hand, monazite-type LaVO4, PrVO4 and NdVO4, under compression, undergo a phase transition to a monoclinic BaWO4-II-type structure (Errandonea et al., Reference Errandonea, Pellicer-Porres, Martínez-García, Ruiz-Fuertes, Friedrich, Morgenroth, Popescu, Rodríguez-Hernández, Muñoz and Bettinelli2016; Panchal et al., Reference Panchal, Errandonea, Manjón, Muñoz, Rodríguez-Hernández, Achary and Tyagi2017; Marqueño et al., Reference Marqueño, Errandonea, Pellicer-Porres, Santamaria-Perez, Martinez-Garcia, Bandiello, Rodriguez-Hernandez, Muñoz, Achary and Popescu2021). Eventually, the monazite-to-scheelite transition has been described for the high-pressure polymorphs of YPO4 and several other REE-free compounds, as SrCrO4 and CaSeO4 (Crichton et al., Reference Crichton, Merlini, Müller, Chantel and Hanfland2012; Gleissner et al., Reference Gleissner, Errandonea, Segura, Pellicer-Porres, Hakeem, Proctor, Raju, Kumar, Rodríguez-Hernández, Muñoz, Lopez-Moreno and Bettinelli2016).

Figure 1. The so-called ‘Bastide diagram’ showing the relationships among structural types as a function of the atomic radii of cations at the A site (r A), T site (r T) and oxygen (r O), within the ATO4 family. The fields corresponding to the SrUO4 and BaWO4-II structures are labelled as orthorhombic (Cmca, Pbcm, Pnma) and monoclinic 14, respectively (2, 10, 12, 14 refer to the space group numbers). The post-baryte field is not reported (modified after López-Solano et al., Reference López-Solano, Rodríguez-Hernández, Muñoz, Gomis, Santamaría-Perez, Errandonea, Manjón, Kumar, Stavrou and Raptis2010).

This study, following the research conducted by Pagliaro et al. (Reference Pagliaro, Lotti, Comboni, Battiston, Guastoni, Fumagalli, Rotiroti and Gatta2022a, Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b), which includes a detailed crystal chemistry description, focuses on the high-pressure and temperature behaviour of four REETO4 mineralogical species from the same locality (Mt. Cervandone, Piedmont, Italy).

Based on experimental single-crystal or powder X-ray diffraction data collected in situ (high-P or high-T) at synchrotron beamlines or in conventional diffraction laboratories, a comparative analysis of the elastic behaviour and structural deformation mechanisms as a function of the crystal chemistry and structure type has been performed. The adoption of up-to-date experimental techniques and crystallographic methods has allowed us to describe a P-induced structural re-arrangement peculiar of the monazite structure type, previously reported by Pagliaro et al. (Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b) for gasparite-(Ce) and here confirmed also for monazite-(Ce). Such an intermediate structural configuration, implying an increase in the number of oxygen atoms bonded to the A site (from 9 to 10), before the occurrence of the phase transition to the post-baryte-type polymorph at higher pressures, has not been reported in earlier literature for monazite-type phosphates.

Crystal structure description

Zircon-type crystal structure

The first studies concerning the crystal structure of zircon date back to the early 20th century and were carried out independently by Vegard (Reference Vegard1916, Reference Vegard1926), Binks (Reference Binks1926), Hassel (Reference Hassel1926) and Wyckoff and Hendricks (Reference Wyckoff and Hendricks1928), in the framework of the pioneering works about the silicate’s structure determination, later gathered by Bragg (Reference Bragg1929) in his Atomic Arrangement in Silicates. After the first studies on zircon, its structural type has been described in several REE-bearing compounds, including xenotime-(Y) (Vegard, Reference Vegard1927) and the synthetic counterpart of chernovite-(Y), YAsO4 (Strada and Schwendimann, Reference Strada and Schwendimann1934).

The zircon-type structure is characterised by a tetragonal I-centred lattice (space group I41/amd). The tetragonal zircon-type structure is constructed by infinite chains of polyhedra, developed along the [001] direction (Fig. 2a and 2d), as the result of the connection, along the polyhedral edges, between the eightfold coordinated A-site dodecahedron (AO8 or REEO8) and the TO4 tetrahedra (Fig. 2c). The AO8 polyhedron displays two independent A–O atomic distances (Fig. 2c), whereas the TO4 is an undistorted tetrahedron defined by a single T–O bond distance. Each chain is in contact with four others on the (001) plane, through connecting edges along an AO8 unit and the surrounding four (Fig. 2b). The atomic coordinates of the A- and T-sites are placed in special, fixed positions, both characterised by a 2m point symmetry. The oxygen atom is also at a special position (m), being its y and z coordinates the sole refinable parameters.

Figure 2. Crystal structure of the zircon-type materials viewed (a) along the [010] and (b) [001] directions and showing (c) the chains running along the c directions and the bond distances configuration among the AO8 polyhedron and (d) a side view of the overall crystal structure. Structure drawings have been made using the software Vesta3 (Momma and Izumi, Reference Momma and Izumi2011).

Monazite-type crystal structure

Parrish (Reference Parrish1939), within the first crystallographic studies on monazites, identified its correct space group as P21/n. The first description of the monazite-type structure has been reported by Mooney (Reference Mooney1948), who investigated the La, Ce, Pr and Nd phosphates as part of the Manhattan project and described the REE atomic site in eightfold coordination. The crystal structure of monazite with the REE site in ninefold coordination has been proposed by Ueda (Reference Ueda1953, Reference Ueda1967), but with non-reliable average P–O bond lengths of ∼1.6 Å. The structure was later described correctly by Beall et al. (Reference Beall, Boatner, Mullica and Milligan1981), Mullica et al. (Reference Mullica, Milligan, Grossie, Beall and Boatner1984) and Ni et al. (Reference Ni, Hughes and Mariano1995), whereas an exhaustive review of the monazite-structure type has been carried out by Boatner (Reference Boatner, Mottana, Paolo Sassi, Thompson and Guggenheim2002) and then by Clavier et al. (Reference Clavier, Podor and Dacheux2011). The monazite-type structure can be described as made by infinite chains running along the [001] direction (c-axis), comprising the alternation of the REE-coordination polyhedra and the T-hosting tetrahedra (Fig. 3).

Figure 3. Crystal structure of monazite, viewed along (a) the [100] and (b) [010] directions; a chain-like unit is highlighted in blue; (c) coordination polyhedron of the REE-bearing A site, with 9 independent A–O bonds; and (d) general view of the monazite structure. Structure drawings have been made using the software Vesta3 (Momma and Izumi, Reference Momma and Izumi2011).

The REE-polyhedron coordination environment has nine corners (oxygen ligands, REEO9, Fig. 3c). According to Mullica et al. (Reference Mullica, Milligan, Grossie, Beall and Boatner1984), the REEO9 polyhedron can be described as an equatorial pentagon (sharing vertices with five TO4 tetrahedra of five adjacent chains in correspondence of the O1b, O2b, O2c, O3b and O4b oxygens), interpenetrated by a tetrahedron (made by the O1a, O2a, O3a and O4a oxygen atoms, see Fig. 3c), which is along the [001] direction in contact with two subsequent TO4 tetrahedra, leading to the formation of the infinite chain units (Fig. 3a,b). The REE–O2a bond length is significantly longer than the other REE–O bonds, contributing to a significant distortion of the REEO9 polyhedron (Beall et al., Reference Beall, Boatner, Mullica and Milligan1981; Ni et al., Reference Ni, Hughes and Mariano1995; Clavier et al., Reference Clavier, Podor and Dacheux2011), which can be considered as 8+1 coordinated.

Samples and experimental methods

The mineral samples of monazite-(Ce), xenotime-(Y) and chernovite-(Y), investigated in this study, originate from the same locality at Mt. Cervandone, Piedmont, Italian western Alps, where they are found as accessory phases in alpine-type fissures within hydrothermal quartz veins (Graeser and Albertini, Reference Graeser and Albertini1995) that cross-cut pegmatitic dykes (Guastoni et al., Reference Guastoni, Pezzotta and Vignola2006). The latter, intruded in leucocratic gneisses, are enriched in REE and have a strong NYF (Niobium-Yttrium-Fluorine enrichment) signature (Černý, Reference Černý1991a, Reference Černý1991b; Černý and Ercit, Reference Černý and Ercit2005). An overview of the geological background of the source rocks is described in Pagliaro et al. (Reference Pagliaro, Lotti, Comboni, Battiston, Guastoni, Fumagalli, Rotiroti and Gatta2022a), along with a detailed chemical analysis of the samples of this study by means of an electron microprobe operating in WDS mode. The experimental chemical formulas of the investigated minerals are summarised in Table 1.

Table 1. Average (and range of the measured) chemical composition (expressed in oxide wt.% and in atoms per formula unit (apfu) calculated on the basis of 4 oxygen atoms) of the chernovite-(Y), xenotime-(Y) and monazite-(Ce) samples under investigation

b.d.l. – below detection limit

In situ high-pressure experiments

In situ high-pressure single-crystal synchrotron X-ray diffraction experiments have been conducted on chernovite-(Y), xenotime-(Y) and monazite-(Ce) at the P02.2 beamline (PETRA-III synchrotron at DESY, Hamburg, Germany) and ID15B beamline (European Synchrotron Radiation Facility, Grenoble, France) using different classes of P-transmitting fluids as shown in Table 2. For all the experiments, the crystals were loaded in membrane-driven diamond anvil cells (DAC), equipped with Boehler-Almax designed diamonds/seats. Metallic foils (steel or rhenium) were pre-indented to ca. 40–70 μm and then drilled by spark-erosion to obtain P-chambers. Ruby spheres were employed as pressure calibrants (pressure uncertainty ±0.05 GPa; Mao et al., Reference Mao, Xu and Bell1986; Chervin et al., Reference Chervin, Canny and Mancinelli2001). All data collections were based on a ω-rotation with 0.5° per step and 0.5 s (monazite and chernovite) or 1 s (xenotime) of exposure time per frame. At ID15B, X-ray diffraction (XRD) data were collected using an Eiger2 9M CdTe detector positioned at 179 mm from the sample with a monochromatic 30.2 keV (λ = 0.4099 Å) beam, whereas XRD patterns at P02.2 were collected on a Perkin Elmer XRD1621 detector at 373 mm from the sample and a monochromatic incident beam with E = 42.67 keV (λ = 0.2906 Å). Further details on the beamlines setups are presented in Merlini and Hanfland (Reference Merlini and Hanfland2013) and Poreba et al. (Reference Poreba, Comboni and Mezouar2022) for ID15B and Rothkirch et al. (Reference Rothkirch, Gatta, Meyer, Merkel, Merlini and Liermann2013), Liermann et al. (Reference Liermann2015) and Bykova et al. (Reference Bykova, Aprilis, Bykov, Glazyrin, Wendt, Wenz, H-P., Roeh, Ehnes, Dubrovinskaia and Dubrovinsky2019) for P02.2. Indexing of the X-ray diffraction peaks, unit-cell refinements and intensity data reductions were performed using the CrysAlisPro package (Rigaku Oxford Diffraction, 2020). Absorption effects, due to the DAC components, were corrected using the semi-empirical ABSPACK routine, implemented in CrysAlisPro.

Table 2. Details pertaining to the in situ high-pressure and high-temperature experiments of this study

* SC-XRD: single-crystal X-ray diffraction; PXRD: powder X-ray diffraction

** He: helium (Klotz et al., Reference Klotz, J-C., Munsch and Le Marchand2009); Ne: neon (Klotz et al., Reference Klotz, J-C., Munsch and Le Marchand2009); m.e.w.: methanol:ethanol:water = 16:3:1 (Angel et al., Reference Angel, Bujak, Zhao, Gatta and Jacobsen2007)

*** See the text for further details

# Institute of Mineralogy and Petrography, University of Innsbruck, Austria

Based on the experimental intensity single-crystal XRD data, the structure refinements were performed using the Jana2020 software (Petříček et al., Reference Petříček, Palatinus, Plàšil and Dušek2023), starting from the structural models reported by Pagliaro et al. (Reference Pagliaro, Lotti, Comboni, Battiston, Guastoni, Fumagalli, Rotiroti and Gatta2022a) for the mineral samples from the same locality. The site occupancy factors of the A (lanthanide-bearing) and tetrahedral sites were fixed according to the average chemical composition obtained from EPMA-WDS analysis (Table 1), disregarding the elements with a concentration lower than 0.03 atoms per formula unit and assuming a full occupancy for both the sites. In addition, for monazite-(Ce), the atomic displacement parameters (ADP) of the oxygen atoms were refined as isotropic. All the refinements converged with no significant correlations among the refined variables. Refined structural models are deposited as crystallographic information files (cifs) and are available as Supplementary material (see below).

For a chernovite-(Y) sample with a slightly larger amount of Ca and Th replacing Y and REE, an in situ high-pressure powder XRD experiment was conducted at the P02.2 beamline of PETRA III (Hamburg, Germany) with a wavelength of λ = 0.2906 Å (42.67 keV) and a Debye-Scherrer geometry. The sample was loaded into a DAC equipped with Boehler-Almax designed diamonds of 400 μm culet size along with the pressure-transmitting medium (see Table 2 for details) and ruby spheres for pressure determination (P – uncertainty ±0.05 GPa; Mao et al., Reference Mao, Xu and Bell1986; Chervin et al., Reference Chervin, Canny and Mancinelli2001). The data collection strategy at any pressure point consisted of a 30° rotation along ω, for an exposure time of 60 s. The X-ray diffraction signals captured by the Perkin Elmer XRD1621 flat panel detector have been finalised and integrated by means of the Dioptas software (Prescher and Prakapenka, Reference Prescher and Prakapenka2015), in order to remove the background noise due to DAC components and extract the 2θ-intensity pattern for any experimental dataset. The unit-cell parameters were determined by fitting the powder XRD data by means of the Rietveld full-profile method using the GSAS-II software (Toby and Von Dreele, Reference Toby and von Dreele2013): the unit-cell parameters, crystallite size, individual scale factor and profile parameters (pseudo-Voigt function) have been refined. Moreover, the background signal has been interpolated through a Chebychev polynomial function, with 4 to 15 terms.

In situ high-temperature experiments

In situ high-temperature single-crystal X-ray diffraction data were collected at the Institute of Mineralogy and Petrography of the University of Innsbruck, using a Stoe IPDS II diffractometer equipped with a Heatstream HT device, providing a continuous flow of hot N2. The primary X-ray beam was generated by an X-ray tube (Mo-anode), operating at 50 kV and 40 mA. A plane graphite monochromator and a multiple pinhole collimator (0.5 mm) were used to guide the beam onto the sample. The image-plate detector was placed at a distance of 100 mm. The temperature calibration had been conducted previously using phase transitions of KNO3, Ag2SO4, K2SO4 and K2CrO4 powders into glass capillaries. The samples, single crystals of monazite-(Ce), xenotime-(Y) and chernovite-Y (ca. 50–120 μm3) have been inserted into SiO2 glass capillaries (0.1 mm in diameter). For all the samples, the data collection consisted in a 180° ω-rotation with a step size of 1° and variable exposure times. The temperature accuracy is ≤ 5°C. Further details about the experimental setting are reported in Krüger and Breil (Reference Krüger and Breil2009). Data collection and reduction has been performed using X-Area (Stoe and Cie, 2008). The indexed cell parameters were always compatible with either the unit cells of chernovite-(Y), xenotime-(Y) or monazite-(Ce).

Structure refinements were performed based on the single-crystal XRD data adopting the same procedure reported previously for the high-pressure data, and refined structural models are available as cifs (supplementary materials, see below). The thermal evolution of significant structural parameters (i.e. A–O bonds, A-coordination polyhedral and T-coordination polyhedral volumes) has been determined from the refined structure models, by means of the tools implemented in the VESTA3 software (Momma and Izumi, Reference Momma and Izumi2011).

In situ high-temperature X-ray powder diffraction experiments were performed on the chernovite sample relatively enriched in Ca and Th at the MCX beamline of the Elettra synchrotron (Basovizza, Trieste Italy), with a wavelength of λ = 0.7293 Å (17 keV) and a Debye-Scherrer setting. The sample, ground to powder in an agate mortar, was loaded in a SiO2 glass capillary (0.3 mm as outer diameter). For any experimental point, the data collection strategy consisted of a 2θ-scan between 8° and 60°. A step size of 0.008° was applied and an equivalent counting time for 1 s/step used. The X-ray diffraction effects were collected by the high-resolution scintillator detector available at the beamline. During the data collection, the sample was spun at a rate of 1000 rotations per minute along the φ-axis. The sample was heated by an air blower, operating between 30 and 1000°C. Further details concerning the experimental setting are reported in Rebuffi et al. (Reference Rebuffi, Plaisier, Abdellatief, Lausi and Scardi2014) and Lausi et al. (Reference Lausi, Polentarutti, Onesti, Plaisier, Busetto, Bais, Barba, Cassetta, Campi, Lamba, Pifferi, Mande, Sarma, Sharma and Paolucci2015).

Results and discussion

Compressional behaviour of the REETO4 minerals

The evolution of the unit-cell parameters of the investigated samples with pressure is reported in Table S3 and shown in Figs 4 and 5. For monazite-(Ce), the evolution of unit-cell parameters vs. P shows no evidence of phase transformation in the entire P-range investigated. Both the zircon-type minerals, on the other hand, show the occurrence of a phase transition. Chernovite-(Y), at pressures higher than ∼10 GPa is no longer stable, undergoing a phase transition to several single-crystal fragments, for which, the position of the peaks in the XRD pattern is compatible with a tetragonal scheelite-type structure. Otherwise, at pressures exceeding ∼17 GPa, xenotime-(Y) undergoes a single-crystal to single-crystal phase transition towards a monazite-type structure. For any pressure ramp, the elastic behaviour of the mineral studied has been described by fitting an isothermal Birch-Murnaghan EoS (BM-EoS), truncated at the second or third order, to the PV data (a comprehensive description of the BM-EoS formalism can be found in Angel, Reference Angel, Hazen and Downs2000) using the EoS-Fit7_GUI software (Gonzalez-Platas et al., Reference Gonzalez-Platas, Alvaro, Nestola and Angel2016). For monazite, similarly to what was observed for the isostructural gasparite-(Ce) by Pagliaro et al. (Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b), a change in the compressional behaviour was detected at ∼18 GPa. The parameters refined by fitting the experimental data by BM-EoS are reported in Table 3, among them: the bulk modulus KV = βV –1 = –V*(∂P/∂V) and its pressure derivative KV’ = (∂KV/∂P)T.

Figure 4. High-pressure evolution of the unit-cell parameters (normalised to ambient conditions values) of (a) the investigated (Ca,Th)-poor and (b) (Ca,Th)-enriched chernovite-(Y) samples and (c) of their respective normalised unit-cell volumes with the refined Birch-Murnaghan equations of state. Empty symbols refer to data collected in decompression.

Figure 5. High-pressure evolution of the unit-cell parameters (normalised to ambient-conditions values) of (a) xenotime-(Y) and (c) monazite-(Ce), (b) the normalised unit-cell volumes of the ambient-pressure and high-pressure polymorphs of xenotime-(Y) and (d) of the monoclinic β angle of monazite-(Ce).

Table 3. Refined equations of state parameters from the fit to the experimental high-pressure and high-temperature unit-cell volume data (see text for further details)

αV: volume thermal expansion coefficient at ambient conditions calculated on the basis of the Holland-Powell EoS.

a 0 and a 1 are two refinable variables in the Holland-Powell EoS (Holland and Powell, Reference Holland and Powell1998).

The evolution of significant structural parameters (i.e., A–O bonds, A-coordination polyhedral and T-coordination polyhedral volumes) with pressure has been determined, based on the refined structure models, by using the VESTA3 software (Momma and Izumi, Reference Momma and Izumi2011). The corresponding values are reported in Table S4.

The results of this study confirm that arsenates are always more compressible than the isostructural phosphates (Table 3), in agreement with the observed relationship that the larger the ionic radii of the A and T sites, the higher the bulk compressibility (e.g. Zhang et al., Reference Zhang, Zhou, Li and Li2008; Li et al., Reference Li, Zhang, Zhou and Cao2009; Lacomba-Perales et al., Reference Lacomba-Perales, Errandonea, Meng and Bettinelli2010; Errandonea et al., Reference Errandonea, Kumar, López-Solano, Rodríguez-Hernández, Muñoz, Rabie and Sáez Puche2011). The refined bulk moduli apparently suggest that monazite-structure type minerals are more compressible than those with a zircon-structure type (Table 3). However, internally consistent theoretical data (Zhang et al., Reference Zhang, Zhou, Li and Li2008; Li et al., Reference Li, Zhang, Zhou and Cao2009) show that there is a clear increase in compressibility along the lanthanoid series from Lu to La, with a discontinuity in the form of a stiffening when transforming from the zircon to the monazite structure type. Therefore, it can be concluded that, for the investigated minerals, the softening induced by the larger A site has a stronger impact than the stiffening induced by the monazite structure type.

The two minerals investigated, with the same zircon structure type, i.e. chernovite-(Y) and xenotime-(Y), undergo different phase transitions paths. Chernovite-(Y) experiences an irreversible transition, from a single-crystal to several crystal fragments, towards a scheelite-type polymorph between ca. 10.5 and 11 GPa. The same phase transition occurs, for synthetic YAsO4 in a powder XRD experiment, in a broader pressure range between 8 and 12 GPa (Errandonea et al., Reference Errandonea, Kumar, López-Solano, Rodríguez-Hernández, Muñoz, Rabie and Sáez Puche2011). The apparent discrepancy between these two studies may be ascribed to: (1) a different kinetics of the phase transition between a single crystal and a polycrystalline material; and (2) slightly different chemical compositions, as a higher phosphorous content in the presently investigated mineral decreases the average radius of the T site; or a combination of both. In general, the same phase transition has already been observed to occur in other ATO4 compounds, where the pressure of transition increases with the decrease of the A and T atoms ionic radii (Wang et al., Reference Wang, Loa, Syassen, Hanfland and Ferrand2004; Zhang et al., Reference Zhang, Wang, Lang, Zhang, Ewing and Boatner2009; Lacomba-Perales et al., Reference Lacomba-Perales, Errandonea, Meng and Bettinelli2010; Errandonea et al., Reference Errandonea, Kumar, López-Solano, Rodríguez-Hernández, Muñoz, Rabie and Sáez Puche2011). The relationship between the ionic radii of the A and T atoms and the type of structure stable at ambient and high-pressure conditions is well known and was first described by Fukunaga and Yamaoka (Reference Fukunaga and Yamaoka1979) and Bastide (Reference Bastide1987), and later discussed in other publications, e.g. in Lopez-Solano et al. (Reference López-Solano, Rodríguez-Hernández, Muñoz, Gomis, Santamaría-Perez, Errandonea, Manjón, Kumar, Stavrou and Raptis2010). As suggested by the Bastide diagram (Fig. 1) and already reported in the literature (Tatsi et al., Reference Tatsi, Stavrou, Boulmetis, Kontos, Raptis and Raptis2008; Zhang et al., Reference Zhang, Wang, Lang, Zhang, Ewing and Boatner2009; Lacomba-Perales et al., Reference Lacomba-Perales, Errandonea, Meng and Bettinelli2010; Musselman, Reference Musselman2017; Heuser et al., Reference Heuser, Palomares, Bauer, Rodriguez, Cooper, Lang, Scheinost, Schlenz, Winkler, Bosbach, Neumeier and Deissmann2018), xenotime-(Y) and isomorphous phosphates undergo a single-crystal to single-crystal phase transition towards a high-pressure polymorph, xenotime-(Y)-II, showing a monazite-type structure. The transition observed in our experiments occurs at a pressure (> 17 GPa) consistent with those reported for synthetic YPO4 compounds (Zhang et al., Reference Zhang, Zhou, Li and Li2008; Lacomba-Perales et al., Reference Lacomba-Perales, Errandonea, Meng and Bettinelli2010). The reversibility of this phase transition is confirmed by this single-crystal study, even though with a significant hysteresis, as the tetragonal polymorph is recovered in decompression only between 6.3 and 1.3 GPa.

Despite a relative scattering between the published values of bulk compressibilities of zircon- and monazite-type phosphates and arsenates, determined on the basis of experimental and theoretical studies, the elastic parameters refined for the mineral species of this study (Table 3) are in agreement with those reported in the literature (Table S1). In the zircon-type minerals chernovite-(Y) and xenotime-(Y), the bulk compression is significantly anisotropic. Indeed, the structure is approximately two times more compressible within the (001) plane than along [001] (Figs 4 and 5, Table S5), which corresponds to the direction of the polyhedral chains evolution. Such a behaviour is strongly controlled by the high tetragonal symmetry, which limits the intra-chain deformation, as confirmed by the behaviour of the two independent A–O bond distances, where the A–Oa bonds, oriented parallel to [001] (Fig. 2), are the less compressible in both the minerals (Table S4). The bulk compression is therefore mainly accommodated by the AO8 coordination polyhedron, as suggested by its bulk modulus, refined by modelling the polyhedral volume data obtained using Vesta3 with a II-BM equation of state by means of the EoS-Fit7-GUI software (Table 4). For both chernovite-(Y) and xenotime-(Y), bulk moduli values of the A-polyhedra are much lower than those obtained for the AsO4 and PO4 tetrahedra, which behave as quasi-rigid units (Table 4). The same conclusion can be drawn for monazite-(Ce) (Table 4) and gasparite-(Ce) (Pagliaro et al., Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b), which, in addition, show that the AsO4 tetrahedra are slightly more compressible than the PO4 ones. As observed previously by Pagliaro et al. (Reference Pagliaro, Lotti, Comboni, Battiston, Guastoni, Fumagalli, Rotiroti and Gatta2022a), the structural features (and responses to external T and P stimuli) of these compounds are strongly controlled by the crystal chemistry, in particular of the T sites. Moreover, not only do PO4 and AsO4 tetrahedra have a different compressional behaviour, but the nature of the prevailing T site controls the size (i.e. the average ionic radius) at ambient conditions of the A polyhedron (Pagliaro et al., Reference Pagliaro, Lotti, Comboni, Battiston, Guastoni, Fumagalli, Rotiroti and Gatta2022a). This leads, given a similar chemical composition of the A site, to slightly larger compressibility for the more ‘expanded’ A-polyhedra of the arsenates among isostructural minerals. The control exerted by the crystal chemistry on the elastic response is also pointed out by the chernovite sample relatively enriched in Ca and Th studied here, which shows a slightly lower bulk modulus value (Table 3). The lack of structure refinements prevents an unambiguous derivation of the structural mechanisms responsible for this behaviour, but it can be suggested that the more ‘expanded’ A-polyhedron (due to the higher content of the larger Th and Ca ions, as reported in Table 1) is coherent with the previous observations that a larger average ionic radius of the A site generates a softer polyhedron.

Table 4. Refined equation-of-state parameters pertaining to the A- and T-sites coordination polyhedra from the high-pressure (BM2 EoS) and high-temperature (Holland-Powell EoS, only A-site coordination polyhedron) experiments

* Due to the high uncertainties, these data should be considered as a qualitative estimation of the TO4 units rigid behaviour

The description of the elastic anisotropy in monazite-(Ce) is less straightforward, as the monoclinic symmetry does not allow us to rely on the axial compressibilities alone, given the variation of the β angle with pressure. Therefore, the finite Eulerian unit-strain tensor of monazite-(Ce) between ambient-P and 18.39 GPa has been calculated using the Win_Strain software (Angel, Reference Angel2011) and with a geometric setting with X//a* and Y//b. The results show that the principal axes of maximum and minimum unit-strain do not correspond with any of the crystallographic axes, as described by the tensor values in the following matrix:

\begin{equation*}\left( {\begin{array}{*{20}{c}} {{\varepsilon _1}} \\ {{\varepsilon _2}} \\ {{\varepsilon _3}} \end{array}} \right)\left( {\begin{array}{*{20}{c}} {158.2\left( 2 \right)^\circ }&{90^\circ }&{56.2\left( 2 \right)^\circ } \\ {90^\circ }&{180^\circ }&{90^\circ } \\ {68.2\left( 2 \right)^\circ }&{90^\circ }&{33.8\left( 2 \right)^\circ } \end{array}} \right) \cdot \left( {\begin{array}{*{20}{c}} a \\ b \\ c \end{array}} \right)\end{equation*}

The analysis of the finite Eulerian unit-strain tensor allowed the determination of the mean compressibility values along the axes of the strain ellipsoid (with ε 1>ε 2>ε 3): ε 1 = 0.003030(11) GPa–1, ε 2 = 0.00200(2) GPa–1, ε 3 = 0.0014(12) GPa–1. The directions of minimum and maximum compressibility are within the (010) plane and the anisotropic scheme is ε 1:ε 2:ε 3=2.16:1.43:1. A comparison with the finite Eulerian unit-strain tensor reported for gasparite-(Ce) by Pagliaro et al. (Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b) points out that these isostructural minerals share a similar elastic anisotropy. The shared compressional behaviour extends to the structural deformation mechanisms acting on the atomic scale: the nine independent A–O bond distances in monazite-(Ce) have different compressional evolutions, as shown in Fig. 6 (see also Table S4). A moderate scattering of the monazite-(Ce) unit-cell parameters can be observed between 10 and 18 GPa, which implied that we should fit the VP data using a Birch-Murnaghan equation of state truncated to the II-order (Table 3). On the basis of the available experimental data and structure refinements, it is not possible to unambiguously detect a change in the compressional behaviour at P < 18 GPa. On the contrary, a change in the response to compression clearly occurs at pressures exceeding ∼18.4 GPa, as evidenced by the significant deviation in the high-pressure evolution of the monoclinic β angle (Fig. 5; Table S3), similar to gasparite-(Ce) (Pagliaro et al., Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b). This behaviour has already been reported for synthetic monazite-type LaPO4 and CePO4 (Huang et al., Reference Huang, J-S., Kung and Lin2010), but not in another high-pressure investigation of CePO4 (Errandonea et al., Reference Errandonea, Gomis, Rodríguez-Hernández, Muñoz, Ruiz-Fuertes, Gupta, Achary, Hirsch, Manjon, Peters, Roth, Tyagi and Bettinelli2018). The structure refinements performed in this study allows us to draw a relationship between the change in the elastic behaviour and the structural re-configuration. Figure 6 and Table S4 show that, at lower pressures, the O3c atom is too far from the A site to be considered to effectively belong to its coordination sphere (Fig. 3). However, the A–O3c interatomic distance shows a significant shortening under compression (Fig. 6; Table S4), so that, at a pressure consistent with the change in the compressional behaviour, the O3c atom is close enough to the A site to enter its coordination sphere, as suggested by the individual bond valences calculated for selected structure refinements and reported in Table S6. As a consequence, the coordination number of the A site would increase from 9 (8+1) to 10 (8+2), even though with different contributions from the individual bonds, given the longer bond distances of the of AO2a and AO3c (Tables S4 and S6). As the same behaviour has already been independently described for gasparite-(Ce) (Pagliaro et al., Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b) and is analogous to what is described to occur at 3.25 GPa in the monazite-type crocoite (PbCrO4, Bandiello et al., Reference Bandiello, Errandonea, Martinez-Garcia, Santamaria-Perez and F.J2012; Errandonea and Kumar, Reference Errandonea and Kumar2014), we are inclined to believe that this is an intrinsic feature of the monazite structure type. Therefore, the known transition to a post-baryte-type polymorph at P higher than 26 GPa (Ruiz-Fuertes et al., Reference Ruiz-Fuertes, Hirsch, Friedrich, Winkler, Bayarjargal, Morgenroth, Peters, Roth and Milman2016) is accomplished through an intermediate structural configuration, still preserving the monazite symmetry and atomic arrangement, but characterised by an increase from 9 to 10 in the number of the oxygen atoms bonded to the lanthanide-bearing site. This intermediate structural configuration is accomplished, with no clear discontinuity, by a smooth approach of the O3c atom to the A site (where O3c corresponds to O3a and O3b in the coordination sphere of two further A atoms). It is worth noting, that the same configuration, with a 10-fold coordinated A-site, is also shared by the monazite-type high-pressure polymorph of CaSO4 (Crichton et al., Reference Crichton, Parise, Antao and Grzechnik2005). The higher pressure at which the change of the elastic behaviour occurs in monazite-(Ce) (ca. 18.4 GPa, vs. ca. 15 GPa in gasparite-(Ce), Pagliaro et al., Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b) is consistent with the higher P-stability expected for the isostructural phosphate with smaller average atomic radii.

Figure 6. (a) Bulk moduli as a function of the A-site atomic radii for several REETO4 (T = As,P,V) minerals, after Li et al. (Reference Li, Zhang, Zhou and Cao2009) and Zhang et al. (Reference Zhang, Zhou, Li and Li2008). (b) High-pressure evolution of the A–O interatomic bond distances in monazite-(Ce).

Despite being affected by larger experimental uncertainties, the same deformation mechanisms described for monazite-(Ce) and gasparite-(Ce) can be derived by the analysis of the refined structural models of the monazite-type high-pressure xenotime-(Y)-II polymorph (Table S4). In this case, the possible occurrence of a change in the compressional behaviour, according to the previous description, should be verified by investigations performed up to higher pressure values, though interatomic AO3c bond distances of ∼2.8–2.9 Å (Table S4) can suggest a 10-fold coordinated A site. A striking and anomalous feature shown by the experimental results of this study concerns the much larger refined compressibility if compared to what is reported in other studies for the same polymorph (KV = 146(5) GPa in this study; KV = 206 and 266 GPa in Zhang et al. (2009) and Lacomba-Perales et al. (Reference Lacomba-Perales, Errandonea, Meng and Bettinelli2010), respectively). The relatively high number of experimental data of this study leads us to the conclusion that the refined bulk modulus value we obtained is very reliable, implying a similar compressibility between the ambient and high-pressure polymorphs of the investigated xenotime-(Y). It is worth mentioning that a similar compressibility can also be found between the zircon- and monazite-type polymorphs of LaVO4, with KV 0 = 93(2) GPa (Yuan et al., Reference Yuan, Wang, Wang, Zhou, Yang, Liu and Zou2015) and 95(5) GPa (Errandonea et al., Reference Errandonea, Pellicer-Porres, Martínez-García, Ruiz-Fuertes, Friedrich, Morgenroth, Popescu, Rodríguez-Hernández, Muñoz and Bettinelli2016), respectively. The thorough re-investigation of the zircon-to-monazite phase transition in xenotime and of the elastic behaviour of the xenotime-II polymorphs appear, therefore, mandatory for a comprehensive understanding.

Thermal behaviour of the REETO4 minerals

The thermal unit-cell parameters evolution from the three in situ single-crystal high-T ramps on monazite-(Ce), chernovite-(Y) and xenotime-(Y) and from the powder ramp on chernovite-(Y), relatively enriched in Ca and Th, are reported in Table S7 and shown in Fig. 7. The thermo-elastic behaviour has been modelled according to the isobaric equation of state modified from Pawley et al. (Reference Pawley, Redfern and Holland1996) and Holland and Powell (Reference Holland and Powell1998) and implemented in the EoS-Fit7_GUI (Gonzalez-Platas et al., Reference Gonzalez-Platas, Alvaro, Nestola and Angel2016). The linear thermal expansion coefficients have also been refined using the TEV software (Langreiter and Kahlenberg, Reference Langreiter and Kahlenberg2015). The refined parameters and the calculated thermal expansion coefficients at ambient conditions αV = 1/V*(∂V/∂T)P are reported in Table 3.

Figure 7. High-temperature evolution of the unit-cell parameters (normalised to ambient-conditions values) of (a) (Ca,Th)-poor and (b) (Ca,Th)-enriched chernovite-(Y), (c) xenotime-(Y) and (d) monazite-(Ce).

A comparison of the linear thermal expansion coefficients refined in this study for xenotime-(Y) and monazite-(Ce) and those already published for the same compounds reveal a good agreement with the literature data (see Table S2 and references therein). However, a significant discrepancy is observed between the values refined for chernovite-(Y) in this study and those reported previously by experimental investigations of YAsO4 (Kahle, Reference Kahle, Schopper, Urban and Wüchner1970; Schopper, Reference Schopper, Urban and Ebel1972; Reddy et al., Reference Reddy, Satyanarayana Murthy and Kistaiah1988). As none of the cited references provide experimental unit-cell parameters, nor structure refinements, it is not possible to discuss such a discrepancy, but it is worth noting that the refined values for both the (Ca,Th)-poor and (Ca,Th)-enriched chernovites of this study are self-consistent and diverge from the literature data. With this in mind, future investigation of other natural or synthetic YAsO4-type zircon compounds seems imperative.

A comparative analysis of the thermo-elastic behaviour of the investigated minerals shows that monazite-(Ce) is much more expansible than the two zircon-type compounds (Table 3). In the latter, the [001] direction, i.e. the least compressible at high-P, is found to be the most expansible at high-T. The thermo-elastic anisotropy of monazite-(Ce) cannot be directly described based on the unit-cell parameters behaviour, given its monoclinic symmetry. The thermal expansion of monazite-(Ce) has been modelled with the TEV software (Langreiter and Kahlenberg, Reference Langreiter and Kahlenberg2015) and at the temperature of 400°C (i.e. ΔT ∼ 380°C) its relationship with the unit-cell axes, with a geometric setting with X//a* and Y//b, is described by the following matrix:

\begin{equation*}\left( {\begin{array}{*{20}{c}} {{\alpha _1}} \\ {{\alpha _2}} \\ {{\alpha _3}} \end{array}} \right)\,\angle \,\left( {\begin{array}{*{20}{c}} {109.02^\circ }&{90^\circ }&{5.47^\circ } \\ {19.02^\circ }&{90^\circ }&{84.53^\circ } \\ {90^\circ }&{0^\circ }&{90^\circ } \end{array}} \right)\, \cdot \,\left( {\begin{array}{*{20}{c}} a \\ b \\ c \end{array}} \right)\,\end{equation*}

The mean thermal expansivity along the axes of the unit-strain ellipsoid, determined at 400°C, is: α1 = 11.56·10–6 K–1, α2 = 9.93·10–6 K–1 and α3 = 7.39·10–6 K–1, leading to the anisotropic scheme α1: α2: α3 = 1.56:1.34:1.

The structure refinements (thermal evolution of selected structural parameters is reported in Table S8) showed that for the zircon-type minerals xenotime-(Y) and chernovite-(Y), the coordination polyhedron hosting the lanthanides ions (A-polyhedron) has a paramount role in accommodating the bulk thermal expansion, its refined thermal expansivity being almost the double of the value referred for the unit-cell volume (Tables 3, 4), whereas the two independent A–O bond distances show a comparable behaviour, different to what is shown at high-P (Table S8). In monazite-(Ce) as well, the A-polyhedron plays a significant role in accommodating the thermal expansion, but of a lesser magnitude with respect to the tetragonal minerals (Tables 3, 4), given the larger degrees of freedom for structure deformation induced by the monoclinic symmetry. In all the cases, it is worth noting that the tetrahedra, being either PO4 or AsO4, appear to behave as rigid units in the temperature range investigated (Table 4).

As a general observation based on the experimental data of this study, xenotime-(Y) appears more expansible with temperature than chernovite-(Y). Even though the discrepancy between our data and the previously published thermal expansion behaviour of other YAsO4 compounds (Kahle, Reference Kahle, Schopper, Urban and Wüchner1970; Schopper, Reference Schopper, Urban and Ebel1972; Reddy et al., Reference Reddy, Satyanarayana Murthy and Kistaiah1988) suggests caution in this regards. However, our observation is consistent with the results reported by Li et al. (Reference Li, Zhang, Zhou and Cao2009) for APO4 and AAsO4 (A = lanthanides) and based on theoretical calculations of lattice energies, where phosphates always show a larger expansibility than isostructural arsenates sharing the same A cation for both the zircon and monazite structural types.

Concluding remarks

The experimental data reported in this study, along with those published by Pagliaro et al. (Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b) on the high-pressure behaviour of gasparite-(Ce), provide a suite of thermo-elastic parameters for natural lanthanide-bearing phosphates and arsenates cropping out in the hydrothermal veins of the Mt. Cervandone.

Consistently with the previous scientific literature, the two zircon-type minerals undergo different P-induced phase transitions. Chernovite-(Y), above ∼10 GPa converts to a scheelite-type polymorph, whereas xenotime-(Y), above ∼17 GPa, transforms into a monazite-type polymorph by a reconstructive phase transition, single-crystal to single-crystal in character.

The monazite-type phosphates and arsenates do not undergo any phase transitions in the explored pressure range of this study. However, the analysis of the compressional evolution of the unit-cell parameters and of the crystal structures of monazite-(Ce) (this study) and gasparite-(Ce) from the same locality (Pagliaro et al., Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b), highlighted a change in the compressional behaviour occurring at ∼18 and 15 GPa, respectively. This is related, for both the minerals, to a structural re-arrangement involving the smooth approach of a tenth oxygen atom into the coordination sphere of the lanthanide-bearing A site, which increases its coordination from 9 (8+1) to 10 (8+2) before the expected phase transition to a high-pressure polymorph. A general conclusion from both the present experiments and previous studies is that, despite the numerous papers publishing high-quality data on these crystalline compounds relevant to Earth and Materials sciences, there are still unexplored regions, whose physical-chemical features should be described by the adoption of up-to-date experimental techniques, facilities and crystallographic methods.

The analysis of the elastic and structural response with pressure and temperature of the investigated minerals of this study, combined with those on gasparite-(Ce) (Pagliaro et al., Reference Pagliaro, Lotti, Guastoni, Rotiroti, Battiston and Gatta2022b), lead us to conclude that their compressional and thermal behaviours are not, as commonly observed, induced by the same mechanisms opposite in sign. For example, phosphates are less compressible than the isostructural arsenates, but at high temperature they are more expansible. In the same way, the zircon-type structure is less compressible along the c crystallographic direction, but at high-T along the same direction the higher expansivity is shown. This behaviour can be explained by the significant control exerted on these compounds by the chemical strain. As described by Pagliaro et al. (Reference Pagliaro, Lotti, Comboni, Battiston, Guastoni, Fumagalli, Rotiroti and Gatta2022a), the chemical nature of the T cations (P or As) has a paramount role in controlling not only the TO4 tetrahedra, but also the other structural parameters: the substitution of P by As not only expands the tetrahedron, but the A coordination polyhedron too (at the same chemical composition of the A site). This study shows that the chemically more ‘expanded’ structures of the arsenates are more compressible than the chemically more ‘compressed’ structures of phosphates, but in response to a thermal perturbation they show the opposite behaviour being less expansible. Such an observation is confirmed by the slightly different behaviours of the two investigated chernovite samples. The sample relatively enriched in Th and Ca, due to the larger ionic radii of these cations with respect to the dominant Y in the A site, show a chemical expansion at ambient conditions (as described in Pagliaro et al., Reference Pagliaro, Lotti, Comboni, Battiston, Guastoni, Fumagalli, Rotiroti and Gatta2022a) that is reflected by a slightly larger compressibility at high pressure and a slightly lower expansivity at high temperature.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1180/mgm.2024.70.

Acknowledgements

This manuscript is dedicated to the memory of our friend and colleague Dr. Alessandro Guastoni (1966–2022), who has given a significant contribution in the conceptualization of this study. FP, GDG and PL are grateful to Alessandro for the numerous and fruitful discussions on this and many other topics. The editors and two anonymous reviewers are acknowledged for the handling of the manuscript and for the fruitful and insightful comments. We acknowledge DESY (Hamburg, Germany), a member of the Helmholtz Association HGF, for the provision of experimental facilities. Parts of this research were carried out at PETRA III. Beamtime was allocated for proposals I-20210452 EC and I-20211179 EC. We acknowledge Elettra Sincrotrone Trieste for providing access to its synchrotron radiation facilities. We acknowledge the European Synchrotron Radiation Facility (ESRF) for provision of synchrotron radiation facilities. The research leading to this result has been supported by the project CALIPSOplus under Grant Agreement 730872 from the EU Framework Programme for Research and Innovation HORIZON 2020. FP, DC, TB, GDG and PL acknowledge the Italian Ministry of University for the support through the project “Dipartimenti di Eccellenza 2023–2027”. DC, GDG and PL acknowledge the University of Milan for the support through the project “Piano di sostegno alla Ricerca 2022”.

Competing interests

The authors declare none.

Footnotes

Guest Editor: Fabrizio Nestola

Dedicated to the memory of Dr. Alessandro Guastoni

References

Angel, R.J. (2000) Equations of State. Pp. 3559 in: High-Temperature and High Pressure Crystal Chemistry (Hazen, Robert M. and Downs, Robert T., editors). Reviews in Mineralogy and Geochemistry, 41. The Mineralogical Society of America and the Geochemical Society, Chantilly, Virginia, USA, doi: 10.2138/rmg.2000.41.2Google Scholar
Angel, R.J. (2011) Win_Strain: A program to calculate strain tensors from unit-cell parameters. http://www.rossangel.com/home.htmGoogle Scholar
Angel, R.J., Bujak, M., Zhao, J., Gatta, G.D. and Jacobsen, S.D. (2007) Effective hydrostatic limits of pressure media for high-pressure crystallographic studies. Journal of Applied Crystallography, 40, 2632, doi: 10.1107/S0021889806045523.Google Scholar
Bandiello, E., Errandonea, D., Martinez-Garcia, D., Santamaria-Perez, D. and F.J, Manjón. (2012) Effects of high-pressure on the structural, vibrational, and electronic properties of monazite-type PbCrO4. Physical Review B, 85, , doi: 10.1103/PhysRevB.85.024108.Google Scholar
Bastide, J.P. (1987) Systématique simplifiée des composés ABX4 (X = O2−, F) et evolution possible de leurs structures cristallines sous pression. Journal of Solid State Chemistry, 71, 115120. doi: 10.1016/0022-4596(87)90149-6Google 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. doi: 10.1016/0022-1902(81)80443-5Google Scholar
Binks, W. (1926) The crystalline structure of zircon. Mineralogical Magazine and Journal of the Mineralogical Society, 21, 176187, doi: 10.1180/minmag.1926.021.115.06Google Scholar
Boatner, L.A. (2002) Synthesis, structure, and properties of monazite, pretulite, and xenotime. Pp. 87121 in: Micas: Crystal Chemistry & Metamorphic Petrology (Mottana, Annibale, Paolo Sassi, Francesco, Thompson, James B. Jr. and Guggenheim, Stephen, editors). Reviews in Mineralogy and Geochemistry, 48. The Mineralogical Society of America and the Geochemical Society, Chantilly, Virginia, USA, doi: 10.2138/rmg.2002.48.4Google Scholar
Bragg, W.L. (1929) Atomic arrangement in the silicates. Transactions of the Faraday Society, 25, , doi: 10.1039/TF9292500291Google Scholar
Brown, I.D. (2002) The Chemical Bond in Inorganic Chemistry: The Bond Valence Model. Oxford University Press, Oxford, pp.Google Scholar
Bykova, E., Aprilis, G., Bykov, M., Glazyrin, K., Wendt, M., Wenz, S., H-P., Liermann, Roeh, J.T., Ehnes, A., Dubrovinskaia, N. and Dubrovinsky, L. (2019) Single-crystal diffractometer coupled with double-sided laser heating system at the Extreme Conditions Beamline P02.2 at PETRAIII. Review of Scientific Instruments, 90, , doi: 10.1063/1.5108881Google Scholar
Černý, P. (1991a) Rare-element granitic pegmatites. Part I: Anatomy and internal evolution of pegmatite deposits. Geoscience Canada, 18, 4967Google Scholar
Černý, P. (1991b) Rare-element granitic pegmatites. Part II: Regional to global environments and petrogenesis. Geoscience Canada, 18, 6881Google Scholar
Černý, P. and Ercit, T.S. (2005) The classification of granitic pegmatites revisited. The Canadian Mineralogist, 43, 20052026, doi: 10.2113/gscanmin.43.6.2005Google Scholar
Chervin, J.C., Canny, B. and Mancinelli, M. (2001) Ruby-spheres as pressure gauge for optically transparent high pressure cells. High Pressure Research, 21, 305314, doi: 10.1080/08957950108202589Google Scholar
Clavier, N., Podor, R. and Dacheux, N. (2011) Crystal chemistry of the monazite structure. Journal of the European Ceramic Society, 31, 941976, doi: 10.1016/j.jeurceramsoc.2010.12.019Google 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, doi: 10.2138/am.2005.1654.Google Scholar
Crichton, W.A., Merlini, M., Müller, H., Chantel, J. and Hanfland, M. (2012) The high-pressure monazite-to-scheelite transformation in CaSeO4. Mineralogical Magazine, 76, 913923. doi: 10.1180/minmag.2012.076.4.08Google Scholar
Davis, J.B., Marshall, D.B., Housley, R.M. and Morgan, P.E.D. (1998) Machinable Ceramics Containing Rare‐Earth Phosphates. Journal of the American Ceramic Society, 81, 21692175, doi: 10.1111/j.1151–2916.1998.tb02602.xGoogle Scholar
Errandonea, D., Kumar, R., López-Solano, J., Rodríguez-Hernández, P., Muñoz, A., Rabie, M.G. and Sáez Puche, R. (2011) Experimental and theoretical study of structural properties and phase transitions in YAsO4 and YCrO4. Physical Review B, 83, , doi: 10.1103/PhysRevB.83.134109Google Scholar
Errandonea, D. and Kumar, R.S. (2014) High-pressure structural transformations of PbCrO4 up to 51.2 GPa: An angle-dispersive synchrotron X-ray diffraction study. Materials Research Bulletin, 60, 206211, doi: 10.1016/j.materresbull.2014.08.041Google Scholar
Errandonea, D., Pellicer-Porres, J., Martínez-García, D., Ruiz-Fuertes, J., Friedrich, A., Morgenroth, W., Popescu, C., Rodríguez-Hernández, P., Muñoz, A. and Bettinelli, M. (2016) Phase stability of lanthanum orthovanadate at high pressure. Journal of Physical Chemistry C, 120, 1374913762, doi: 10.1021/acs.jpcc.6b04782Google Scholar
Errandonea, D. (2017) High‐pressure phase transitions and properties of MTO4 compounds with the monazite‐type structure. Physica Status Solidi B, 254, , doi: 10.1002/pssb.201700016Google Scholar
Errandonea, D., Gomis, O., Rodríguez-Hernández, P., Muñoz, A., Ruiz-Fuertes, J., Gupta, M., Achary, S.N., Hirsch, A., Manjon, F.J., Peters, L., Roth, G., Tyagi, A.K. and Bettinelli, M. (2018) High-pressure structural and vibrational properties of monazite-type BiPO4, LaPO4, CePO4, and PrPO4. Journal of Physics: Condensed Matter, 30, , doi: 10.1088/1361-648X/aaa20dGoogle Scholar
Finch, R.J. and Hanchar, J.M.(2003) Structure and chemistry of zircon and zircon-group minerals. Pp. 125 in: Zircon (Hanchar, John M. and Hoskin, Paul W.O., editors). Reviews in Mineralogy and Geochemistry, 53. The Mineralogical Society of America and the Geochemical Society, Chantilly, Virginia, USA, doi: 10.2113/0530001Google Scholar
Fukunaga, O. and Yamaoka, S. (1979) Phase transformations in ABO4 type compounds under high pressure. Physics and Chemistry of Minerals, 5, 167177, doi: 10.1007/BF00307551Google Scholar
Gleissner, J., Errandonea, D., Segura, A., Pellicer-Porres, J., Hakeem, M.A., Proctor, J.E., Raju, S.V., Kumar, R.S., Rodríguez-Hernández, P., Muñoz, A., Lopez-Moreno, S. and Bettinelli, M. (2016) Monazite-type SrCrO4 under compression. Physical Review B, 94, , doi: 10.1103/PhysRevB.94.134108Google Scholar
Gonzalez-Platas, J., Alvaro, M., Nestola, F. and Angel, R.J. (2016) EosFit7-GUI: a new graphical user interface for equation of state calculations, analyses and teaching. Journal of Applied Crystallography, 49, 13771382, doi: 10.1107/S1600576716008050.Google Scholar
Graeser, S. and Albertini, C. (1995) Wannigletscher und Conca Cervandone. Lapis, 20, 4164.Google Scholar
Guastoni, A., Pezzotta, F. and Vignola, P. (2006) Characterization and genetic inferences of arsenates, sulfates and vanadates of Fe, Cu, Pb, Zn from Mount Cervandone (Western Alps, Italy). Periodico di Mineralogia, 75, 141150.Google Scholar
Harrison, T.M., Catlos, E.J. and Montel, J-M. (2002) U-Th-Pb dating of phosphate minerals. Pp. 524558 in: Micas: Crystal Chemistry & Metamorphic Petrology (Mottana, Annibale, Paolo Sassi, Francesco, Thompson, James B. Jr. and Guggenheim, Stephen, editors). Reviews in Mineralogy and Geochemistry, 48. The Mineralogical Society of America and the Geochemical Society, Chantilly, Virginia, USA, doi: 10.2138/rmg.2002.48.14.Google Scholar
Hassel, O. (1926) XIV. Die Kristallstruktur einiger Verbindungen von der Zusammensetzung MRO4. I. Zirkon ZrSiO4. Zeitschrift für Kristallographie – Crystalline Materials, 63, 247254, doi: 10.1524/zkri.1926.63.1.247Google Scholar
Hay, R.S., Mogilevsky, P. and Boakye, E. (2013) Phase transformations in xenotime rare-earth orthophosphates. Acta Materialia, 61, 69336947, doi: 10.1016/j.actamat.2013.08.005Google Scholar
Heuser, J.M., Palomares, R.I., Bauer, J.D., Rodriguez, M.L., Cooper, J., Lang, M., Scheinost, A.C., Schlenz, H., Winkler, B., Bosbach, D., Neumeier, S. and Deissmann, G. (2018) Structural characterization of (Sm,Tb)PO4 solid solutions and pressure-induced phase transitions. Journal of the European Ceramic Society, 38, 40704081, doi: 10.1016/j.jeurceramsoc.2018.04.030Google Scholar
Holland, T.J.B. and Powell, R. (1998) An internally consistent thermodynamic data set for phases of petrological interest. Journal Metamorphic Geology, 16, 309343, doi: 10.1111/j.1525-1314.1998.00140.xGoogle Scholar
Huang, T., J-S., Lee, Kung, J. and Lin, C-M. (2010) Study of monazite under high pressure. Solid State Communications, 150, 18451850, doi: 10.1016/j.ssc.2010.06.042Google Scholar
Jayaraman, A., Kourouklis, G.A., Espinosa, G.P., Cooper, A.S. and van Uitert, L.G. (1987) A high-pressure Raman study of yttrium vanadate (YVO4) and the pressure-induced transition from the zircon-type to the scheelite-type structure. Journal of Physics and Chemistry of Solids, 48, 755759, doi: 10.1016/0022-3697(87)90072-2Google Scholar
Kahle, H.G., Schopper, H.C., Urban, W. and Wüchner, W. (1970) Temperature effects on zircon structure lattice parameters and zero‐field resonance for substituted Gd3+. Physica Status Solidi B, 38, 815819, doi: 10.1002/pssb.19700380231Google Scholar
Klotz, S., J-C., Chervin, Munsch, P. and Le Marchand, G. (2009) Hydrostatic limits of 11 pressure transmitting media. Journal of Physics D, 42, , http://dx.doi.org/10.1088/0022-3727/42/7/075413Google 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, doi: 10.1127/0935-1221/2004/0016-0117Google Scholar
Krüger, H. and Breil, L. (2009) Computer-controlled high-temperature single-crystal X-ray diffraction experiments and temperature calibration. Journal of Applied Crystallography, 42, 140142, doi: 10.1107/S0021889808035607Google Scholar
Lacomba-Perales, R., Errandonea, D., Meng, Y. and Bettinelli, M. (2010) High-pressure stability and compressibility of APO4 (A=La, Nd, Eu, Gd, Er, and Y) orthophosphates: An X-ray diffraction study using synchrotron radiation. Physical Review B, 81, , doi: 10.1103/PhysRevB.81.064113Google Scholar
Langreiter, T. and Kahlenberg, V. (2015) TEV—A program for the determination of the thermal expansion tensor from diffraction data. Crystals, 5, 143153, doi: 10.3390/cryst5010143Google Scholar
Lausi, A., Polentarutti, M., Onesti, S., Plaisier, J.R., Busetto, E., Bais, G., Barba, L., Cassetta, A., Campi, G., Lamba, D., Pifferi, A., Mande, S.C., Sarma, D.D., Sharma, S.M. and Paolucci, G. (2015) Status of the crystallography beamlines at Elettra. European Physical Journal Plus, 130, , doi: 10.1140/epjp/i2015-15043-3Google Scholar
Li, H., Zhang, S., Zhou, S. and Cao, X. (2009) Bonding characteristics, thermal expansibility, and compressibility of RXO(4) (R = rare earths, X = P, As) within monazite and zircon structures. Inorganic Chemistry, 48, 45424548, doi: 10.1021/ic900337jGoogle Scholar
Liermann, H-P. et al. (2015) The Extreme Conditions Beamline P02.2 and the Extreme Conditions Science Infrastructure at PETRA III. Journal of Synchrotron Radiation, 22, , doi: 10.1107/s1600577515005937.Google Scholar
López-Solano, J., Rodríguez-Hernández, P., Muñoz, A., Gomis, O., Santamaría-Perez, D., Errandonea, D., Manjón, F.J., Kumar, R.S., Stavrou, E. and Raptis, C. (2010) Theoretical and experimental study of the structural stability of TbPO4 at high pressures. Physical Review B, 81:21, doi: 10.1103/PhysRevB.81.144126Google Scholar
Manjón, F.J., Rodríguez-Hernández, P., Muñoz, A., Romero, A.H., Errandonea, D., Syassen, K. (2010) Lattice dynamics of YVO4 at high pressures. Physical Review B, 81, . doi: 10.1103/PhysRevB.81.075202Google Scholar
Mao, H.K., Xu, J., Bell, P.M. (1986) Calibration of the ruby pressure gauge to 800 kbar under quasi‐hydrostatic conditions. Journal of Geophysical Research, 91, 46734676, doi: 10.1029/JB091iB05p04673Google Scholar
Marqueño, T., Errandonea, D., Pellicer-Porres, J., Santamaria-Perez, D., Martinez-Garcia, D., Bandiello, E., Rodriguez-Hernandez, P., Muñoz, A., Achary, S.N., Popescu, C. (2021) Polymorphism of praseodymium orthovanadate under high pressure. Physical Review B, 103, doi: 10.1103/PhysRevB.103.134113Google Scholar
Merlini, M. and Hanfland, M. (2013) Single-crystal diffraction at megabar conditions by synchrotron radiation. High Pressure Research 33:511522. doi: 10.1080/08957959.2013.831088Google Scholar
Momma, K. and Izumi, F. (2011) VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. Journal of Applied Crystallography, 44, 12721276, doi: 10.1107/S0021889811038970Google Scholar
Mooney, R.C.L. (1948) Crystal structures of a series of rare earth phosphates. The Journal of Chemical Physics, 16, , doi: 10.1063/1.1746668Google Scholar
Morgan, P.E.D., Marshall, D.B. and Housley, R.M. (1995) High-temperature stability of monazite-alumina composites. Materials Science and Engineering: A, 195, 215222, doi: 10.1016/0921-5093(94)06521-7Google Scholar
Mullica, D.F., Milligan, W.O., Grossie, D.A., Beall, G.W. and Boatner, L.A. (1984) Ninefold coordination LaPO4: Pentagonal interpenetrating tetrahedral polyhedron. Inorganica Chimica Acta, 95, 231236, doi: 10.1016/S0020-1693(00)87472-1Google Scholar
Musselman, M.A. (2017) In situ Raman spectroscopy of pressure-induced phase transformations in DyPO4 and GdxDy (1–x)PO4. MSc thesis, Colorado School of Mines, USA.Google Scholar
Ni, Y., Hughes, J.M., Mariano, A.N. (1995) Crystal chemistry of the monazite and xenotime structures. American Mineralogist, 80, 2126, doi: 10.2138/am-1995-1-203Google Scholar
Oelkers, E.H. and Montel, J-M. (2008) Phosphates and nuclear waste storage. Elements, 4, 113116, doi: 10.2113/GSELEMENTS.4.2.113Google Scholar
Orlova, A.I. and Ojovan, M.I. (2019) Ceramic mineral waste-forms for nuclear waste immobilization. Materials (Basel), 12, doi: 10.3390/ma12162638Google Scholar
Pagliaro, F., Lotti, P., Comboni, D., Battiston, T., Guastoni, A., Fumagalli, P., Rotiroti, N. and Gatta, G.D. (2022a) High-pressure behaviour of gasparite-(Ce) (nominally CeAsO4), a monazite-type arsenate. Physics and Chemistry of Minerals, 49, , doi: 10.1007/s00269-022-01222-5Google Scholar
Pagliaro, F., Lotti, P., Guastoni, A., Rotiroti, N., Battiston, T. and Gatta, G.D. (2022b) Crystal chemistry and miscibility of chernovite-(Y), xenotime-(Y), gasparite-(Ce) and monazite-(Ce) from Mt. Cervandone, Western Alps, Italy. Mineralogical Magazine, 86, 150167, doi: 10.1180/mgm.2022.5Google Scholar
Panchal, V., Errandonea, D., Manjón, F.J., Muñoz, A., Rodríguez-Hernández, P., Achary, S.N. and Tyagi, A.K. (2017) High-pressure lattice-dynamics of NdVO4. Journal of Physics and Chemistry of Solids, 100, 126133, doi: 10.1016/j.jpcs.2016.10.001Google Scholar
Parrish, W. (1939) Unit cell and space group of monazite, (La,Ce,Y)PO4. American Mineralogist, 24, 651652.Google Scholar
Pawley, A.R., Redfern, S.A.T. and Holland, T.J.B. (1996) Volume behaviour of hydrous minerals at high pressure and temperature; I, Thermal expansion of lawsonite, zoisite, clinozoisite, and diaspore. American Mineralogist, 81, 335340, doi: 10.2138/am-1996-3-407Google Scholar
Perrière, L., Bregiroux, D., Naitali, B., Audubert, F., Champion, E., Smith, D.S. and Bernache-Assollant, D. (2007) Microstructural dependence of the thermal and mechanical properties of monazite LnPO4 (Ln=La to Gd). Journal of the European Ceramic Society, 27, 32073213, doi: 10.1016/j.jeurceramsoc.2006.12.005Google Scholar
Petříček, V., Palatinus, L., Plàšil, J. and Dušek, M. (2023) JANA2020 – a new version of the crystallographic computing system JANA. Zeitschrift für Kristallographie – Crystalline Materials, 238, 271282. doi: 10.1515/zkri-2023-0005Google Scholar
Poreba, T., Comboni, D. and Mezouar, M., et al (2022) Tracking of structural phase transitions via single crystal x-ray diffraction at extreme conditions: advantages of extremely brilliant source. Journal of Physics: Condensed Matter, 35, doi: 10.1088/1361-648X/aca50bGoogle Scholar
Prescher, C. and Prakapenka, V.B. (2015) DIOPTAS: a program for reduction of two-dimensional X-ray diffraction data and data exploration. High Pressure Research, 35, 223230, doi: 10.1080/08957959.2015.1059835Google Scholar
Rapp, R.P. and Watson, E.B. (1986) Monazite solubility and dissolution kinetics: implications for the thorium and light rare earth chemistry of felsic magmas. Contributions to Mineralogy and Petrology, 94, 304316, doi: 10.1007/BF00371439Google Scholar
Rebuffi, L., Plaisier, J.R., Abdellatief, M., Lausi, A. and Scardi, P. (2014) MCX: a synchrotron radiation beamline for X‐ray diffraction line profile analysis. Zeitschrift für anorganische und allgemeine Chemie, 640, 31003106. doi: 10.1002/zaac.201400163Google Scholar
Reddy, C.V.V., Satyanarayana Murthy, K. and Kistaiah, P. (1988) X-ray study of the thermal expansion anisotropy in YVO4 and YAsO4 compounds. Solid State Communications, 67, 545547. doi: 10.1016/0038-1098(84)90179-0Google Scholar
Rigaku Oxford Diffraction (2020) CrysAlisPro version 171.41.93. Wroclaw, PolandGoogle Scholar
Rothkirch, A., Gatta, G.D., Meyer, M., Merkel, S., Merlini, M. and Liermann, H.P. (2013) Single-crystal diffraction at the Extreme Conditions beamline P02.2: procedure for collecting and analyzing high-pressure single-crystal data. Journal of Synchrotron Radiation, 20, 711720. doi: 10.1107/S0909049513018621Google Scholar
Ruiz-Fuertes, J., Hirsch, A., Friedrich, A., Winkler, B., Bayarjargal, L., Morgenroth, W., Peters, L., Roth, G. and Milman, V. (2016) High-pressure phase of LaPO4 studied by X-ray diffraction and second harmonic generation. Physical Review B, 94, . doi: 10.1103/PhysRevB.94.134109Google Scholar
Schopper, H.C., Urban, W. and Ebel, H. (1972) Measurements of the temperature dependence of the lattice parameters of some rare earth compounds with zircon structure. Solid State Communications, 11, 955958, doi: 10.1016/0038-1098(72)90297-9Google Scholar
Sousa Filho, P.C. de and Serra, O.A. (2009) Red, green, and blue lanthanum phosphate phosphors obtained via surfactant-controlled hydrothermal synthesis. Journal of Luminescence, 129, 16641668, doi: 10.1016/j.jlumin.2009.04.075Google Scholar
Stangarone, C., Angel, R.J., Prencipe, M., Mihailova, B. and Alvaro, M. (2019) New insights into the zircon-reidite phase transition. American Mineralogist, 104, 830837.Google Scholar
Stoe and Cie (2008) WinXpose 1.7.6. Darmstadt, Germany.Google Scholar
Strada, M. and Schwendimann, G. (1934) La struttura cristallina di alcuni fosfati ed arseniati di metalli trivalenti. II. Arseniato e fosfato di yttrio. Gazzetta Chimica Italiana, 1934, 662674.Google Scholar
Strzelecki, A.C., Zhao, X., Estevenon, P., Xu, H., Dacheux, N., Ewing, R.C. and Guo, X. (2024) Crystal chemistry and thermodynamic properties of zircon structure-type materials. American Mineralogist, 109, 225242, doi: 10.2138/am-2022-8632Google Scholar
Subbarao, E.C., Agrawal, D.K., McKinstry, H.A., Sallese, C.W. and Roy, R. (1990) Thermal Expansion of Compounds of Zircon Structure. Journal of the American Ceramic Society, 73, 12461252, doi: 10.1111/j.1151-2916.1990.tb05187.xGoogle Scholar
Tatsi, A., Stavrou, E., Boulmetis, Y.C., Kontos, A.G., Raptis, Y.S. and Raptis, C. (2008) Raman study of tetragonal TbPO4 and observation of a first-order phase transition at high pressure. Journal of Physics: Condensed Matter, 20, , doi: 10.1088/0953-8984/20/42/425216Google Scholar
Toby, B.H. and von Dreele, R.B. (2013) GSAS-II: the genesis of a modern open-source all purpose crystallography software package. Journal of Applied Crystallography, 46, 544549, doi: 10.1107/S0021889813003531Google Scholar
Ueda, T. (1953) The crystal structure of monazite (CePO4). Memoirs of the College of Science, University of Kyoto, Series B, 1953, 227246.Google Scholar
Ueda, T. (1967) Reexamination of the crystal structure of monazite. The Journal of the Japanese Association of Mineralogists, Petrologists and Economic Geologists, 58, 170179, doi: 10.2465/ganko1941.58.170Google Scholar
Ushakov, S.V., Helean, K.B., Navrotsky, A. and Boatner, L.A. (2001) Thermochemistry of rare-earth orthophosphates. Journal of Materials Research, 16, 26232633, doi: 10.1557/JMR.2001.0361Google Scholar
Vegard, L. (1916) VI. Results of crystal analysis. The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science, 32, 6596, doi: 10.1080/14786441608635544Google Scholar
Vegard, L. (1926) CIV. Results of crystal analysis. The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science, 1, 11511193, doi: 10.1080/14786442608633716Google Scholar
Vegard, L. (1927) XLVII. The structure of xenotime and the relation between chemical constitution and crystal structure. The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science, 4, 511525, doi: 10.1080/14786440908564357Google Scholar
Vorres, K.S. (1962) Correlating ABO4 compound structures. Journal of Chemical Education, 39, , doi: 10.1021/ed039p566Google Scholar
Wang, X., Loa, I., Syassen, K., Hanfland, M. and Ferrand, B. (2004) Structural properties of the zircon- and scheelite-type phases of YVO4 at high pressure. Physical Review B, 70, , doi: 10.1103/PhysRevB.70.064109Google Scholar
Wyckoff, R.W.G. and Hendricks, S.B. (1928) IV. Die Kristallstruktur von Zirkon und die Kriterien für spezielle Lagen in tetragonalen Raumgruppen. Zeitschrift für Kristallographie - Crystalline Materials, 66, 73102, doi: 10.1524/zkri.1928.66.1.73Google Scholar
Yuan, H., Wang, K., Wang, C., Zhou, B., Yang, K., Liu, J. and Zou, B. (2015) Pressure-Induced Phase Transformations of Zircon-Type LaVO4 Nanorods. Journal of Physical Chemistry C, 119, 83648372, doi: 10.1021/acs.jpcc.5b01007Google Scholar
Zhang, S., Zhou, S., Li, H. and Li, L. (2008) Investigation of thermal expansion and compressibility of rare-earth orthovanadates using a dielectric chemical bond method. Inorganic Chemistry, 47, 78637867, doi: 10.1021/ic800672hGoogle Scholar
Zhang, F.X., Wang, J.W., Lang, M., Zhang, J.M., Ewing, R.C. and Boatner, L.A. (2009) High-pressure phase transitions of ScPO4 and YPO4. Physical Review B, 80, , doi: 10.1103/PhysRevB.80.184114Google Scholar
Figure 0

Figure 1. The so-called ‘Bastide diagram’ showing the relationships among structural types as a function of the atomic radii of cations at the A site (rA), T site (rT) and oxygen (rO), within the ATO4 family. The fields corresponding to the SrUO4 and BaWO4-II structures are labelled as orthorhombic (Cmca, Pbcm, Pnma) and monoclinic 14, respectively (2, 10, 12, 14 refer to the space group numbers). The post-baryte field is not reported (modified after López-Solano et al., 2010).

Figure 1

Figure 2. Crystal structure of the zircon-type materials viewed (a) along the [010] and (b) [001] directions and showing (c) the chains running along the c directions and the bond distances configuration among the AO8 polyhedron and (d) a side view of the overall crystal structure. Structure drawings have been made using the software Vesta3 (Momma and Izumi, 2011).

Figure 2

Figure 3. Crystal structure of monazite, viewed along (a) the [100] and (b) [010] directions; a chain-like unit is highlighted in blue; (c) coordination polyhedron of the REE-bearing A site, with 9 independent A–O bonds; and (d) general view of the monazite structure. Structure drawings have been made using the software Vesta3 (Momma and Izumi, 2011).

Figure 3

Table 1. Average (and range of the measured) chemical composition (expressed in oxide wt.% and in atoms per formula unit (apfu) calculated on the basis of 4 oxygen atoms) of the chernovite-(Y), xenotime-(Y) and monazite-(Ce) samples under investigation

Figure 4

Table 2. Details pertaining to the in situ high-pressure and high-temperature experiments of this study

Figure 5

Figure 4. High-pressure evolution of the unit-cell parameters (normalised to ambient conditions values) of (a) the investigated (Ca,Th)-poor and (b) (Ca,Th)-enriched chernovite-(Y) samples and (c) of their respective normalised unit-cell volumes with the refined Birch-Murnaghan equations of state. Empty symbols refer to data collected in decompression.

Figure 6

Figure 5. High-pressure evolution of the unit-cell parameters (normalised to ambient-conditions values) of (a) xenotime-(Y) and (c) monazite-(Ce), (b) the normalised unit-cell volumes of the ambient-pressure and high-pressure polymorphs of xenotime-(Y) and (d) of the monoclinic β angle of monazite-(Ce).

Figure 7

Table 3. Refined equations of state parameters from the fit to the experimental high-pressure and high-temperature unit-cell volume data (see text for further details)

Figure 8

Table 4. Refined equation-of-state parameters pertaining to the A- and T-sites coordination polyhedra from the high-pressure (BM2 EoS) and high-temperature (Holland-Powell EoS, only A-site coordination polyhedron) experiments

Figure 9

Figure 6. (a) Bulk moduli as a function of the A-site atomic radii for several REETO4 (T = As,P,V) minerals, after Li et al. (2009) and Zhang et al. (2008). (b) High-pressure evolution of the A–O interatomic bond distances in monazite-(Ce).

Figure 10

Figure 7. High-temperature evolution of the unit-cell parameters (normalised to ambient-conditions values) of (a) (Ca,Th)-poor and (b) (Ca,Th)-enriched chernovite-(Y), (c) xenotime-(Y) and (d) monazite-(Ce).

Supplementary material: File

Pagliaro et al. supplementary material 1

Pagliaro et al. supplementary material
Download Pagliaro et al. supplementary material 1(File)
File 3.3 MB
Supplementary material: File

Pagliaro et al. supplementary material 2

Pagliaro et al. supplementary material
Download Pagliaro et al. supplementary material 2(File)
File 28.1 KB
Supplementary material: File

Pagliaro et al. supplementary material 3

Pagliaro et al. supplementary material
Download Pagliaro et al. supplementary material 3(File)
File 88.4 KB