Next Article in Journal
Mechanism of Rake Frame Shear Drainage during Gravity Dewatering of Ultrafine Unclassified Tailings for Paste Preparation
Next Article in Special Issue
The Role of Scandium Substitution in Babingtonite Group Minerals
Previous Article in Journal
Effect and Mechanism of CaO on Iron Recovery and Desulfurization by Reduction Roasting-Magnetic Separation of High-Sulfur Cyanide Tailings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Breakdown of Natural Epidote and Clinozoisite in High-T and Low-P Conditions and Characterization of Its Products

1
Department of Mineralogy, Petrology and Economic Geology, Faculty of Natural Sciences, Comenius University, Ilkovičova 6, 842 15 Bratislava, Slovakia
2
Institute of Nuclear and Physical Engineering, Faculty of Electrical Engineering and Information Technology, Slovak University of Technology, Ilkovičova 3, 812 19 Bratislava, Slovakia
3
Department of Nuclear Reactors, Faculty of Nuclear Sciences and Physical Engineering, Czech Technical University in Prague, V Holešovičkách 2, 180 00 Prague, Czech Republic
4
Center STU for Nano-Diagnostics, Slovak University of Technology, Vazovova 5, 812 43 Bratislava, Slovakia
5
Department of Geological Sciences, Masaryk University, Kotlářská 2, 611 37 Brno, Czech Republic
*
Author to whom correspondence should be addressed.
Minerals 2022, 12(2), 238; https://doi.org/10.3390/min12020238
Submission received: 22 December 2021 / Revised: 4 February 2022 / Accepted: 8 February 2022 / Published: 12 February 2022
(This article belongs to the Special Issue Mineralogy, Petrology and Crystallography of Silicate Minerals)

Abstract

:
A heat treatment was performed on selected epidote and clinozoisite crystals to establish the nature of any changes in the optical and crystal-chemical properties and to identify a breakdown product using a wide spectrum of analytical methods. Natural samples were heated from 900 to 1200 °C under atmospheric pressure in ambient oxidation conditions for 12 h. Epidote and clinozoisite were stable at 900 °C; those heated at 1000 °C, 1100 °C, and 1200 °C exhibited signs of breakdown, with the development of cracks and fissures. The average chemical composition of epidote is Ca2.000Al2.211Fe0.742Si2.994O12(OH), while that of clinozoisite is Ca2.017A12.626Fe0.319Si3.002O12(OH). The breakdown products identified by electron microanalysis, powder X-ray diffraction, Raman spectroscopy, and high-resolution transmission electron microscopy were anorthite, pyroxene compositionally close to esseneite, and wollastonite. The decomposition of the epidote-clinozoisite solid solution is controlled by the following reaction: 4 epidote/clinozoisite → 2 pyroxene + 2 wollastonite + 4 anorthite + 2 H2O. Pyroxene likely contains a significant proportion of tetrahedral Fe3+ as documented by the Mössbauer spectroscopy. Moreover, the presence of hematite in the Mössbauer spectrum of the clinozoisite sample heated at 1200 °C can result from the following reaction: 4 epidote → pyroxene + 3 wollastonite + 4 anorthite + hematite + 2 H2O.

1. Introduction

Epidote and clinozoisite are monoclinic sorosilicate minerals belonging to the epidote supergroup; their formulae can be written as A1CaA2CaM1AlM2AlM3Fe3+(Si2O7)(SiO4)O(OH) and A1CaA2CaM1AlM2AlM3Al(Si2O7)(SiO4)O(OH), respectively [1]. The A(1) and A(2) sites with a coordination number (CN) >6 are mainly occupied by Ca, while M(1), M(2), and M(3) sites with CN = 6 primarily contain Al and Fe3+ [2]. The crystal structure of epidote is complex and consists of single silicate tetrahedra SiO4 and double silicate tetrahedra Si2O7, as well as continuous chains of AlO6 and AlO4(OH)2 octahedra running parallel to the b axis [3]. Several studies showed that Fe3+ preferentially occupies the M3 site [2,3,4], and that the amount of Fe3+ at the M1 and M2 sites is not very extensive [5].
The crystal structure and crystal chemistry of epidote-clinozoisite series at room or low temperature have been investigated extensively using several analytical methods and theoretical models [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23]. However, not many studies have been devoted to high-temperature (HT) epidote-clinozoisite behavior during in situ HT experiments, despite the importance of these minerals in the field of metamorphic and/or magmatic petrology [4]. A single-crystal diffraction study in situ HT was only performed on Sr-piemontite [24]. This X-ray diffraction study was not sufficient to monitor the T-induced disordering phenomenon in epidote [24]. A unit-cell thermal expansion of clinozoisite at high temperature [25] and a unit-cell compression of epidote at high pressure [26] were described. A low-T experiment was performed on epidote with a composition of (Cal.98Mn0.02)(Fe0.81A12.19)Si3O12(OH) with a neutron structure refinement at 15 K [12]. In addition, an examination of OH stretching modes and their pleochroic scheme was performed using single-crystal polarized infrared (FTIR) spectroscopy at room T [2].
Epidote supergroup minerals have a wide range of stability. Epidotes are, however, stable over a wide range of pressure and temperature conditions in continental and oceanic crust [27,28]. They were found in shallow geothermal fields at a few hundred meters of depth, suggesting a pressure of much less than 0.1 GPa and a temperature below 200 °C [29]. Commonly, epidote breaks down in the range of 600 to 700 °C, but incorporation of REE can stabilize it [3]. Allanite may occur in a variety of igneous rocks with high crystallization temperature and has been reported at 1050 °C, 3 to 4.5 GPa [30]. An increasing number of experimental studies have revealed that the epidote-supergroup minerals have huge stability fields which extend to 7 GPa pressure and to more than 1200 °C temperature [27]. The stability of epidote is controlled by pressure, temperature, the Al/Fe3+ ratio, oxygen fugacity, and fluid composition [7,23,31].
A few experimental studies on epidote stability have been published. The effects of temperature on the crystal structure of epidote at 293 and 1070 K (~20 °C, ~797 °C) [4], the behavior of epidote at high pressure (10 GPa) and high temperature (1200 K; ~927 °C) [28], and the impact of low temperature on epidote structure [2] were studied. Powder X-ray diffraction data at 1070 K showed epidote stability within the temperature range, and crystallinity was maintained [4]. The thermal stability of epidote up to 1070 K was observed in this study; the absence of dehydration and/or nonconvergent ordering of Al and Fe3+ between different octahedral sites and/or convergent ordering on M3 were likely due to the slow kinetics of the aforementioned processes [4]. The instability of epidote was defined at T > 1200 K with breakdown products including anorthite and hedenbergitic pyroxene with no corundum or hematite found; the breakdown reaction was as follows: epidote → anorthite + pyroxene ± Al2O3 ± Fe2O3 + H2O [2]. The pressure plays an important role in epidote stability; the breakdown is more rapid with increasing pressure.
This experimental study was focused on the breakdown reactions and products of epidote-clinozoisite series after heating at temperatures on the upper edge of their stability at various pressures under variable experimental conditions. The natural samples were selected, in contrast to previous works mostly done on synthetic samples, to describe the possible effects of chemical inhomogeneity and diffusion on the mineral breakdown. Firstly, the influence of variable Fe/Al ratio on low-P experiment products was determined. After the experiments, several analytical methods were applied to study structural changes and phase transitions including the nanoscale identification of breakdown products by high-resolution transmission electron microscopy (HRTEM).

2. Materials and Methods

Optically and chemically very homogeneous single crystals were selected for experiments. Crystals of natural dark-green epidote were collected from Pfarrerb near the Sobotín metabasite massif in the southern part of Hrubý Jeseník Mountains, Czech Republic (ES). Natural brownish green clinozoisite crystals were from Alchuri, Shigar Valley, Skardu district, Baluchistan, Pakistan (CP).
Samples were divided into several groups required for different heating procedures. One sample from each group was not heat-treated and was used as a standard (marked as 0), while other samples were treated (labeled according to temperature). Both sets of samples were heated in a muffle furnace in an air atmosphere for 12 h. Samples were subsequently cooled to ambient temperature for 12 h. A list of samples is provided in Table 1.
Chemical analysis was performed with the Cameca SX100 electron microprobe (Cameca company, part of Ametek group, Gennevilliers Cedex, France) operated in wavelength-dispersion mode at the Masaryk University, Brno, Czech Republic, under the following conditions: accelerating voltage 15 kV, beam current 20 nA, and beam diameter 5 μm. The samples were analyzed with the following standards: wollastonite (SiKα. CaKα), TiO2 (TiKα), Al2O3 (AlKα), pure Cr (CrKα), fayalite (FeKα), rhodonite (MnKα), MgO (Mg Kα), pure Ni (NiKα), albite (NaKα), orthoclase (KKα), Rb2ZnSi5O12 glass (RbLα), pollucite (CsLα), barite (BaLα), SrTiO3 (SrLα), BaF2 (FKα), and NaCl (ClKα). Detection limits of the measured elements varied between 0.01 and 0.05 wt.% with Ba, F, and Cl always below the detection limit. Peak counting times were 10 to 40 s during measurements depending on the expected concentration of the element in the mineral phase. Major elements were measured using shorter times, whereas longer times were applied for elements with a low concentration. The chemical formula of epidote-supergroup minerals was calculated on the basis of ∑ (A + M + T) = 8 cations with the procedure following epidote-supergroup nomenclature [1].
Powder X-ray diffraction analyses were determined using a BRUKER D8 Advance diffractometer (Bruker AXS GmbH, Karlsruhe, Germany) at the Laboratory of X-ray diffraction SOLIPHA (Comenius University in Bratislava, Faculty of Natural Sciences, Bratislava, Slovakia) under the following conditions: Bragg-Brentano geometry (Theta-2Theta), Cu anticathode (Kα1 = 1.5406 Å), accelerating voltage 40 kV, and beam current 40 mA. Ni Kβ filters were used for stripping Kβ radiation, and data were obtained using the BRUKER LynxEye detector. The step size was 0.01° 2θ, the counting time was 3 s per step, and measurements ranged from 2° to 65° 2θ. Analyzed scans were fitted, and lattice parameters were refined with BRUKER DIFFRACplus TOPAS software using the structural models for appropriate mineral phases—anorthite [32], pyroxene esseneite [33], and wollastonite [34].
The 57Fe Mössbauer spectra were obtained at room temperature (298 K) using a standard constant acceleration spectrometer with a 57Co/Rh radioactive source. Calibration of the apparatus was performed by a 12.5 μm foil of metallic α-Fe and a commercial (Amersham) calibration pellet of sodium nitroprusside (surface density ~5 mg Fe/cm2) for high- (±10.5 mm/s) and low-velocity (±4.5 mm/s) regions, respectively. However, all isomer shifts herein are quoted with respect to the room-temperature Mössbauer spectrum of α-Fe. Errors in determination of particular spectral components including relative area, isomer shift (IS), quadrupole splitting (QS) (and/or quadrupole shift), line width (Γ), and hyperfine magnetic field (Bhf) were estimated to be ±1%, ±0.02 mm/s, ±0.03 mm/s, ±0.03 mm/s, and ±0.3 T, respectively (Institute of Nuclear and Physical Engineering, Faculty of Electrical Engineering and Information Technology, The Slovak University of Technology, Bratislava).
Raman spectra of low-P-treated epidote and clinozoisite were acquired on a LabRAM HR Evolution (Horiba, Jobin Yvon, Palaiseau, France) Raman spectrometer system with a Peltier-cooled CCD detector and Olympus BX-41 microscope (Masaryk University, Department of Geological Sciences, Brno, Czech Republic). The Raman spectra were excited by a 532 nm (diode) laser and collected in the range between 100 cm−1 and 4000 cm−1 with a resolution of 2 cm−1. Multiple spot analyses on different areas of epidote produced similar spectra and confirmed the spectral reproducibility. No surface damage was observed after the laser illumination of the measurement. The acquired Raman spectra were processed using the Peakfit (Systat software, Inc., San Jose, CA, USA) software package. Band fitting was done using the Lorentz function with variable width; the fitting was gradually refined until it produced reproducible results with a square regression coefficient greater than 0.995.
High-resolution transmission electron microscopy (HRTEM) characterization was performed on a Jeol JEM ARM 200cF (JEOL Ltd., Akishima, Tokyo, Japan) analytical transmission electron microscope (TEM) operated at 200 kV that was equipped with a JEOL JED 2300 SDD detector for X-ray energy-dispersive spectroscopy (EDS) microanalysis (Center STU for Nanodiagnostic, The Slovak University of Technology, Bratislava, Slovakia). EDS microanalysis was used to study breakdown products and to distinguish between individual phases. Samples for TEM examination were prepared using a standard technique, which involved mechanical grinding of parent material followed by making a suspension of produced powders and ethanol, before undergoing sonication for 10 min. A drop of suspension was then deposited on a carbon-coated TEM Cu-grid. After drying in the air, the samples were examined by TEM.

3. Results

Both ES and CP samples displayed only weak chemical zoning in backscattered electron (BSE) images with small variations in Al and Fe3+ content (Figure 1 and Figure 2) and the average chemical composition listed in Table 2. The main difference between samples was in their Fe/Al ratio. Epidote from Sobotín was M3Fe3+-dominant (0.68–0.80 apfu), while clinozoisite samples had Fe3+ in the range 0.29–0.33 apfu (Table 2, Figure 3). The content of other cations including Ca, Mg, Ti, V, and Cr was very low or below the detection limit in both types of samples.
After heating at 900 °C, neither sample showed signs of phase transitions, whereas, in samples heated at more than 1000 °C, a new mineral association with plagioclase (anorthite), pyroxene (with a composition close to esseneite), and wollastonite evolved. The breakdown products formed a homogeneous mass retaining the original composition of epidote and clinozoisite in samples heated up to 1100 °C, but samples heated at 1200 °C contained domains enriched or depleted in Fe3+ (Figure 1 and Figure 2; Table 3 and Table 4).
This suggests that breakdown products were intergrown on the nanoscale. Moreover, in both sets of samples heated from 1000 °C to 1200 °C, a system of crystallographically oriented cracks and fissures developed (Figure 1c,d and Figure 2c,d).
The formation of cracks was connected with volume changes after the transformation and dehydration with a visible increase in size with rising temperature. This phenomenon was probably related to structural changes in the transition of ES and CP to the breakdown products. The breakdown process is well documented in the XRD pattern (Figure 4 and Figure 5) and Raman spectra (Figure 6 and Figure 7). The unit-cell dimensions of breakdown products are provided in Table 5 and Table 6. The Raman spectra of samples heated at 900 °C clearly show that at this temperature did not take place (Figure 6, Table 7); however, spectra measured from the matrix developed after heating at 1000–1200 °C could be interpreted as a composite of anorthite, pyroxene and wollastonite (Figure 7, Table 8), confirming the results from powder XRD.
Both sets of samples heated at 1000 °C or more displayed significant changes in the Mössbauer spectra compared to original samples and samples heated at 900 °C (Table 9 and Table 10, Figure 8). According to the mineral phases in breakdown products, it is most likely that Fe was incorporated into the pyroxene structure at both octahedral and tetrahedral sites according to isomer shift (IS) and quadrupole splitting (QS) of measured doublets. Doublets with QS above 0.97 mm/s were attributed to Fe3+ at the T site, whereas those with QS below 0.70 mm/s were interpreted as Fe3+ at the M1 site. Moreover, in the CP sample heated at 1200 °C, a magnetic component occurred, which could be interpreted as hematite.
Epidote-clinozoisite structural breakdown and breakdown products were documented on the nanoscale by HRTEM. Breakdown products formed particles several tens of nm (Figure 9a). It was also used to confirm the presence of both pyroxene and wollastonite, because these XRD peaks and Raman bands were partly overlapping (not in the case of well-evidenced anorthite); therefore, this observation could be taken as their definitive proof. The particle with O, Si, and Ca identified in the EDS spectrum was wollastonite (Figure 9b), whereas the nanocrystal with O, Si, Ca, Fe, Al, and Ti peaks was pyroxene (Figure 9c).

4. Discussion

In the present experiment, the epidote-clinozoisite solid solution stability was studied under ambient conditions, atmospheric pressure, and without addition of any fluid phase. There was no evidence of phase transition or breakdown of epidote or clinozoisite at 900 °C; however, at T > 1000 °C, a breakdown of both epidote and clinozoisite was observed. The Raman spectra of samples heated at temperatures greater than 1000 °C contained relatively broad bands and were probably the composite of both anorthite and esseneite-like pyroxene contributions. The broadness of the bands suggested a very low structural order of the breakdown product, which was supported by other analytical data.
All products were convincingly identified by the selected analytical methods. Anorthite and wollastonite presumably had low chemical variability. The original epidote and clinozoisite had low Na content, below the EMPA detection limit. This limited possible albite substitution.
Pyroxene was the most intriguing breakdown product. It had the largest compositional variability and could accommodate Si, Ca, and both Al and Fe. This composition indicates that pyroxene in our samples had QUAD composition [47]. The composition of hedenbergite proposed for epidote breakdown [28] was not very likely for our reaction because hedenbergite contains only divalent Fe, and, without Na, there is no substitution that could balance the increased charge of Fe3+ indicated by the Mössbauer study of our samples. However, a tschermakite-type substitution with trivalent cation substituting for Si4+ could charge-balance the exclusive Fe3+ content in studied pyroxene. The tschermakite-type substitution is the most probable mechanism allowing the incorporation of trivalent cations such as Fe3+ into Ca pyroxene, similarly to esseneite [33].
The doublets in the Mössbauer spectra of breakdown products could be divided into two groups. Both had an isomer shift below 0.40 mm/s, indicating that all Fe in pyroxene was ferric [37,48,49,50]. However, there were differences between both groups according to their quadrupole splitting. Doublets with quadrupole splitting between 0.60 and 0.70 mm/s were similar to those of octahedral Fe3+ at the M1 site of aegirinic pyroxene [37,48,50], has albeit with slightly higher QS than in pyroxenes from a hedenbergite-aegirine solid solution (up to 0.49 mm/s [37]). This could have result from the different distortion level of M1 octahedron in the studied pyroxene.
The second group of doublets had distinctly higher quadrupole splitting than octahedral Fe3+ doublets. Their QS and IS values were similar to tetrahedral Fe3+ in Fe-esseneitic or Ca-Fe-tschermakitic pyroxenes [49,51,52]. The presence of Fe3+ at the tetrahedral site caused an expansion of the T tetrahedron. This expansion was anisotropic; the length of nonbridging T-O1 and T-O2 bonds was very similar in pyroxene with tetrahedral Al and Fe3+, while the bridging bonds were larger by 0.02 Å in pyroxene with tetrahedral Fe3+ than those in CaTs [49].
The quadrupole splitting had a negative correlation to the area of tetrahedral Fe3+ doublets but was positively correlated for octahedral Fe3+ (Figure 10a,b). The decrease in tetrahedral Fe3+ quadrupole splitting could be attributed to the reduced distortion of the tetrahedra with increasing TFe3+ or TAl occupancy due to M1Fe2+ + TSi ↔ M1Fe3+ + TFe3+ substitution [49,53]. The positive correlation of octahedral Fe3+ in studied samples was unusual. The difference to the published data [49] could be explained by the different substitution. The electric field was more strongly perturbed by changing the valence of neighboring M1 or M2 ions, and the change in quadrupole splitting of M1Fe3+ was also influenced by a change in valence of the tetrahedral site ions [48]. In the synthetic Si-deficient pyroxenes of CaFe3+AlSiO6CaTiAl2O6 composition, the octahedral distortion and change in quadrupole splitting of M1Fe3+ were influenced by the M1Fe3+M1Ti4+ and TSi ↔ TAl substitutions [49]. However, in the studied samples, the Ti content was low. Therefore, the accommodation of Fe3+ at the octahedral site was allowed by M1R2+ + TSi ↔ M1R3+ + TR3+, where R is either Fe or Al. Consequently, if a random distribution of Al and Fe3+ is assumed, an increase in the octahedral distortion and quadrupole splitting would have been in a linear dependence of the M1Fe3+ proportion. The evolution of quadrupole splitting values against the heating temperature is shown in Figure 10c,d.
Pyroxenes in the studied samples showed some similarities, as well as some differences. The a size was similar to hedenbergite [37] in the ES samples, whereas, in the CP samples heated at >1100 °C, it was significantly smaller, even smaller than that of aegirine [37], shifting toward jadeite [38]. The b size of ES pyroxene was very similar to the published esseneite [35] and aegirine [37] data. The data for CP pyroxene were very variable, increasing with the temperature from a value similar to jadeite [38] for the sample heated at 1000 °C to a value approaching that of hedenbergite [37]. The c size in the CP samples was very similar to aegirine [37] or esseneite [35], whereas, in ES pyroxene, it was even larger than esseneite.
On the basis of the lattice parameters of studied samples, it can be concluded that a and b indicate variable Al3+ and Fe3+ occupancy at the M sites. Observed variations could result from the various degree of ordering of these cations. The c sizes in all samples clearly evidence a relatively high substitution of trivalent cations for Si4+ at the T site. The CP pyroxene likely had a composition close to esseneite. Esseneite was described with the composition of (Ca1.01Na0.01) (Fe3+0.72Mg0.16Al0.04Ti0.03Fe2+0.02) Si1.19Al0.81O6.00 [33] and end-member formula CaFe3+AlSiO6. However, the ES pyroxene with the c size larger than esseneite very likely had Fe3+ at the T site, expanding the tetrahedra and, as the result of its larger ionic radius (0.49 Å, [54]) than Si (0.26 Å, [54]) and Al (0.39 Å, [54]), extending the tetrahedral chain. There was no detectable Fe3+ in the tetrahedra of natural esseneite, but experimentally synthesized CaFe3+AlSiO6 contained tetrahedral Fe3+, implying a cooling rate dependence of Fe3+ and Al ordering at the M1 and T sites [33]. Consequently, the proportion of Fe3+ and Al in pyroxene varied to replicate the composition of original epidote and clinozoisite. On the basis of the Fe3+ content in the original samples and with regard to proposed reactions, there was ca. 0.60 apfu and 0.20 apfu of tetrahedral Fe3+ in ES and CP samples, respectively.
Taking into account the observed associations of breakdown products, we can propose variable temperature-induced transformation reactions of the epidote-clinozoisite solid solution:
4 clinozoisite → 2 pyroxene(1) + 2 wollastonite + 4 anorthite + 2 H2O,
4 Ca2Al3Si3O12(OH) → 2 CaAl(AlSi)O8 + 2 CaSiO3 + 4 CaAl2Si2O8 + 2 H2O,
for clinozoisite, and
4 epidote → 2 pyroxene(2) + 2 wollastonite + 4 anorthite + 2 H2O,
4 Ca2Al2Fe3+Si3O12(OH) → 2 CaFe3+(SiFe3+)O6 + 2 CaSiO3 + 4 CaAl2Si2O8 + 2 H2O,
for epidote.
It is important to stress that these reactions were calculated for ideal end-member compositions; the composition of products was also ideal. However, the variations in Al:Fe ratio could be balanced not only by the different proportion of breakdown products in each reaction but also by the change in the composition of breakdown products.
Moreover, hematite was observed in the Mössbauer spectrum of the CP sample heated at 1200 °C, although it was not clearly identified in powder XRD. This can be explained by the large peak broadening induced by the small size of particles observed in the studied sample. In contrast, in the Mössbauer spectroscopy, the proportion of phases was determined by the proportion of Fe contained in them. Consequently, hematite was more significantly reflected in the Mössbauer spectrum than in the powder XRD pattern when associated with a large volume of Fe-poorer (pyroxene) or even almost Fe-free (anorthite, wollastonite) minerals.
The breakdown of the epidote component in the CP sample at 1200 °C with the production of hematite can be driven by following reaction:
4 epidote → 1 pyroxene(2) + 3 wollastonite + 4 anorthite + 1 hematite + 2 H2O,
4 Ca2Al2Fe3+Si3O12(OH) → 1 CaFe3+(SiFe3+)O6 + 3 CaSiO3 + 4 CaAl2Si2O8 + Fe2O3 + 2 H2O.
This reaction is similar to the previous one; the production of hematite is managed by the partial transformation of Fe-bearing clinozoisite through pyroxene(2) resulting in hematite and wollastonite formation at 1200 °C. Hematite is the final product of high-temperature heating of all Fe-oxidic compounds under oxidizing conditions [55]. This also includes Fe-bearing silicates such as tourmalines, which, when heated at a temperature of 900 °C, break down producing a mass containing mullite and hematite [56]. However, at temperatures below 1200 °C, pyroxene(2) seems to be stable, and no hematite is produced.
Breakdown of epidote probably had a subsolidus character with recrystallization of primary mineral with deformation mechanism induced by crystal-structural changes resulting in grain-size reduction. The texture of samples was maintained and contained decomposed minerals. Samples were without signs of any differentiation or partial melting. The temperature of 1200 °C allowed partial chemical diffusion as a result of the reaction of product division. This could also indicate a subsolidus character of the reaction.
The resulting breakdown association with anorthite, wollastonite, and esseneitic pyroxene also has implications for natural processes. Esseneite is a product of either high-temperature recrystallization of silicate-bearing rocks in strongly oxidizing conditions usually induced by coal burning [33,57] or high-temperature metomorphic and metasomatic processes connected with volcanic extrusions and magmatic intrusions in form of skarns and skarnoids [35,58,59]. Wollastonite and anorthite occur in very high-temperature parageneses, such as at Nanga Parbat, India, where this association replaces grossular and quartz [60]. Associations with clinopyroxene (En15-43Fs14-36Wo41-51) + plagioclase (An37-100) ± magnetite and with wollastonite ± clinopyroxene (En17-38Fs8-34Wo49-59) ± plagioclase (An46-100) ± garnet (Grs0-65Adr24-75Sch0-76) ± quartz were observed in calc-silicate xenoliths from Merapi Volcano, Indonesia upon contact of lava with limestones [61]. Therefore, it is possible to assume that epidote-group minerals can also be precursors for tschermakitic pyroxenes and pyroxenoids, if the temperature exceeds their stability.

5. Conclusions

In the present experiment, the epidote-clinozoisite solid solution stability was studied under ambient conditions, atmospheric pressure and without addition of any fluid. The structural breakdown was observed in samples heated at 1000–1200 °C. The breakdown mineral association comprised anorthite, pyroxene, and wollastonite with additional hematite occurring after heating at 1200 °C. The XRD and Mössbauer spectroscopy data suggest that pyroxene contained a significant proportion of Ca-tschermakitic molecules, resulting in a composition similar to esseneite but with the substitution of Al3+ and Fe3+ at both tetrahedral and octahedral sites depending on their proportion in the original epidote and clinozoisite sample. The high-temperature and low-pressure breakdown of epidote to the specific mineral association suggests that a similar process could also take place naturally.

Author Contributions

Conceptualization, P.K. and P.B.; methodology, P.K., P.B., M.M., M.Č. and R.Š.; data curation, P.K. and P.B.; writing—original draft preparation, P.K.; writing—review and editing, P.B. and M.M.; supervision, P.B.; project administration, P.B. and M.M.; funding acquisition, P.B. and M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Education of Slovak Republic grant agency under the contracts VEGA 1/0137/20 and VEGA 1/0130/20, by the European Regional Development Fund-Project “Center for Advanced Applied Sciences” (grant number CZ.02.1.01/0.0/0.0/16_019/0000778), and on the basis of the major project “Advancing University Capacity and Competence in Research, Development and Innovation” (ITMS project code: 313021X329) supported by Operational Program Integrated Infrastructure and funded by the European Regional Development Fund.

Data Availability Statement

Data supporting the reported results can be obtained from the authors upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Armbruster, T.; Bonazzi, P.; Akasaka, M.; Bermanec, V.; Chopin, C.; Gieré, R.; Heuss-Assbichler, S.; Liebscher, A.; Menchetti, S.; Pan, Y.; et al. Recommended nomenclature of epidote-group minerals. Eur. J. Mineral. 2006, 18, 551–567. [Google Scholar] [CrossRef] [Green Version]
  2. Gatta, G.D.; Alvaro, M.; Bromiley, G. A low temperature X-ray single-crystal diffraction and polarised infra-red study of epidote. Phys. Chem. Miner. 2012, 39, 1–15. [Google Scholar] [CrossRef] [Green Version]
  3. Franz, G.; Liebscher, A. Physical and Chemical Properties of the Epidote Minerals: An Introduction. Rev. Mineral. Geochem. 2004, 56, 1–81. [Google Scholar] [CrossRef]
  4. Gatta, G.D.; Meven, M.; Bromiley, G. Effects of temperature on the crystal structure of epidote: A neutron single-crystal diffraction study at 293 and 1,070 K. Phys. Chem. Miner. 2010, 37, 475–485. [Google Scholar] [CrossRef] [Green Version]
  5. Gabe, E.J.; Portheine, J.A.N.C.; Whitlone, S.H. A Reinvestigation of the Epidote Structure: Confirmation of the Iron Location. Am. Mineral. 1973, 58, 218–223. [Google Scholar]
  6. Langer, K.; Raith, M. Infrared spectra of Al-Fe(III)-epidotes and zoisites, Ca2(Al1-pFe3+p)Al2O(OH)[Si2O7][SiO4]. Am. Mineral. 1974, 59, 1249–1258. [Google Scholar]
  7. Bird, D.K.; Helgeson, H.C. Chemical interaction of aqueous solutions with epidote-feldspar mineral assemblages in geologic systems; 1, Thermodynamic analysis of phase relations in the system CaO-FeO-Fe2O3-Al2O3-SiO2-H2O-CO2. Am. J. Sci. 1980, 280, 907–941. [Google Scholar] [CrossRef]
  8. Carbonin, S.; Molin, G. Crystal-chemical considerations of eight metamorphic epidotes. Neues Jahrb. Mineral. Abh. 1980, 139, 205–215. [Google Scholar]
  9. Apted, M.J.; Liou, J.G. Phase relations among greenschist, epidote-amphibolite, and amphibolite in a basaltic system. Am. J. Sci. 1983, 283 A, 328–354. [Google Scholar]
  10. Stergiou, A.C.; Rentzeperis, P.J.; Sklavounos, S. Refinement of the crystal structure of a medium iron epidote. Z. Krist.-Cryst. Mater. 2014, 178, 297–305. [Google Scholar] [CrossRef]
  11. Bird, D.K.; Cho, M.; Janik, C.J.; Liou, J.G.; Caruso, L.J. Compositional, order/disorder, and stable isotope characteristics of Al-Fe epidote, State 2-14 drill hole, Salton Sea geothermal system. J. Geophys. Res. 1988, 93, 13135–13144. [Google Scholar] [CrossRef]
  12. Kvick, Å.; Pluth, J.J.; Richardson, J.W.; Smith, J.V. The ferric ion distribution and hydrogen bonding in epidote: A neutron diffraction study at 15 K. Acta Crystallogr. Sect. B 1988, 44. [Google Scholar] [CrossRef]
  13. Bonazzi, P.; Menchetti, S. Monoclinic members of the epidote group: Effects of the AI ⇌ Fe3+ ⇌ Fe2+ substitution and of the entry of REE3+. Mineral. Petrol. 1995, 53, 133–153. [Google Scholar] [CrossRef]
  14. Comodi, P.; Zanazzi, P.F. The pressure behavior of clinozoisite and zoisite: An X-ray diffraction study. Am. Mineral. 1997, 82, 61–68. [Google Scholar] [CrossRef]
  15. Giuli, G.; Bonazzi, P.; Menchetti, S. Al-Fe disorder in synthetic epidotes: A single-crystal X-ray diffraction study. Am. Mineral. 1999, 84, 933–936. [Google Scholar] [CrossRef]
  16. Bonazzi, P.; Menchetti, S. Manganese in monoclinic members of the epidote group: Piemontite and related minerals. Rev. Mineral. Geochem. 2004, 56, 495–552. [Google Scholar] [CrossRef]
  17. Liebscher, A. Spectroscopy of epidote minerals. Rev. Mineral. Geochem. 2004, 56, 125–170. [Google Scholar] [CrossRef]
  18. Dollase, W.A. Refinement and comparison of the structures of zoisite and clinozoisite. Am. Mineral. 1968, 53, 1882–1898. [Google Scholar]
  19. Dollase, W.A. Refinement of the Crystal Structures of Epidote, Allanite and Hancockite. Am. Mineral. 1971, 56, 447–464. [Google Scholar]
  20. Dollase, W.A. Crystal structure and cation ordering of piemontite. Am. Mineral. 1969, 54, 710–717. [Google Scholar]
  21. Dollase, W.A. Mössbauer spectra and iron distribution in the epidote-group minerals*. Z. Krist. 1973, 138, 41–63. [Google Scholar] [CrossRef]
  22. Ito, T.; Morimoto, N.; Sadanaga, R. On the structure of epidote. Acta Crystallogr. 1954, 7, 53–59. [Google Scholar] [CrossRef]
  23. Liou, J.G. Synthesis and stability relations of epidote, Ca2Al2FeSi3O12(OH). J. Petrol. 1973, 14. [Google Scholar] [CrossRef]
  24. Catti, M.; Ferraris, G.; Ivaldi, G. Thermal behavior of the crystal structure of strontian piemontite. Am. Mineral. 1988, 73, 1370–1376. [Google Scholar]
  25. Pawley, A.R.; Redfern, S.A.T.; Holland, T.J.B. Volume behavior of hydrous minerals at high pressure and temperature: I. Thermal expansion of lawsonite, zoisite, clinozoisite, and diaspore. Am. Mineral. 1996, 81, 335–340. [Google Scholar] [CrossRef]
  26. Holland, T.J.B.; Redfern, S.A.T.; Pawley, A.R. Volume behavior of hydrous minerals at high pressure and temperature: II. Compressibilities of lawsonite, zoisite, clinozoisite, and epidote. Am. Mineral. 1996, 81, 341–348. [Google Scholar] [CrossRef]
  27. Poli, S.; Schmidt, M.W. Experimental subsolidus studies on epidote minerals. Rev. Mineral. Geochem. 2004, 56, 171–195. [Google Scholar] [CrossRef]
  28. Gatta, G.D.; Merlini, M.; Lee, Y.; Poli, S. Behavior of epidote at high pressure and high temperature: A powder diffraction study up to 10 GPa and 1,200 K. Phys. Chem. Miner. 2011, 38, 419–428. [Google Scholar] [CrossRef]
  29. Bird, D.K.; Spieler, A.R. Epidote in geothermal systems. Rev. Mineral. Geochem. 2004, 56, 235–300. [Google Scholar] [CrossRef]
  30. Hermann, J. Allanite: Thorium and light rare earth element carrier in subducted crust. Chem. Geol. 2002, 192, 289–306. [Google Scholar] [CrossRef]
  31. Holdaway, M.J. Thermal stability of Al-Fe epidote as a function of fO2 and Fe content. Contrib. Mineral. Petrol. 1972, 37, 307–340. [Google Scholar] [CrossRef]
  32. Wenk, H.R.; Joswig, W.; Tagai, T.; Korekawa, M.; Smith, B.K. The average structure of An62–66 labradorite. Am. Mineral. 1980, 65, 81–95. [Google Scholar]
  33. Cosca, M.A.; Peacor, D.R. Chemistry and structure of esseneite (CaFe3+AlSiO6), a new pyroxene produced by pyrometamorphism. Am. Mineral. 1987, 72, 148–156. [Google Scholar]
  34. Ohashi, Y. Polysynthetically-twinned structures of enstatite and wollastonite. Phys. Chem. Miner. 1984, 10, 217–229. [Google Scholar] [CrossRef]
  35. Yakubovich, O.V.; Zayakina, N.V.; Oleinikov, O.B.; Kostin, A.V. Esseneite from Xenoliths in Dacite Lavas: Crystal Structure and Genesis. Geol. Ore Depos. 2019, 61, 689–695. [Google Scholar] [CrossRef]
  36. Raudsepp, M.; Hawthorne, F.C.; Turnock, A.C. Crystal chemistry of synthetic pyroxenes on the join CaNiSi2O6-CaMgSi2O6 (diopside); a Rietveld structure refinement study. Am. Mineral. 1990, 75, 1274–1281. [Google Scholar]
  37. Redhammer, G.J.; Amthauer, G.; Roth, G.; Tippelt, G.; Lottermoser, W. Single-crystal X-ray diffraction and temperature dependent 57Fe Mössbauer spectroscopy on the hedenbergite-aegirine (Ca,Na)(Fe2+,Fe3+)Si2O6 solid solution. Am. Mineral. 2006, 91, 1271–1292. [Google Scholar] [CrossRef]
  38. Nestola, F.; Tribaudino, M.; Boffa Ballaran, T.; Liebske, C.; Bruno, M. The crystal structure of pyroxenes along the jadeite-hedenbergite and jadeite-aegirine joins. Am. Mineral. 2007, 92, 1492–1501. [Google Scholar] [CrossRef]
  39. Ohashi, Y.; Finger, L.W. The role of octahedral cations in pyroxenoid crystal chemistry. I. Bustamite,, wollastonite, and the pectolite-schizolite-serandite series, sample Mn-WO, from Broken Hill, New South Wales, Australia. Am. Mineral. 1978, 63, 274–288. [Google Scholar]
  40. Angel, R.J. High-pressure structure of anorthite. Am. Mineral. 1988, 73, 1114–1119. [Google Scholar]
  41. Nagashima, M.; Akasaka, M. X-ray Rietveld and 57Fe Mössbauer studies of epidote and piemontite on the join Ca2Al2Fe3+Si3 O12(OH)-Ca2Al2Mn3+Si3O12(OH) formed by hydrothermal synthesis. Am. Mineral. 2010, 95, 1237–1246. [Google Scholar] [CrossRef]
  42. Lafuente, B.; Downs, R.T.; Yang, H.; Stone, N. The power of databases: The RRUFF project. In Highlights in Mineralogical Crystallography; Armbruster, T., Danisi, R.M., Eds.; Walter De Gruyter GmbH: Berlin, Germany, 2015; pp. 1–29. [Google Scholar]
  43. Wang, A.; Yu, J.; Han, J.; Guo, L.; Zeng, P. Database of Standard Raman Spectra of Minerals and Related Inorganic Crystals. Appl. Spectrosc. 1994, 48, 959–968. [Google Scholar] [CrossRef]
  44. Makreski, P.; Jovanovski, G.; Kaitner, B.; Gajović, A.; Biljan, T. Minerals from Macedonia. XVIII. Vibrational spectra of some sorosilicates. Vib. Spectrosc. 2007, 44, 162–170. [Google Scholar] [CrossRef]
  45. Sharma, S.K.; Simons, B.; Yoder, H.S. Raman study of anorthite, calcium Tschermak’s pyroxene, and gehlenite in crystalline and glassy states. Am. Mineral. 1983, 68, 1113–1125. [Google Scholar]
  46. Buzatu, A.; Buzgar, N. The Raman study of single-chain silicates. An. Stiintifice Univ. AI Cuza Din Iasi. Sect. 2 Geol. 2010, 56, 107–125. [Google Scholar]
  47. Morimoto, N. Nomenclature of Pyroxenes. Mineral. Petrol. 1988, 39, 55–76. [Google Scholar] [CrossRef]
  48. Dollase, W.A.; Gustafson, W.I. 57Fe Mössbauer spectra analysis of sodic clinopyroxenes. Am. Mineral. 1982, 67, 311–327. [Google Scholar]
  49. Akasaka, M. 57Fe Mössbauer study of clinopyroxenes in the join CaFe3+AlSiO6-CaTiAl2O6. Phys. Chem. Miner. 1983, 9, 205–211. [Google Scholar] [CrossRef]
  50. De Grave, E.; Van Alboom, A.; Eeckhout, S.G. Electronic and magnetic properties of a natural aegirine as observed from its Mössbauer spectra. Phys. Chem. Miner. 1998, 25, 378–388. [Google Scholar] [CrossRef]
  51. Ghose, S.; Okamura, F.P.; Ohashi, H. The crystal structure of CaFe3+SiAlO6 and the crystal chemistry of Fe3+-Al3+ substitution in calcium Tschermak’s pyroxene. Contrib. Mineral. Petrol. 1986, 92, 530–535. [Google Scholar] [CrossRef]
  52. Skogby, H.; Rossman, G.R. OH- in pyroxene: An experimental study of incorporation mechanisms and stability. Am. Mineral. 1989, 74, 1059–1069. [Google Scholar]
  53. Hafner, S.S.; Huckenholz, H.G. Mössbauer Spectrum of Synthetic Ferri-diopside. Nat. Phys. Sci. 1971, 233, 9–11. [Google Scholar] [CrossRef]
  54. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  55. Cornell, R.M.; Schwertmann, U. The Iron Oxides: Structure, Properties, Reactions, Occurrences and Uses; WILEY-VCH GmbH&Co. KGaA: Weinheim, Germany, 2003; Volume 39. [Google Scholar]
  56. Bačík, P.; Ozdín, D.; Miglierini, M.; Kardošová, P.; Pentrák, M.; Haloda, J. Crystallochemical effects of heat treatment on Fe-dominant tourmalines from Dolní Bory (Czech Republic) and Vlachovo (Slovakia). Phys. Chem. Miner. 2011, 38, 599–611. [Google Scholar] [CrossRef]
  57. Foit, F.F.; Hooper, R.L.; Rosenberg, P.E. An unusual pyroxene, melilite, and iron oxide mineral assemblage in a coal-fire buchite from Buffalo, Wyoming. Am. Mineral. 1987, 72, 137–147. [Google Scholar]
  58. Pascal, M.L.; Katona, I.; Fonteilles, M.; Verkaeren, J. Relics of high-temperature clinopyroxene on the join Di-CaTs with up to 72 mol.% Ca(Al,Fe3+)AlSiO6 in the skarns of Ciclova and Magureaua Vatei, Carpathians, Romania. Can. Mineral. 2005, 43, 857–881. [Google Scholar] [CrossRef] [Green Version]
  59. Reato, L.; Huraiová, M.; Konečný, P.; Marko, F.; Hurai, V. Formation of Esseneite and Kushiroite in Tschermakite-Bearing Calc-Silicate Xenoliths Ejected in Alkali Basalt. Minerals. 2022, 12, 156. [Google Scholar] [CrossRef]
  60. Misch, P. Stable association wollastonite-anorthite, and other cale-silicate assemblages in amphibolite-facies crystalline schists of Nanga Parbat, Northwest Himalayas. Beiträge Mineral. Petrogr. 1964, 10, 315–356. [Google Scholar] [CrossRef]
  61. Whitley, S.; Halama, R.; Gertisser, R.; Preece, K.; Deegan, F.M.; Troll, V.R. Magmatic and metasomatic effects of magma–carbonate interaction recorded in calc-silicate xenoliths from Merapi volcano (Indonesia). J. Petrol. 2020, 61, 48. [Google Scholar] [CrossRef]
Figure 1. BSE images of heat-treated epidote: (a) ES-0; (b) ES-9; (c) ES-10; (d) ES-11. The untreated natural sample (a) and sample treated at 900 °C (b) displayed only the weak chemical zoning caused by small variations in Al and Fe3+ content; samples heated over 1000 °C formed breakdown products firstly as a homogeneous mass (c), while those heated at 1100 °C formed domains with a slightly different composition (d).
Figure 1. BSE images of heat-treated epidote: (a) ES-0; (b) ES-9; (c) ES-10; (d) ES-11. The untreated natural sample (a) and sample treated at 900 °C (b) displayed only the weak chemical zoning caused by small variations in Al and Fe3+ content; samples heated over 1000 °C formed breakdown products firstly as a homogeneous mass (c), while those heated at 1100 °C formed domains with a slightly different composition (d).
Minerals 12 00238 g001
Figure 2. BSE images of heat-treated clinozoisite: (a) CP-0; (b) CP-9; (c) CP-10; (d) CP-11. The untreated natural sample (a) and sample treated at 900 °C (b) displayed only the weak chemical zoning caused by small variations in Al and Fe3+ content; as in previous samples, heating over 1000 °C led to the formation of breakdown products firstly as a homogeneous mass (c), whereas, at 1100 °C, domains with slightly different composition were formed (d).
Figure 2. BSE images of heat-treated clinozoisite: (a) CP-0; (b) CP-9; (c) CP-10; (d) CP-11. The untreated natural sample (a) and sample treated at 900 °C (b) displayed only the weak chemical zoning caused by small variations in Al and Fe3+ content; as in previous samples, heating over 1000 °C led to the formation of breakdown products firstly as a homogeneous mass (c), whereas, at 1100 °C, domains with slightly different composition were formed (d).
Minerals 12 00238 g002
Figure 3. Diagram of Al-Fe3+ substitution in studied samples. Epidote was M3Fe3+-dominant in contrast to Al-dominant clinozoisite samples, as documented by the different positions of ES and CP analyses in the diagram.
Figure 3. Diagram of Al-Fe3+ substitution in studied samples. Epidote was M3Fe3+-dominant in contrast to Al-dominant clinozoisite samples, as documented by the different positions of ES and CP analyses in the diagram.
Minerals 12 00238 g003
Figure 4. Powder X-ray diffraction patterns of epidote experimental samples (ES) compared to patterns of wollastonite [39], anorthite [40], esseneite [33], and epidote [41] (modeled from structural data using pseudo-Voigt function).
Figure 4. Powder X-ray diffraction patterns of epidote experimental samples (ES) compared to patterns of wollastonite [39], anorthite [40], esseneite [33], and epidote [41] (modeled from structural data using pseudo-Voigt function).
Minerals 12 00238 g004
Figure 5. Powder X-ray diffraction patterns of clinozoisite experimental samples (CP) compared to patterns of wollastonite [39], anorthite [40], esseneite [33], and clinozoisite [14] (modeled from structural data using pseudo-Voigt function).
Figure 5. Powder X-ray diffraction patterns of clinozoisite experimental samples (CP) compared to patterns of wollastonite [39], anorthite [40], esseneite [33], and clinozoisite [14] (modeled from structural data using pseudo-Voigt function).
Minerals 12 00238 g005
Figure 6. The Raman spectra of epidote (a) and clinozoisite (b), natural samples and heated at 900 °C in the 1200–100 cm−1 region; our spectra were compared with the published spectrum of epidote [42].
Figure 6. The Raman spectra of epidote (a) and clinozoisite (b), natural samples and heated at 900 °C in the 1200–100 cm−1 region; our spectra were compared with the published spectrum of epidote [42].
Minerals 12 00238 g006
Figure 7. The Raman spectra of epidote (a) and clinozoisite (b) samples, heat-treated at temperatures higher than 1000 °C, compared with the published spectra of anorthite, esseneite, and wollastonite [42].
Figure 7. The Raman spectra of epidote (a) and clinozoisite (b) samples, heat-treated at temperatures higher than 1000 °C, compared with the published spectra of anorthite, esseneite, and wollastonite [42].
Minerals 12 00238 g007
Figure 8. Room-temperature Mössbauer spectra of the ES and CP samples in their original state (ES-0, CP-0) and after annealing from 900 to 1200 °C. Experimental points are plotted as solid symbols. Fitting curves for all data points are gray. Decomposed peaks are also colored with different shades of gray. Shaded sub-spectra correspond in color to Fe3+ sites from Table 7 and Table 8.
Figure 8. Room-temperature Mössbauer spectra of the ES and CP samples in their original state (ES-0, CP-0) and after annealing from 900 to 1200 °C. Experimental points are plotted as solid symbols. Fitting curves for all data points are gray. Decomposed peaks are also colored with different shades of gray. Shaded sub-spectra correspond in color to Fe3+ sites from Table 7 and Table 8.
Minerals 12 00238 g008
Figure 9. The HRTEM images of epidote breakdown products: (a) nanocrystals of breakdown products; (b) detailed view of wollastonite with its EDS spectrum; (c) detailed view of pyroxene with its EDS spectrum. Note that Cu in the EDS spectrum is from the Cu TEM grid.
Figure 9. The HRTEM images of epidote breakdown products: (a) nanocrystals of breakdown products; (b) detailed view of wollastonite with its EDS spectrum; (c) detailed view of pyroxene with its EDS spectrum. Note that Cu in the EDS spectrum is from the Cu TEM grid.
Minerals 12 00238 g009
Figure 10. Comparison of quadrupole splitting and area of doublets for (a) ES and (b) CP samples and quadrupole splitting and heating temperature for (c) ES and (d) CP samples.
Figure 10. Comparison of quadrupole splitting and area of doublets for (a) ES and (b) CP samples and quadrupole splitting and heating temperature for (c) ES and (d) CP samples.
Minerals 12 00238 g010
Table 1. List of the studied samples.
Table 1. List of the studied samples.
TemperatureEpidote, Czech Republic (Sobotín)Clinozoisite, Pakistan (Baluchistan)
25 °CES-0CP-0
900 °CES-9CP-9
1000 °CES-10CP-10
1100 °CES-11CP-11
1200 °CES-12CP-12
Table 2. Electron microprobe analyses (in wt.% of oxides and apfu of cations) of epidote (ES) and clinozoisite (CP).
Table 2. Electron microprobe analyses (in wt.% of oxides and apfu of cations) of epidote (ES) and clinozoisite (CP).
ESESESESCPCPCP
P2O50.000.000.000.040.070.040.02
SiO237.1038.2338.0837.9438.7538.9338.97
TiO20.010.070.060.160.140.160.08
Al2O321.9625.1425.5425.1129.2029.3729.41
V2O30.000.090.050.040.030.080.09
Cr2O30.000.020.000.010.050.080.07
Fe2O315.0111.4210.4010.675.435.645.17
MnO0.010.020.050.040.050.010.03
MgO0.020.000.010.020.100.080.07
NiO0.010.000.050.000.000.010.03
CaO24.1623.8623.9923.9724.2924.3624.52
Na2O0.020.000.010.000.050.000.02
K2O0.000.000.000.010.000.020.02
H2O *1.881.911.911.901.951.951.95
Total100.17100.74100.1499.86100.10100.71100.47
Si4+2.9612.9952.9902.9942.9892.9862.990
P5+0.0000.0000.0000.0020.0050.0020.001
Al3+0.0390.0050.0100.0030.0070.0120.009
T3.0003.0003.0003.0003.0003.0003.000
Al3+1.0001.0001.0001.0001.0001.0001.000
M21.0001.0001.0001.0001.0001.0001.000
Mg2+0.0020.0000.0010.0030.0120.0090.008
Ni2+0.0000.0000.0030.0000.0000.0000.002
Fe3+0.9010.6730.6140.6340.3150.3260.299
Mn3+0.0010.0020.0030.0020.0040.0010.002
V3+0.0000.0050.0030.0030.0020.0050.005
Cr3+0.0000.0010.0000.0010.0030.0050.004
Al3+0.0950.3190.3750.3580.6640.6540.681
M31.0001.0001.0001.0001.0001.0001.000
Al3+0.9320.9980.9780.9750.9840.9890.969
Ti4+0.0010.0040.0030.0090.0080.0090.004
M10.9321.0030.9810.9840.9920.9980.973
Ca2+0.9961.0000.9980.9990.9921.0000.997
Na+0.0040.0000.0020.0010.0080.0000.003
A11.0001.0001.0001.0001.0001.0001.000
Ca2+1.0701.0031.0201.0281.0161.0031.019
A21.0701.0031.0201.0291.0161.0051.021
OH1.0001.0001.0001.0001.0001.0001.000
* H2O calculated for 1 OH per formula unit.
Table 3. Electron microprobe analyses (in wt.% of oxides) of the epidote breakdown products (bulk from area with 5 μm diameter).
Table 3. Electron microprobe analyses (in wt.% of oxides) of the epidote breakdown products (bulk from area with 5 μm diameter).
ES-10ES-10ES-10ES-11ES-11ES-11ES-11ES-11ES-12ES-12ES-12ES-12
P2O50.010.040.000.190.010.000.000.000.010.060.050.01
SiO237.0438.7536.4537.2437.4937.4937.2538.6938.7038.1638.2137.96
TiO20.000.000.030.160.020.010.040.000.030.000.030.19
Al2O323.3521.7323.1523.4226.5824.1421.8127.3821.4124.2928.3427.89
V2O30.000.050.010.040.000.060.030.070.010.000.020.08
Cr2O30.010.000.020.000.010.020.030.000.010.010.000.00
Fe2O313.0814.8212.2412.0110.6312.4516.349.1216.3512.409.2512.99
MnO0.050.010.020.580.040.040.020.070.030.020.080.06
MgO0.020.000.010.040.030.030.010.030.000.020.000.00
NiO0.000.010.000.020.030.010.000.010.000.000.010.00
CaO23.9623.7723.4825.7723.0424.6824.1322.5524.2124.0222.4923.00
Na2O0.010.020.020.050.010.010.000.010.010.040.000.05
K2O0.000.000.000.000.000.010.010.000.020.000.000.01
Total97.5399.1995.4199.5197.8998.9399.6797.92100.7899.0398.46102.24
Table 4. Electron microprobe analyses (in wt.% of oxides) of the clinozoisite breakdown products (bulk from area with 5 μm diameter).
Table 4. Electron microprobe analyses (in wt.% of oxides) of the clinozoisite breakdown products (bulk from area with 5 μm diameter).
CP-10CP-10CP-10CP-10CP-10CP-10CP-11CP-11CP-11CP-11CP-12CP-12CP-12
P2O50.060.050.000.000.020.010.230.040.000.010.030.050.02
SiO241.8240.5837.5938.7835.4839.3533.2338.6741.8139.3539.0339.1743.16
TiO20.060.070.150.140.170.150.610.170.030.030.060.140.09
Al2O328.5327.7828.7428.9722.6526.1215.3930.9033.1229.1231.0628.0429.96
V2O30.090.000.030.000.060.120.290.010.000.070.040.150.09
Cr2O30.070.060.080.030.060.090.280.010.000.020.050.100.03
Fe2O30.003.854.805.765.553.9913.933.700.005.044.255.300.00
MnO0.000.040.090.020.130.020.210.050.000.020.060.120.06
MgO0.040.040.100.030.150.030.160.070.000.060.070.060.02
NiO0.000.000.030.010.000.010.000.000.000.010.000.020.02
CaO22.8723.3722.4225.4334.3629.6335.3523.0820.3423.4125.4224.8823.50
Na2O0.000.000.000.000.000.000.000.000.030.000.020.000.01
K2O0.000.000.000.010.010.010.000.000.000.010.000.010.01
Total93.5395.8494.0399.1798.6399.5399.6696.6995.3497.13100.0798.0396.96
Table 5. Unit-cell parameters of the main breakdown products after the heat treatment of epidote (ES) and clinozoisite (CP).
Table 5. Unit-cell parameters of the main breakdown products after the heat treatment of epidote (ES) and clinozoisite (CP).
AnorthiteES-10ES-11ES-12CP-10CP-11CP-12
a (Å)8.1832(4)8.1847(4)8.1872(4)8.1819(4)8.1809(4)8.1820(4)
b (Å)12.876(3)12.8824(4)12.88044)12.778(3)12.8789(4)12.8774(4)
c (Å)7.0869(4)7.0879 (4)7.0877(4)7.033(3)7.0857(4)7.0887(4)
α (°)93.223(3)93.136(3)93.153(3)93.04(2)93.094(3)93.124(3)
β (°)115.720(3)115.769(3)115.763(3)115.59(2)115.766(3)115.803(3)
γ (°)91.059(3)91.188(3)91.191(3)91.147(3)91.227(3)91.232(3)
V3)670.94(2)671.2(1)671.3(2)656.4(1)670.5(1)670.6(1)
WollastoniteES-10ES-11ES-12CP-10CP-11CP-12
a (Å)8.011(3)7.940 (3)7.932 (3)7.923(3)7.945(3)7.911(3)
b (Å)7.309(3)7.318(3)7.317(3)7.222(3)7.341(3)7.279(3)
c (Å)6.753(3)6.755(3)7.058(3)7.105(3)7.051(3)7.055(3)
α (°)90.99(2)91.98(2)90.23(2)91.21(2)89.58 (2)90.74(2)
β (°)95.61(2)95.11(2)95.17(2)94.92(2)95.47 (2)95.09(2)
γ (°)103.49(2)103.18(2)103.36(2)103.67(2)103.33 (2)102.62(2)
V3)382.353(3)380.05 (2)396.811 (3)393.233(3)398.274(3)394.7(1)
Table 6. Unit-cell parameters of pyroxene as a breakdown product after the heat treatment of epidote (ES) and clinozoisite (CP) compared to esseneite (Ess), diopside (Di), hedenbergite (Hd), aegirine (Aeg), and jadeite (Jd).
Table 6. Unit-cell parameters of pyroxene as a breakdown product after the heat treatment of epidote (ES) and clinozoisite (CP) compared to esseneite (Ess), diopside (Di), hedenbergite (Hd), aegirine (Aeg), and jadeite (Jd).
ES-10ES-11ES-12CP-10CP-11CP-12Ess Di Hd Aeg Jd
Ref.This Work[35][36][37][37][38]
a (Å)9.815(3)9.8329(4)9.809(3)9.808(3)9.606(3)9.689 (3)9.7610(12)9.7470(7)9.8357(8)9.6554(8)9.4278(3)
b (Å)8.830(3)8.8334(4)8.8212(4)8.690(3)8.874(3)8.923(3)8.8223(8)8.9235/4)9.0078(8)8.7998(8)8.5651(4)
c (Å)5.381(4)5.3904(4)5.3815(4)5.293(3)5.326(3)5.289(3)5.3360(5)5.2524(4)5.2408(5)5.2900(6)5.2262(2)
β (°)105.99(2)105.848(3)105.909(3)105.43(2)104.96(2)105.45(2)105.918(13)105.939(4)104.664(6)107.304107.624(2)
V3)448.363(3)450.41(2)447.81(2)434.88(2)438.534(3)440.8(1)441.89(8)439.28(5)449.2429.12402.21(2)
Table 7. Assignment of epidote and clinozoisite bands in the Raman spectra of studied samples.
Table 7. Assignment of epidote and clinozoisite bands in the Raman spectra of studied samples.
ES-9CP-9Assignment[43]Assignment[44]Assignment
136 ν(M-O)130 134
170166ν(M-O)165 168
232235ν(M-O)230 230
273272ν(M-O)275 276
310 ν(M-O)- 314
345347ν(M-O)350ν(M-O)350ν(M-O)
398390ν(Si-Ob-Si)390ν(M-O)390ν(M-O)
424ν(Si-Ob)430ν(Si-Ob-Si)430ν(Si-Ob-Si)
447450ν(Si-Ob)450ν(Si-Ob-Si)452ν(Si-Ob-Si)
505506ν(Si-Ob)510ν(M-O)508ν(M-O)
557569ν(Si-Ob)565ν(Si-Ob-Si)565ν(Si-Ob-Si)
607605ν(Si-Onb)600ν(Si-Ob)599ν(Si-Ob)
843 ν(Si-Onb)830ν(Si-Onb)832ν(Si-Onb)
-868 --864ν(Si-Onb)
882887ν(Si-Onb)880ν(Si-Onb)885ν(Si-Onb)
913912ν(Si-Onb)920ν(Si-Onb)914ν(Si-Onb)
975981ν(Si-Onb)980ν(Si-Onb)980ν(Si-Onb)
10381039ν(Si-Onb)1040ν(Si-Onb)1040ν(Si-Onb)
10871091ν(Si-Onb)1085ν(Si-Onb)1084ν(Si-Onb)
Table 8. Assignment of breakdown products bands in the Raman spectra of studied samples compared to anorthite (An), Ca-tschermakitic pyroxene (Ca-Tsch Px), aegirine (Aeg), and wollastonite (Wol).
Table 8. Assignment of breakdown products bands in the Raman spectra of studied samples compared to anorthite (An), Ca-tschermakitic pyroxene (Ca-Tsch Px), aegirine (Aeg), and wollastonite (Wol).
ES-10ES-11ES-12CP-10CP-11CP-12An[45]Ca-Tsch Px[45]Aeg[46]Wol[46]
137136133139lattice mode132lattice mode
183182181178178176182lattice mode
199199202200lattice mode
209210208 206lattice mode
258252252258253lattice mode
286282289275279282273lattice mode 273M-O stretch/bend
326325327317319317316lattice mode
368369367369369lattice mode370lattice mode
400405396397400396400lattice mode
423426423423427νs (T-O-T)
480478479481484νs (T-O-T)
495497499499498499 499O-Si-O bend
543544545549549548553 549 544Si-Obr stretch
632631 645 651νs (T-O-T) 635Si-Obr stretch
668671671675675672681 676 678Si-Obr stretch
708
737736735736733 741
760756753757755748755
897889887 887Si-Onbr stretch
909906915908νas (Si-O-Si)
922νs (T-O)
945938942944939 949νas (Si-O-Si)953νs (T-O)953Si-Onbr stretch
965965966970964973 968Si-Onbr stretch
997997991997998νas (Si-O-Si)
10221008 1044 1032 1044Si-Onbr stretch1042Si-Onbr stretch
Table 9. Parameters of 57Mössbauer spectra for epidote samples. Color of dublets is for their identification in the Figure 8.
Table 9. Parameters of 57Mössbauer spectra for epidote samples. Color of dublets is for their identification in the Figure 8.
No.Doublet AreaIS (mm/s)QS (mm/s)Γ (mm/s)MineralSite
ES-0D 1 1.000.362.040.30epidoteFe3+, M3
ES-9D 1 0.600.352.090.26epidoteFe3+, M3
D 2 0.400.381.900.42epidoteFe3+, unspecified
ES-10D 1 0.080.251.670.22pyroxeneFe3+, T
D 2 0.280.311.340.38pyroxeneFe3+, T
D 3 0.400.360.970.35pyroxeneFe3+, T
D 4 0.240.380.560.31pyroxeneFe3+ M1
ES-11D 1 0.120.251.650.26pyroxeneFe3+, T
D 2 0.190.291.330.31pyroxeneFe3+, T
D 3 0.380.341.010.37pyroxeneFe3+, T
D 4 0.310.380.610.34pyroxeneFe3+ M1
ES-12D 1 0.080.251.680.25pyroxeneFe3+, T
D 2 0.180.291.410.33pyroxeneFe3+, T
D 3 0.360.351.090.37pyroxeneFe3+, T
D 4 0.370.380.680.38pyroxeneFe3+ M1
Table 10. Parameters of 57Mössbauer spectra for clinozoisite samples. Color of dublets is for their identification in the Figure 8.
Table 10. Parameters of 57Mössbauer spectra for clinozoisite samples. Color of dublets is for their identification in the Figure 8.
No.Doublet AreaIS (mm/s)QS (mm/s)Γ (mm/s)MineralSite
CP-0D 1 0.370.232.320.56epidoteFe3+, M3’
D 2 0.310.362.030.29epidoteFe3+, M3
D 3 0.150.501.780.39epidoteFe3+, M1
D 4 0.170.841.170.50epidoteFe3+, 6
CP-9D 1 0.400.352.110.25epidoteFe3+, M3
D 2 0.600.371.960.35epidoteFe3+, M1
CP-10D 1 0.160.251.590.32pyroxeneFe3+, T
D 2 0.430.331.150.41pyroxeneFe3+, T
D 3 0.410.380.690.38pyroxeneFe3+, M1
CP-11D 1 0.200.191.630.54pyroxeneFe3+, T
D 2 0.380.341.140.41pyroxeneFe3+, T
D 3 0.420.380.670.39pyroxeneFe3+, M1
CP-12S 1 * 0.220.41−0.200.34hematiteFe2O3
D 1 0.380.361.120.63pyroxeneFe3+, T
D 2 0.400.380.580.48pyroxeneFe3+, M1
* Hyperfine magnetic field, Bhf = 51.6 T.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kozáková, P.; Miglierini, M.; Čaplovičová, M.; Škoda, R.; Bačík, P. Structural Breakdown of Natural Epidote and Clinozoisite in High-T and Low-P Conditions and Characterization of Its Products. Minerals 2022, 12, 238. https://doi.org/10.3390/min12020238

AMA Style

Kozáková P, Miglierini M, Čaplovičová M, Škoda R, Bačík P. Structural Breakdown of Natural Epidote and Clinozoisite in High-T and Low-P Conditions and Characterization of Its Products. Minerals. 2022; 12(2):238. https://doi.org/10.3390/min12020238

Chicago/Turabian Style

Kozáková, Petra, Marcel Miglierini, Mária Čaplovičová, Radek Škoda, and Peter Bačík. 2022. "Structural Breakdown of Natural Epidote and Clinozoisite in High-T and Low-P Conditions and Characterization of Its Products" Minerals 12, no. 2: 238. https://doi.org/10.3390/min12020238

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop