Next Article in Journal
Investigation of Gold Recovery and Mercury Losses in Whole Ore Amalgamation: Artisanal Gold Mining in Nambija, Ecuador
Next Article in Special Issue
An Intensity Tensor and Electric Field Gradient Tensor for Fe3+ at M1 Sites of Aegirine–Augite Using Single-Crystal Mössbauer Spectroscopy
Previous Article in Journal
Amp-TB2 Protocol and Its Application to Amphiboles from Recent, Historical and Pre-Historical Eruptions of the Bezymianny Volcano, Kamchatka
Previous Article in Special Issue
Advances in Analysis of the Fe-Ni-Co Alloy and Iron-Bearing Minerals in Meteorites by Mössbauer Spectroscopy with a High Velocity Resolution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

57Fe Mössbauer Spectroscopy and X-ray Diffraction of Annealed Highly Metamict Perrierite: Activation Energy and Recrystallization Processes

by
Dariusz Malczewski
1,*,
Agnieszka Grabias
2,
Maria Dziurowicz
1 and
Tomasz Krzykawski
1
1
Institute of Earth Sciences, University of Silesia, Będzińska 60, 41-200 Sosnowiec, Poland
2
Łukasiewicz Research Network—Institute of Microelectronics and Photonics, Al. Lotników 32/46, 02-668 Warszawa, Poland
*
Author to whom correspondence should be addressed.
Minerals 2023, 13(11), 1395; https://doi.org/10.3390/min13111395
Submission received: 11 September 2023 / Revised: 27 October 2023 / Accepted: 28 October 2023 / Published: 30 October 2023

Abstract

:
This paper presents the results of 57Fe Mössbauer spectroscopy and X-ray diffraction analysis of highly metamict perrierite (REE,Ca,Th)4(Fe2+,Mg)2(Ti,Fe3+)3Si4O22 after annealing in argon from 673 to 1273 K for one hour. Radioactive elements in metamict minerals damage crystal structure on geologic time scales primarily due to recoil nuclei from α-decay of 238U, 232Th, 235U, and their daughter products. Metamict minerals are widely used in geochronology and can serve as natural analogs for the study of radiation effects in high-level nuclear waste. Analyses were performed on fragments of a perrierite sample collected from granitoids near Amherst, Virginia (USA). Electron microprobe and gamma-ray spectrometry recorded Fe concentrations of 4.7 wt.% and Th and U concentrations of 0.64 and 0.06 wt.%, respectively. The calculated total absorbed α-dose was 7.8 × 1015 α-decay mg−1. The Mössbauer spectrum of the untreated sample can be fitted to two Fe2+ and two Fe3+ doublets in octahedral coordination with a relative ΣFe2+/ΣFe of 0.63. For samples annealed at 1173 K and 1273 K, spectra show a decrease in the total contribution of Fe2+ to 0.58 due to dehydroxylation associated with the simultaneous oxidation of post-metamict Fe2+ to Fe3+. In the examined perrierite, Fe2+ occurs in structural positions B and C(1). The broad, predominant Fe3+ doublet observed in the spectrum of the unannealed sample splits into two components at 973 K interpreted to represented positions C(1) and C(2) in the perrierite structure. The Mössbauer spectra show a prominent decrease in the width of the high-energy absorption peak representing Fe2+ components with increasing temperature. The variation in the width of this peak versus the annealing temperature seems to be an indicator of thermally induced recrystallization. Based on the exponential dependence of the derivative function of the parameter with the inverse temperature and using an Arrhenius plot, an activation energy (EA) of 0.73 eV was determined for thermally-induced recrystallization. Corresponding XRD data show progressive recrystallization with increasing annealing temperature. The XRD pattern of the fragment annealed at 1273 K indicates that highly metamict perrierite recrystallized to the pre-metamict state that can be indexed to the C2/m space group.

1. Introduction

Perrierite is a REE-Ti-bearing accessory mineral often found in a metamict state due to the presence of Th and, to a lesser degree, U in its crystal structure. It was first recorded in shore sands of Nettuno, Italy in 1950 [1]. Perrierite shows a strong chemical and structural similarity to chevkinite, and previous works have suggested that it is the same mineral [2]. Later studies reported that these are distinct minerals with slight structural differences [3,4,5,6,7]. The perrierite-chevkinite mineral group (sorosilicates) can be represented by the idealized formula A3+4B2+C3+2Ti4+2Si4O22, where A = REE, Th, Ca, Sr, Na, and K in 8- and 10-fold coordination, B = Fe2+, Mg, Mn, and Ca in 6-fold coordination, and C = Ti, Mg, Mn, Fe3+, Fe2+, Nb, Al, and Zr in 6-fold coordination. Cerium (Ce) represents the dominant rare earth element, and many substitutions are possible [8,9,10,11].
The proposed polyhedral representation of chevkinite–perrierite group minerals [11] consists of four octahedrally coordinated M-sites, where the M(1) site is predominantly occupied by Fe2+ (position B), and the remaining M(2), M(3), and M(4) sites are occupied primarily by Ti and Fe3+ (positions C(2) and C(1)) [8]. Position A consists of two polyhedra, A(1) and A(2), with respective coordination numbers of 8 and 10. The remarkable similarity of both minerals raises the possibility that they may be polymorphs. Studies on synthetic chevkinites and perrierites have shown that the size of the cations at the A, B, and C positions determines which phase will be formed, and they do not qualify as polymorphs according to the formal definition [6]. Substitutions of a wide range of cations make natural minerals more complicated, however, and the crystal structure in question generally exhibits behaviors similar to synthetic equivalents. Chevkinite–perrierite group minerals exhibit polymorphism or polymorphic behavior during temperature-induced transitions. The report [4] describes how three out of five samples of chevkinite annealed under atmospheric air at temperatures of 1000 °C and 1350 °C transformed into perrierite within one hour. The transformations did not occur among samples annealed in argon, nor was the reverse transition (perrierite to chevkinite) observed during annealing under atmospheric air at the same temperatures. These observations have led to the suggestion that perrierite is oxidized chevkinite. This result was partially confirmed for a sample of chevkinite from Virginia [5]. Extensive studies of synthetic chevkinites and perrierites have shown the opposite behavior. Compounds crystallized as perrierites transition to the chevkinite after heating at 1070 °C [7]. In general, among natural and synthetic minerals, thermal polymorphism can be observed for specimens or compounds whose chemical composition lies close to a phase boundary. These may transform from one phase to another due to a small change in the ratio of ionic sizes.
This study used Mössbauer spectroscopy and X-ray diffraction to trace the recrystallization process of highly metamict perrierite following high-temperature annealing in argon. To our knowledge, no published literature has addressed transformation of local crystal structure around Fe ions during annealing of metamict perrierite. Mössbauer spectroscopy results were also used to determine the activation energy for thermally induced recrystallization of highly metamict perrierite. The activation energy helps determine the sensitivity of metamict phases to amorphization and thermal recrystallization. The higher the activation energy for thermal recrystallization, the easier it will be for a mineral to become amorphized by radiation damage due to recoil nuclei from α-decay of U and Th series constituents. Table 1 summarizes the limited data available on activation energies (obtained by different methods) for recrystallization processes in metamict minerals.

2. Materials and Methods

2.1. Sample Description and Chemical Analysis

The perrierite–(Ce) example investigated here exhibits mostly grayish black and vitreous luster with some dull areas (Figure 1). It was collected from a pegmatite near Amherst, Virginia (USA) [5]. After splitting the large sample into fragments, the pieces were placed in quartz tubes, sealed under argon, and annealed for one hour in a muffle furnace at 673, 773, 973, 1073, 1173, and 1273 K. Each temperature was stabilized within ±2 K. After annealing, the samples were quenched and ground to a powder.
The composition of the sample was determined using a JEOL JSM-6480 scanning electron microscope with an energy-dispersive X-ray spectrometer (SEM–EDS) operated at an accelerating voltage of 20 kV and a beam current of 30 μA. Data were analyzed using the EDS2006 software package. Electron-microprobe data were obtained from two inner fragments of the specimen subjected to 26 total analytical points. The concentrations of 232Th, 238U, and 235U were calculated based on gamma-ray activities of 228Ac (232Th), 226Ra, 214Pb, and 214Bi (238U), and assuming a natural 238U/235U ratio of 137.88 [19]. Gamma-ray spectra were collected using a GX4018 system consisting of a coaxial HPGe detector (45.2% efficiency) in a lead and cooper shield (102 mm) with a Lynx multichannel buffer. Table 2 lists the chemical composition and the calculated α-doses.
The studied perrierite exhibited a total absorbed α-dose of 7.8 × 1015 α-decay mg−1 (Table 2). The calculated 232Th α-dose of 5.5 × 1015 α-decay mg−1 was more than twice that of 238U (2.1 × 1015 α-decay mg−1). The sample also exhibited high Nd content (6.2 wt.%). Similar Nd concentrations were noted for samples from two localities near Amherst in one of the first published reports on perrierites and chevkinites from Virginia [5].

2.2. Mössbauer Spectroscopy and X-ray Diffraction

For 57Fe Mössbauer spectroscopy, the powder fragments were prepared as a thin disc absorber with surface densities close to 10 mg cm−2. Mössbauer transmission spectra of the untreated and annealed fragments were recorded at room temperature using a constant acceleration spectrometer with a linear velocity reference signal of triangular shape in the range of ±4.5 mm s−1, and a linear arrangement of a 57Co/Rh source (=50 mCi, 295 K), an absorber, and a proportional counter. For each sample, a few million counts per channel were collected by a multichannel analyzer in 512 channels and then folded to 256 channels. Mössbauer spectra were numerically analyzed by the fitting software Recoil, and all doublets were fitted with Lorentz lines. The velocity scale was calibrated using α-Fe foil about 25 µm thick as a reference absorber at 295 K, and the calibration spectrum was fitted using Lorentz lines with the full width at a half maximum of 0.27 mm s−1.
The untreated and annealed fragments were analyzed for their X-ray diffraction (XRD) patterns using a PANanalytical X’Pert PRO MPD diffractometer in the Θ−Θ system and CoKα radiation in the scan mode with a 0.02° step size and a 298 s counting time for each step. Computational analyses were performed using HIGHSCORE+ PANanalytical v.4.6 software. Numerical calculations to determine the activation energy were performed using the OriginPro 2023b software (OriginLab Corporation, Northampton, MA, USA).

3. Results and Discussion

3.1. Mössbauer Spectroscopy

Figure 2 and Figure 3 show 57Fe Mössbauer spectra and the corresponding XRD patterns of the respective untreated and annealed perrierite samples as a function of annealing temperature. Table 3 lists the hyperfine parameters derived from the fitting procedure for each annealed sample. Isomer shift values are given relative to the α-Fe standard.
The Mössbauer spectrum of the untreated sample shows two quadrupole doublets assigned to Fe2+ (nos. 1 and 2) and two quadrupole doublets assigned to Fe3+ (nos. 3 and 4), all in octahedral coordination (Figure 2a, Table 3). The powder XRD pattern (Figure 2b) shows that the untreated sample is highly metamict with an absorbed α-dose of 7.8 × 1015 α-decay mg−1. Previous studies suggest that the amorphization dose approaches 1016 α-decay mg−1 for metamict silicates and 1017 α-decay mg−1 for metamict oxides [20,21,22,23,24,25].

3.1.1. Changes in Fe2+ Components

For the unannealed sample, doublet no. 1 with δ = 1.13 mm s−1 and Δ = 2.25 mm s−1 makes the largest contribution (0.39 on average) to Mössbauer spectra across all annealing temperatures. This doublet’s δ and Δ values are almost the same for samples annealed from 673 K to 1273 K and average 1.17 mm s−1 and 2.19 mm s−1, respectively. These values indicate the relatively regular coordinating octahedra. According to the accepted description of the structure of the chevkinite–perrierite group, this doublet should be assigned to Fe2+ in position B and to the M1 site in the polyhedral representation.
Relative to doublet no. 1, doublet no. 2 consists of a Fe2+ component with noticeably smaller values representing an isomer shift and a quadrupole splitting. The unannealed sample and samples annealed from 673 K to 1173 K yielded average δ and Δ values of 1.07 mm s−1 and 1.68 mm s−1, respectively. This contribution decreased with increasing annealing temperature and a simultaneous increase in doublet no. 6 from Fe3+ (Figure 3, Table 3). The Fe2+ doublets with low δ and Δ values mostly appeared in octahedral positions showing tetragonal deformation, including radiation-damaged metamict minerals [26,27,28,29]. After recrystallization at 1273 K, this doublet exhibited higher quadrupole splitting (2.02 mm s−1) values, while the isomer shift did not change (Table 3). Hyperfine parameters for doublet no. 2 and the observed increase in the doublet no. 6 contribution from Fe3+ indicate that this component represents Fe2+ at the C(1)/M2 position. The perrierite structure shows that both Fe2+ and Fe3+ are present in this position. As seen in Figure 4, the relative contribution of ƩFe2+ varies from 0.56 to 0.67. The observed decrease in the contribution of Fe2+ above 773 K most likely reflects Fe3+ reduction to Fe2+ in damaged mineral regions due to post-damage diffusion of H2 (as OH) into the structure. Conversely, recrystallization in argon can be accompanied by dehydroxylation with simultaneous oxidation of excess post-metamict Fe2+ to Fe3+. This effect was observed in davidite and titanite [18,30]. The large contribution of doublet no. 2 in spectra for the unannealed sample and those annealed at lower temperatures results from the metamictization process. In untreated allanite–(Ce) specimens with various degrees of metamictization, the ratio of ƩFe2+ to ƩFe varied from 0.40 to 0.73 [31,32]. For a sample of fully metamict allanite, heating in atmospheric air at 1000 K causes only partial oxidation of Fe2+ to Fe3+ [32]. The metamict state strengthens the Fe2+ fraction.
According to the perrierite structure, the bond lengths of Fe with apex oxygens, e.g., Fe−O(7) or Fe–O(6) [8,11], should be significantly shorter, by about 0.15–0.20 Å, relative to the average bond length of the other four coordinating oxygens in the M1 site. In the Mössbauer spectrum, the component derived from such a Fe2+ position should be characterized by hyperfine parameters similar to those observed for doublet no. 2 up to 1173 K. However, according to the δ and Δ values, doublet no. 1 represents a more regular octahedral site.
The only known published report of the Mössbauer spectrum of a chevkinite sample shows two Fe2+ doublets with a total contribution of 0.43 [33]. The first with a contribution of 0.09 and δ and Δ values of 1.25 mm s−1 and 2.21 mm s−1, respectively, was assigned to Fe2+ in the eight-coordinated A site. A feature with these parameters did not appear in our data. The second doublet assigned to position B, representing a contribution of 0.34 (within error), and with values of δ = 1.14 mm s−1 and Δ = 2.25 mm s−1, corresponds exactly to doublet no. 1 in the perrierite spectra. The authors did not describe the degree of metamictization of the chevkinite sample, but fitted line widths suggest that the sample was partially metamict.

3.1.2. Changes in Fe3+ Components

As previously noted, in the spectrum of the unannealed perrierite sample, Fe3+ iron is represented by two doublets numbered 3 and 4. The large width of doublet no. 3 absorption line indicates that it results from a larger number of slightly different but overlapping Fe3+ components with a continuous δ and Δ distribution. This would result from the deformation of the coordinating octahedrons during metamictization. Starting at 973 K, doublet no. 3 was divided into two new components labeled as doublets 5 and 6 (Figure 3c–f, Table 3). Doublet no. 5 for samples annealed from 973 K to 1273 K exhibits a low isomer shift averaging 0.22 mm s−1 and a relatively high quadrupole splitting averaging 1.29 mm s−1. The contribution of this doublet changed slightly from 0.17 to 0.19. The isomer shift and quadrupole splitting values indicate that this doublet represents the C(2)/M4 position in the perrierite structure. Based on studies of synthetic perrierite crystals [8], position C(2) may have extremely different bond lengths of 1.79 Å and 2.42 Å for the two coordinating oxygens. This causes a very strong interaction with the former and a weak interaction with the latter. Such asymmetric octahedra can yield the Fe3+ component with hyperfine parameters observed for doublet no. 5.
Doublet no. 6 is characterized by an average δ of 0.29 mm s−1 and an average Δ of 0.75 mm s−1, as well as an increasing contribution from 0.12 to 0.23 in the spectrum of the sample annealed at 1273 K (Figure 3f, Table 3). The increase in the contribution of this doublet reflects the decrease in the doublet no. 2 Fe2+ contribution. These observations, along with the hyperfine parameter values, suggest a doublet assignment to the C(1)/M2 position [8,9,10,11]. The average value of isomer shifts and quadrupole splittings of doublets 5 and 6 indicate octahedral coordination of Fe3+. The Fe3+ in tetrahedral coordination should be characterized by δ ~ 0.17 mm s−1 and Δ~ 0.50 mm s−1, respectively. Such components do not occur in the examined perrierite [34,35].
Doublet no. 4 present in the Mössbauer spectra of samples annealed at 1173 K exhibits the highest average δ value of 0.42 mm s−1, Δ equal to 1.92 mm s−1 and also the smallest contribution of previously discussed Fe3+ components (Figure 3, Table 3). As indicated by its absence in the fully crystalline sample after annealing at 1273 K and the slight increase in doublet 5 and 6 contributions, this doublet probably represents the most distorted Fe3+ octahedra at the M2 and M4 sites.
In chevkinite, two Fe3+ doublets were reported with hyperfine parameter values of δ = 0.43 mm s−1 and Δ = 0.57 mm s−1, and δ = 0.33 mm s−1 and Δ = 1.34 mm s−1, respectively, with contributions of 0.36 and 0.21, respectively. These correspond to doublets 6 and 5 in the spectra of the studied perrierite. The data indicate lower isomer shifts by the main perrierite Fe3+ components relative to their structural counterparts in the chevkinite analyzed thus far. The relations also suggest potentially smaller-sized Fe3+ octahedra in perrierite relative to those in chevkinite [33].

3.2. Unit Cell Parameters

Figure 3g–l shows XRD diffraction patterns for perrierite samples after one hour of annealing in argon at 673, 773, 973, 1073, 1173, and 1273 K. The data show samples recrystallizing with increasing temperature and complete crystallization with sample annealing at 1273 K.
Table 4 lists the cell parameters for the best fit obtained for the C2/m space group after XRD data processing. Figure 5 shows the high degree of agreement between experimental XRD patterns and those calculated for the C2/m space group.
The unit cell parameters of the annealed perrierite sample match those reported for natural and synthetic perrierites. The main lattice parameter that distinguishes perrierite from chevkinite is the “β” lattice constant, which ranges from 113.4° to 113.9° for perrierite and from 100.2° to 101° for chevkinite. In the case of the perrierite sample annealed in this study, the parameter’s value is 113.6°.

3.3. Determination of Activation Energy

The metamictization process causes deformation and displacement of structural polyhedra. Thermal recrystallization restores the structural units to position, obtaining crystalline forms. The activation energy for thermally induced recrystallization of a metamict mineral represents the minimum energy required to restore structural polyhedra and, consequently, the original long-range order that defines the crystalline state. As seen in Figure 3a–f, the absorption line width of the high energy peak from Fe2+ components (nos. 1 and 2) decreases with increasing annealing temperature. This allows a parameter R (Figure 6a,b) to be defined so as to represent the ratio of the line width of the untreated sample (ΓUT) to the line width of samples annealed at temperature T (ΓT). The R term can then serve as input data for estimating the activation energy for recrystallization from Mössbauer spectroscopy data based on existing annealing models [14,36,37].
Figure 6c shows the variation in R as a function of temperature. The data indicate progressive recrystallization between 673 K and 1273 K in argon characterized by a single-stage mechanism. After annealing at 1273 K, the absorption line of the high-energy peak from the superposition of two Fe2+ components was only half the width of the corresponding line observed for the unannealed sample. The derivative of R with respect to temperature, dR/dT, is an exponential function of the inverse temperature, 1/T, and can represent the annealing rate. The derivative function shown in Figure 7a can be fitted by an Arrhenius-type equation as:
R T = A e x p E A k T
where A = 4.592 is a fitted parameter, EA is the activation energy, and k is the Boltzmann’s constant.
The logarithm of the function given by Equation (1) (Figure 7b) plotted vs. the inverse temperature is a straight line:
l n R T = b 1 T a
with slope a = 8436 and b = 1.254. Since a = EA/k, the activation energy EA = 0.73 ± 0.04 eV.
The calculated activation energy of 0.73 eV for highly metamict perrierite falls well below values reported for fully metamict allanite, gadolinite, and zircon (Table 1). However, the value of 0.73 eV for metamict perrierite is comparable to the values of 0.45, 0.58, and 0.33 eV calculated for partially metamict davidite, gadolinite, and microlite [13,15,18]. In a partially metamict state, crystalline domains coexist with irregular amorphous regions. This heterogeneity can lead to lower activation energy measured during thermal recrystallization relative to that obtained from a completely metamict phase. As is the case of gadolinite, the activation energy for fully metamict perrierite is expected to be much higher. The higher activation energy estimated for perrierite relative to that reported for monazite comports with the latter’s structural integrity in the presence of internal radiation and its typical natural, intact crystalline state [38,39].

4. Conclusions

57Fe Mössbauer spectroscopy and XRD patterns of highly metamict perrierite annealed from 673 K to 1273 K in argon revealed progressive recrystallization with increasing annealing temperature. After annealing at 1273 K, metamict perrierite was found to be fully recrystallized and able to be indexed in the C2/m space group. The Mössbauer spectrum of the sample annealed at 1273 K consisted of two Fe2+ doublets representing two different structural positions and two Fe3+ doublets representing the same structural position but two different octahedral sites. The decrease in the width of the line of the high-energy peak from the superposition of the two Fe2+ components with increasing temperature may indicate the annealing rate. Based on changes in this parameter and using an Arrhenius-type equation, the activation energy for the thermal recrystallization of perrierite was determined to be 0.73 eV. Comparison of the perrierite hyperfine spectral parameters with those measured for the one sample of chevkinite analyzed thus far indicates that, within measurement uncertainty, both minerals exhibit nearly the same values for the dominant Fe2+ doublet, but slightly different values for the Fe3+ components.

Author Contributions

Conceptualization, D.M.; methodology and software, D.M., A.G. and T.K.; validation, D.M., A.G. and T.K.; investigation and measurements, D.M., M.D., A.G. and T.K., formal analysis, D.M. and T.K.; resources, D.M., M.D., T.K. and A.G.; data curation, D.M., T.K. and M.D.; writing-original draft preparation, D.M.; writing-review and editing D.M.; visualization, D.M.; funding acquisition, D.M. and M.D.; project administration, D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Centre of Poland grant No. 2018/29/B/ST10/01495.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bonatti, S. Chevkinite, perrierite and epidotes. Am. Miner. 1959, 44, 115–137. [Google Scholar]
  2. Jaffe, H.W.; Evans, H.T., Jr.; Chapman, R.W. Occurrence and age of chevkinite from the Devil’s Slide fayalite quartz syenite near Stark, New Hampshire. Am. Miner. 1956, 41, 474–487. [Google Scholar]
  3. Gottardi, G. The crystal structure of perrierite. Am. Miner. 1960, 43, 1–14. [Google Scholar]
  4. Lima de Faria, J. Identification of Metamict Minerals by X-ray Powder Photographs. In Estudos, Ensaios e Documentos; Junta de Investigações do Ultramar: Lisbon, Portugal, 1964; Volume 112. [Google Scholar]
  5. Mitchell, R.S. Virginia metamict minerals: Perrierite and chevkinite. Am. Miner. 1966, 51, 1394–1405. [Google Scholar]
  6. Ito, J. A study of chevkinite and perrierite. Am. Miner. 1967, 52, 1094–1104. [Google Scholar]
  7. Ito, J.; Arem, J.E. Chevkinite and perrierite: Synthesis, crystal growth and polymorphism. Am. Miner. 1971, 56, 307–319. [Google Scholar]
  8. Calvo, C.; Faggiani, R. A re-investigation of the crystal structures of chevkinite and perrierite. Am. Miner. 1974, 59, 1277–1285. [Google Scholar]
  9. Segalstad, T.V.; Larsen, A.O. Chevkinite and perrierite from the Oslo region, Norway. Am. Miner. 1978, 63, 499–505. [Google Scholar]
  10. Parodi, G.C.; Della Ventura, G.; Mottana, A.; Raudsepp, M. Zr-rich non metamict perrierite-(Ce) from holocrystalline ejecta in the Sabatini volcanic complex (Latium, Italy). Miner. Mag. 1994, 58, 607–613. [Google Scholar] [CrossRef]
  11. Sokolova, E.; Hawthorne, F.C.; Della Ventura, G.; Kartashov, P.M. Chevkinite-(Ce): Crystal structure and the effect of moderate radiation-induced damage on site-occupancy refinement. Can. Miner. 2004, 42, 1013–1025. [Google Scholar] [CrossRef]
  12. Saini, H.S.; Lal, N.; Nagpaul, K.K. Annealing studies of fission tracks in allanite. Contrib. Miner. Petrol. 1975, 52, 143–145. [Google Scholar] [CrossRef]
  13. Lumpkin, G.R.; Foltyn, E.M.; Ewing, R.C. Thermal recrystallization of alpha-recoil damaged minerals of the pyrochlore structure type. J. Nucl. Mater. 1986, 139, 113–120. [Google Scholar] [CrossRef]
  14. Virk, S.H. Single activation energy model of radiation damage in solid state nuclear track detectors. Radiat. Eff. Defects Solids 1995, 133, 87–95. [Google Scholar] [CrossRef]
  15. Janeczek, J.; Eby, R.K. Annealing in radiation damage in allanite and gadolinite. Phys. Chem. Miner. 1993, 19, 343–356. [Google Scholar] [CrossRef]
  16. Meldrum, A.; Boatner, L.A.; Ewing, R.C. Displacive radiation effects in the monazite- and zircon-structure orthophosphates. Phys. Rev. B 1997, 56, 805–814. [Google Scholar] [CrossRef]
  17. Malczewski, D.; Janeczek, J. Activation energy of annealed metamict gadolinite from 57Fe Mössbauer spectroscopy. Phys. Chem. Miner. 2002, 29, 226–232. [Google Scholar] [CrossRef]
  18. Malczewski, D.; Grabias, A.; Dziurowicz, M. Activation energy of annealed, partially metamict davidite by 57Fe Mössbauer spectroscopy. J. Geosci. 2020, 65, 37–44. [Google Scholar] [CrossRef]
  19. Steiger, R.H.; Jager, E. Subcommission on Geochronology: Convention on the use of decay onstants in geo- and cosmochronology. Earth Planet. Sci. Lett. 1977, 36, 359–362. [Google Scholar] [CrossRef]
  20. Murakami, T.; Chakoumakos, B.C.; Ewing, R.C.; Lumpkin, G.R.; Weber, W.J. Alpha-decay event damage in zircon. Am. Miner. 1991, 76, 1510–1532. [Google Scholar]
  21. Woodhead, J.A.; Rossman, G.R.; Silver, L.T. The metamictization of zircon: Radiation dose-dependent structural characteristics. Am. Miner. 1991, 76, 74–82. [Google Scholar]
  22. Weber, W.J. Self-radiation damage and recovery in Pu-doped zircon. Radiat. Eff. Defects Solids 1991, 115, 341–349. [Google Scholar] [CrossRef]
  23. Nasdala, L.; Wenzel, M.; Vavra, G.; Irmer, G.; Wenzel, T.; Kober, B. Metamictisation of natural zircon: Accumulation versus thermal annealing of radioactivity-induced damage. Contrib. Miner. Petrol. 2001, 141, 125–144. [Google Scholar] [CrossRef]
  24. Malczewski, D.; Dziurowicz, M. 222Rn and 220Rn emanations as a function of the absorbed α-doses from select metamict minerals. Am. Miner. 2015, 100, 1378–1385. [Google Scholar] [CrossRef]
  25. Malczewski, D.; Dziurowicz, M.; Krzykawski, T.; Grabias, A. Spectroscopic characterization and thermal recrystallization study of an unknown metamict phrase from Tuften Quarry, southern Norway. Can. Miner. 2018, 56, 365–373. [Google Scholar] [CrossRef]
  26. Dollase, W.A. Mössbauer spectra and iron distribution in the epidote-group minerals. Z. Krist. Cryst. Mater. 1973, 138, 41–63. [Google Scholar]
  27. Ito, J.; Hafner, S.S. Synthesis and study of gadolinites. Am. Miner. 1974, 59, 700–708. [Google Scholar]
  28. Kartashov, P.M.; Ferraris, G.; Ivaldi, G.; Sokolova, E.; McCammon, C.A. Ferriallanite-(Ce), CaCeFe3+AlFe2+(SiO4)(Si2O7)O(OH), a new member of the epidote group: Description, X-Ray and Mössbauer study. Can. Miner. 2002, 40, 1641–1848. [Google Scholar] [CrossRef]
  29. Malczewski, D. Recrystallization in fully metamict gadolinite from Ytterby (Sweden), annealed in air and studied by 57Fe Mössbauer Spectroscopy. Am. Miner. 2010, 95, 463–471. [Google Scholar] [CrossRef]
  30. Hawthorne, F.C.; Groat, L.A.; Rausepp, M.; Ball, N.A.; Kimata, M.; Spike, F.D.; Gaba, R.; Halden, M.N.; Lumokin, G.R.; Ewing, R.C.; et al. Alpha-decay damage in titanite. Am. Miner. 1991, 76, 370–396. [Google Scholar]
  31. Malczewski, D.; Grabias, A. 57Fe Mössbauer Spectroscopy of radiation damaged allanites. Acta Phys. Pol. A 2008, 114, 1683–1690. [Google Scholar] [CrossRef]
  32. Reissner, C.E.; Bismayer, U.; Kern, D.; Reissner, M.; Park, S.; Zhang, J.; Ewing, R.C.; Shelyug, A.; Navrotsky, A.; Paulmann, C.; et al. Mechanical and structural properties of radiation-damaged allanite-(Ce) and the effects of thermal annealing. Phys. Chem. Miner. 2019, 46, 921–933. [Google Scholar] [CrossRef]
  33. Li, Z.; Jin, M.; Liu, M.; Liu, X. Iron distribution in chevkinite. Hyperfine Interact. 1992, 70, 1057–1060. [Google Scholar] [CrossRef]
  34. Hawthorne, F.C. (Ed.) Spectroscopic methods in mineralogy and geology. Reviews in mineralogy. Miner. Soc. Am. 1988, 18, 698. [Google Scholar]
  35. Rancourt, D.G.; Dang, M.Z.; Lalonde, A.E. Mössbauer Spectroscopy of tetrahedral Fe3+ in trioctahedral micas. Am. Miner. 1992, 77, 34–43. [Google Scholar]
  36. Primak, W. Kinetics of processes distributed in activation energy. Phys. Rev. 1955, 100, 1677–1689. [Google Scholar] [CrossRef]
  37. Primak, W. Large temperature range annealing. J. Appl. Phys. 1960, 31, 1524–1533. [Google Scholar] [CrossRef]
  38. Ewing, R.C.; Haaker, R.F. The metamict state: Implications for radiation damage in crystalline waste forms. Nucl. Chem. Waste Manag. 1980, 1, 51–57. [Google Scholar] [CrossRef]
  39. Meldrum, A.; Boatner, L.A.; Weber, W.J.; Ewing, R.C. Radiation damage in zircon and monazite. Geochim. Cosmochim. Acta. 1998, 62, 2509–2520. [Google Scholar] [CrossRef]
Figure 1. Photo of one of the perrierite specimens from Amherst (Virginia) analyzed in this study. The long edge is 2.5 cm in length.
Figure 1. Photo of one of the perrierite specimens from Amherst (Virginia) analyzed in this study. The long edge is 2.5 cm in length.
Minerals 13 01395 g001
Figure 2. (a) 57Fe Mössbauer spectrum of untreated perrierite sample. Solid dots = experimental data, thick solid line = fitted curve, thin solid line = fitted doublets; (b) corresponding XRD pattern.
Figure 2. (a) 57Fe Mössbauer spectrum of untreated perrierite sample. Solid dots = experimental data, thick solid line = fitted curve, thin solid line = fitted doublets; (b) corresponding XRD pattern.
Minerals 13 01395 g002
Figure 3. (af) 57Fe Mössbauer spectra at room temperature of perrierite samples annealed in argon for 1 h at the given annealing temperatures. Solid dots = experimental data, thick solid line = fitted curve, thin solid line = fitted doublets; (gl) corresponding XRD patterns.
Figure 3. (af) 57Fe Mössbauer spectra at room temperature of perrierite samples annealed in argon for 1 h at the given annealing temperatures. Solid dots = experimental data, thick solid line = fitted curve, thin solid line = fitted doublets; (gl) corresponding XRD patterns.
Minerals 13 01395 g003
Figure 4. Relative contributions of Fe2+ vs. annealing temperature.
Figure 4. Relative contributions of Fe2+ vs. annealing temperature.
Minerals 13 01395 g004
Figure 5. Comparison of XRD pattern for a perrierite sample annealed in argon at 1273 K for one hour (top) and a calculated XRD pattern for the C2/m space group (bottom).
Figure 5. Comparison of XRD pattern for a perrierite sample annealed in argon at 1273 K for one hour (top) and a calculated XRD pattern for the C2/m space group (bottom).
Minerals 13 01395 g005
Figure 6. (a,b) represent the width at half maximum of the high-energy peak (Γ) from Fe2+ components fitted with the Lorentzian singlet; (c) variations in the parameter R with annealing temperature. The solid line represents the exponential fit: R = R0 + A exp(b T), where R0 = 1.11(2), A = 6.2(2) × 10−5, and b = 7.48(4) × 10−3. Coefficient of determination R2 = 0.99.
Figure 6. (a,b) represent the width at half maximum of the high-energy peak (Γ) from Fe2+ components fitted with the Lorentzian singlet; (c) variations in the parameter R with annealing temperature. The solid line represents the exponential fit: R = R0 + A exp(b T), where R0 = 1.11(2), A = 6.2(2) × 10−5, and b = 7.48(4) × 10−3. Coefficient of determination R2 = 0.99.
Minerals 13 01395 g006
Figure 7. (a) Plot of R / T vs. T−1. The solid line shows the exponential fit based on Equation (1); (b) corresponding Arrhenius plot of l n ( R / T ) vs. T−1. The solid line represents the linear regression given by Equation (2).
Figure 7. (a) Plot of R / T vs. T−1. The solid line shows the exponential fit based on Equation (1); (b) corresponding Arrhenius plot of l n ( R / T ) vs. T−1. The solid line represents the linear regression given by Equation (2).
Minerals 13 01395 g007
Table 1. Previously reported activation energies for metamict minerals.
Table 1. Previously reported activation energies for metamict minerals.
ReferenceMineralActivation Energy (eV)Methods
[12]Allanite2.3 Fission track annealing
[13]Microlite (partially metamict)0.33 DTA
Betafite (metamict)0.97
[14]Zircon2.87Fission track annealing
3.59 Heavy ion track annealing
[15]Gadolinite (partially metamict)0.58X-ray diffraction
[16]Monazite0.08Irradiation by 800 keV Kr2+
[17]Gadolinite (metamict)1.97Mössbauer spectroscopy
[18]Davidite (partially metamict)0.45Mössbauer spectroscopy
Table 2. Age, chemical composition (wt.%), and calculated α-doses for the perrierite sample examined in this study.
Table 2. Age, chemical composition (wt.%), and calculated α-doses for the perrierite sample examined in this study.
Age1.1(1) Ga a
O26.3(14)
F1.08(43)
Al2.7(10)
Si13.9(9)
Ca4.6(18)
Ti8.3(17)
Fe4.7(10)
Ge0.72(20)
Se1.49(24)
Rb1.81(10)
Zr0.42(24)
Nb0.54(26)
Ce15.9(17)
Nd6.2(10)
La, Gd, Er, Yb, Lu7.9(12)
Ta1.97(41)
Th0.64(3)
U0.060(3)
Total99.23
Calculated total dose (DT) b
(α-decay mg−1)
7.77(78) × 1015
Calculated dose from 232Th (D232) (α-decay mg−1)5.54(53) × 1015
Calculated dose from 238U (D238) (α-decay mg−1)2.09(23) × 1015
Calculated dose from 235U (D235) (α-decay mg−1)0.14(2) × 1014
a Mesoproterozoic age. b Doses were calculated as: D232 = 6 × N232(et λ232 − 1), D238 = 8 × N238(et λ238 − 1), D235 = 7 × N235(et λ235 − 1), and DT = D232 + D238 + D235. N232, N238, and N235 are the present number of atoms of 232Th, 238U, and 235U per milligram, λ238, λ235, and λ232 are the decay constants of 232Th, 238U, and 235U (respectively), and t is the geologic age.
Table 3. Parameters for 57Fe Mössbauer spectra for untreated sample (UT) and annealed perrierite samples (Figure 3).
Table 3. Parameters for 57Fe Mössbauer spectra for untreated sample (UT) and annealed perrierite samples (Figure 3).
Annealing Temperature (K)Doublet
No.
χ2δ
(mm s−1)
Δ
(mm s−1)
Γ
(mm s−1)
Assignment
(CN)
Rel. Area *
UT
11.611.13(1)2.25(2)0.44(1)Fe2+(6)0.37
2 1.07(1)1.73(2)0.44(2)Fe2+(6)0.30
3 0.32(2)1.02(5)0.70(3)Fe3+(6)0.26
4 0.49(4)1.99(4)0.50(3)Fe3+(6)0.07
673
10.971.15(1)2.20(1)0.42(1)Fe2+(6)0.33
2 1.06(1)1.67(2)0.44(1)Fe2+(6)0.32
3 0.28(2)1.04(1)0.58(1)Fe3+(6)0.25
4 0.39(1)1.90(2)0.38(2)Fe3+(6)0.10
773
11.801.13(1)2.23(2)0.44(1)Fe2+(6)0.42
2 1.05(1)1.72(1)0.44(1)Fe2+(6)0.25
3 0.26(1)1.03(3)0.62(2)Fe3+(6)0.27
4 0.44(3)1.87(4)0.40(3)Fe3+(6)0.06
973
11.341.17(1)2.18(1)0.38(1)Fe2+(6)0.36
2 1.07(1)1.66(1)0.42(1)Fe2+(6)0.25
4 0.38(1)1.89(1)0.36(1)Fe3+(6)0.10
5 0.21(1)1.26(3)0.46(3)Fe3+(6)0.17
6 0.27(1)0.77(4)0.42(1)Fe3+(6)0.12
1073
11.151.18(1)2.17(1)0.38(1)Fe2+(6)0.38
2 1.08(1)1.64(2)0.44(2)Fe2+(6)0.24
4 0.39(1)1.93(1)0.36(3)Fe3+(6)0.08
5 0.21(1)1.27(1)0.48(3)Fe3+(6)0.18
6 0.28(1)0.78(1)0.40(1)Fe3+(6)0.12
1173
10.951.18(1)2.17(1)0.38(1)Fe2+(6)0.37
2 1.08(1)1.67(2)0.56(3)Fe2+(6)0.20
4 0.40(2)1.92(3)0.34(2)Fe3+(6)0.06
5 0.24(1)1.29(1)0.48(3)Fe3+(6)0.17
6 0.30(1)0.74(2)0.44(2)Fe3+(6)0.20
1273
11.541.19(1)2.19(1)0.32(1)Fe2+(6)0.38
2 1.08(1)2.02(2)0.56(3)Fe2+(6)0.20
5 0.23(1)1.34(3)0.44(2)Fe3+(6)0.19
6 0.29(1)0.72(3)0.44(2)Fe3+(6)0.23
* Estimated uncertainty is ΔA/A ≤ 10%.
Table 4. Unit cell dimensions calculated for metamict perrierite after one-hour annealing at 1273 K in argon.
Table 4. Unit cell dimensions calculated for metamict perrierite after one-hour annealing at 1273 K in argon.
NameValue
Unit CellC 1 2/m 1
a (Å)13.5841(8)
b (Å)5.6340(3)
c (Å)11.6722(3)
alpha (°)90
beta (°)113.38(2)
gamma (°)90
V (Å3)818.3592
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Malczewski, D.; Grabias, A.; Dziurowicz, M.; Krzykawski, T. 57Fe Mössbauer Spectroscopy and X-ray Diffraction of Annealed Highly Metamict Perrierite: Activation Energy and Recrystallization Processes. Minerals 2023, 13, 1395. https://doi.org/10.3390/min13111395

AMA Style

Malczewski D, Grabias A, Dziurowicz M, Krzykawski T. 57Fe Mössbauer Spectroscopy and X-ray Diffraction of Annealed Highly Metamict Perrierite: Activation Energy and Recrystallization Processes. Minerals. 2023; 13(11):1395. https://doi.org/10.3390/min13111395

Chicago/Turabian Style

Malczewski, Dariusz, Agnieszka Grabias, Maria Dziurowicz, and Tomasz Krzykawski. 2023. "57Fe Mössbauer Spectroscopy and X-ray Diffraction of Annealed Highly Metamict Perrierite: Activation Energy and Recrystallization Processes" Minerals 13, no. 11: 1395. https://doi.org/10.3390/min13111395

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