Next Article in Journal
Atomized Reagent Addition with Synchronized Jet Pre-Mineralization to Enhance the Flotation Process: Study on Atomization Parameters and Mechanisms of Enhancement
Next Article in Special Issue
The Crystal Chemistry and Structure of V-Bearing Silicocarnotite from Andradite–Gehlenite–Pseudowollastonite Paralava of the Hatrurim Complex, Israel
Previous Article in Journal
Accumulated Copper Tailing Solid Wastes with Specific Compositions Encourage Advances in Microbial Leaching
Previous Article in Special Issue
A New Mineral Calcioveatchite, SrCaB11O16(OH)5·H2O, and the Veatchite–Calcioveatchite Isomorphous Series
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Crystal Chemistry of Boussingaultite, (NH4)2Mg(SO4)2·6H2O, and Its Derivatives in a Wide Temperature Range

by
Elena S. Zhitova
1,2,
Rezeda M. Sheveleva
1,2,
Andrey A. Zolotarev
1,*,
Roman Yu. Shendrik
3,
Elizaveta A. Pankrushina
4,5,6,
Konstantin A. Turovsky
1,
Margarita S. Avdontceva
1,
Maria G. Krzhizhanovskaya
1,
Natalia S. Vlasenko
7,
Anatoly A. Zolotarev
1,
Mikhail A. Rassomakhin
8 and
Sergey V. Krivovichev
1,6
1
Institute of Earth Sciences, St. Petersburg State University, 199034 St. Petersburg, Russia
2
Institute of Volcanology and Seismology FEB RAS, 683006 Petropavlovsk-Kamchatsky, Russia
3
Vinogradov Institute of Geochemistry SB RAS, 664033 Irkutsk, Russia
4
Zavaritsky Institute of Geology and Geochemistry UB RAS, 620110 Ekaterinburg, Russia
5
Institute of Physics and Technology, Ural Federal University, 620002 Ekaterinburg, Russia
6
Kola Science Center, Russian Academy of Sciences, 184209 Apatity, Russia
7
Centre for Geo-Environmental Research and Modelling, St. Petersburg State University, 199034 St. Petersburg, Russia
8
South Urals Federal Research Center of Mineralogy and Geoecology UB RAS, 456317 Miass, Russia
*
Author to whom correspondence should be addressed.
Minerals 2024, 14(10), 1052; https://doi.org/10.3390/min14101052
Submission received: 2 September 2024 / Revised: 17 October 2024 / Accepted: 18 October 2024 / Published: 20 October 2024

Abstract

:
The crystal structure, thermal behavior, and vibrational spectra of the anthropogenic analogue of boussingaultite, (NH4)2Mg(SO4)2·6H2O, and its dehydrated counterpart efremovite, (NH4)2Mg2(SO4)3, were studied in detail. The sample from the Chelyabinsk burning coal dumps has the composition of (NH4)1.92(Mg1.02Mn0.01Fe0.01)∑1.04(SO4)2·6H2O and crystallizes in the space group P21/a, with a = 9.3183(4), b = 12.6070(4), c = 6.2054(3) Å, β = 107.115(5)°, V = 696.70(5) Å3 (at 20 °C), Z = 2. The thermal evolution steps are as follows: boussingaultite (NH4)2Mg(SO4)2·6H2O (25–90 °C) → X-ray amorphous phase (100–150 °C) → efremovite (NH4)2Mg2(SO4)3 (160–340 °C) → MgSO4 Cmcm + Pbnm (340–580 °C) → MgSO4 Pbnm (580–700 °C). Thermal expansion is anisotropic, with the coefficients (×106 °C−1) α11 = 52(2), α22 = 68(2), α33 = –89(3), and αv = 31(3) at T = –123 °C; and α11 = 53(2), α22 = 67(2), α33 = 15(1), and αv = 136(3) at T = 60 °C. The maximal thermal expansion is along the b-axis and is due to straightening of corrugated pseudolayers (within the ab plane) of Mg(H2O)6 octahedra and SO4 tetrahedra with NH4 groups in the interlayer space. Vibrational spectroscopy outlines the general trend of dehydration and deammonization as the difference in the temperature intervals of these transformation steps allows separation of O–H and N–H vibrations in the process of dehydration by infrared and Raman spectroscopy. The intermediate partially dehydrated modification of boussingaultite was detected by in situ Raman spectroscopy at 110 °C that may correspond to ammonium leonite.

Graphical Abstract

1. Introduction

Boussingaultite, (NH4)2Mg(SO4)2·6H2O, belongs to the picromerite group of minerals, with the general formula A2M(XO4)2·6H2O, where, in natural species, A = K or NH4; M = Mg, Cu, Zn, Fe, and Ni; and X = S [1]. The synthetic inorganic compounds with this stoichiometry are known as Tutton’s salts, with broad chemical variability [2]. Only two ammonium and magnesium sulfates are known in nature as minerals: hydrated form, boussingaultite; and dehydrated form, efremovite, (NH4)2Mg(SO4)2. They are usually found in association with each other [3]. Boussingaultite has been known of since 1863 [4], but it can be considered a rare mineral species, in accordance with its known findings (in 14 countries). This is probably the major reason for the fact that most of the structural data in the literature have been reported for the synthetic phases (Tutton’s ammonium salts). The crystal structure of naturally occurring boussingaultite was revealed very recently by neutron diffraction [5]. Boussingaultite, efremovite, and the majority of other ammonium minerals are associated with specific geological environments, such as volcanism (fumaroles) [6,7,8,9] and coal fires (including pseudofumaroles) [10,11,12].
The astrobiological aspect of the crystal chemical study of ammonium minerals and phases is in developing methodologies for nitrogen (that is the fifth most common element in the Solar system) detection in extraterrestrial bodies, like planets and comets [13,14]. The Ni-bearing boussingaultite has been identified recently in carbonaceous chondrite meteorites [15]. The essential feature of nitrogen in ammonium form is its biological accessibility, in contrast to most of nitrogen, which is triple-bonded atmospheric and biologically inaccessible [16]. Thus, ammonium minerals are of high biochemical importance as (1) material prototypes for agricultural fertilizers (see the literature survey by [5]) and (2) the component of prebiotic life, and as evidence of biologically mediated process [16]. On Earth, most of the nitrogen is concentrated in solid material form [17] and transported in NH3 form by magmatic liquids and fumarolic fluids [13]. The latter can produce rare associations of ammonium minerals, as well as the emission of coal-fire pseudofumaroles. Often, the phases formed as a result of coal fires are represented by more perfect crystals suitable for detailed crystal chemical studies, while, as shown by a structural analysis, burning coal dump phases are full structural analogues of fumarolic counterparts [18]. Recently, nitrogen was detected in Martian sediments by the Curiosity rover in the form of nitrate [19]. Taking into account the abundance of sulfates (including K-sulfate jarosite) on Martian surface [14] that could be of post-volcanic origin and approximating to terrestrial conditions, the fixation of nitrogen in the ammonium form seems promising. At the same time, the detection of nitrogen in natural phases by analytical methods is still problematic even for ammonium-dominant ones because nitrogen is a light element, and its peak in energy-dispersive spectra can be almost invisible [14].
Another analytical complication is due to the overlapping positions of N–H and O–H vibrations in Raman and infrared spectra [20,21,22]. Nitrogen detection becomes even more complex when taking into account that most of the ammonium occurs as an impurity in potassium minerals. Therefore, detecting ammonium in extraterrestrial minerals is a complex scientific task that can be solved by the detailed crystal chemical characterization of terrestrial well-crystallized phases, including those found on burnt coal dumps [23,24,25].
Mineralogically, the formation of boussingaultite is often confined to fumarolic and pseudofumarolic environments, which are characterized by different ranges of elevated temperatures at atmospheric pressure. The high-temperature experiments, such as those reported below, can be interpreted as a simulation of boussingaultite formation and transformation under natural conditions. Previously, the high-temperature behavior of the synthetic analogue of boussingaultite (ammonium magnesium sulfate hexahydrate) was studied by calorimetric and thermogravimetric methods [26], and the main transformation steps were revealed: (1) incongruent melting of (NH4)2Mg(SO4)2·6H2O compound detected as two effects at 150–170 °C, (2) evaporation of water ~280 °C, (3) formation of efremovite ~290 °C and (4) phase transition of ammonium sulfate at ~370 °C. The study of Kosova and coauthors [26] revealed anomalies on differential scanning calorimetry (DSC), dielectric and permittivity curves interpreted as hypothesized reversible order–disorder phase transition that can be responsible for the appearance of ferroelectric properties of the material.
The aim of the present research is to characterize the crystal structure of boussingaultite and partly its derivatives under low (similar to Martian conditions) and high (similar to conditions of volcanic fumaroles and coal-fire pseudofumaroles) temperatures and to define N–H and O–H vibrations via the dehydration of boussingaultite.

2. Occurrence

The Chelyabinsk Coal Basin (CCB) is located east of Chelyabinsk City (from the northeast to the south of the city) and has an area of 1300 km2. The CCB area is dominated mainly by brown coal of the Triassic–Jurassic age. Similar to other coal basins, the CCB deposits are subject to spontaneous combustion with the formation of so-called pseudofumaroles, around which some phases crystallize as sublimates or as part of hypergene processes. The status (mineral or technogenic phase) of such sublimates is discussed by the Commission on New Minerals Nomenclature and Classification (CNMNC). Eight minerals from CCB, bazhenovite, godovikovite, dmisteinbergite, svyatoslavite, rorisite, efremovite, srebrodolskite, and fluorellestadite, were approved by the CNMNC during the 1985–1990 period, with the later decision not to consider such phases as minerals. Despite the fact that the CNMNC recently revised its approach to the phases of burnt dumps and began to approve minerals from such environments, in our work, the studied samples are considered technogenic phases.
Burnt dumps of the Chelyabinsk Coal Basin are a unique object of technogenic phase formation. More than 240 different mineral-like compounds were described there, and about 50 of them were unique at the time of their first description. Of particular interest are ammonium-containing phases formed as a result of contact with organic matter at elevated temperatures. At least 16 ammonium-containing phases have been described within the burnt CCB dumps of which two phases have been confirmed as valid mineral species: godovikovite, (NH4)Al(SO4)2 [27]; and efremovite, (NH4)2Mg2(SO4)3 [28]. It is worth noting that the mineral status of godovikovite and efremovite has been confirmed by their findings in solely natural environments such as volcanic fumaroles.
The investigated therein technogenic analogue of boussingaultite (in the text, called boussingaultite for the sake of simplicity) from CCB has been taken from the personal collection of B.V. Chesnokov that is stored in the Natural Science Museum of the Ilmen State Reserve (Miass, Russia), sample No. 057–sch–12 (catalog number). It originates from the burnt dumps of mine No. 43bis, where boussingaultite was described in significant quantities, forming ammonium-type sulfate crusts [29,30]. Boussingaultite from burnt dumps of CCB is considered to be the final product of the hydration of efremovite, (NH4)2Mg2(SO4)3, but the direct crystallization of boussingaultite from ammonium-containing solutions at temperatures up to 100 °C is a second possible scenario [31]. Another ammonium mineral mentioned in association with boussingaultite and efremovite is the ammonium analogue of leonite, K2Mg(SO4)2·4H2O.
The second sample of boussingaultite comes from the Kladno mine (Central Bohemian Region, Czech Republic)—the burning coal waste piles that were stored in the same collection as the sample from CCB.

3. Materials and Methods

3.1. Chemical Composition

Several crystals of boussingaultite were mounted in epoxy blocks and polished. Some crystals were deposited onto carbon tape, and all samples were coated with carbon. The elemental analyses were carried out using a Hitachi S–3400N scanning electron microscope equipped with an energy-dispersive spectrometer Oxford X–Max 20, at an accelerating voltage of 20 kV and a probe current of 1 nA, and with various electron beam diameters of minimum 5 μm due to possible dehydration under the electron beam. The standards are given in Table 1. The energy-dispersive spectra were processed automatically using the AzTec Energy software package.

3.2. Low-Temperature Single-Crystal X-Ray Diffraction Analysis (LT SCXRD)

The crystal structure of boussingaultite at different temperatures was studied by performing a single-crystal X-ray diffraction analysis in the range from −173 up to −73 °C (temperature step was 25 °C). Additional data were obtained at 52 °C to record the presence or absence of order–disorder phase transitions assumed by Kosova and coauthors [26]. The analysis was carried out using the single-crystal four-circle Agilent Technologies (Oxford Diffraction) “Supernova” diffractometer (MoKα radiation, 40 kV and 1.5 mA, frame widths of 1° in ω, and 45 s counting time for each frame) equipped with a low-temperature Oxford Cryostream system. Single-crystal X-ray diffraction analysis data were processed in the CrysAlisPro software package [32], an empirical absorption correction was calculated based on spherical harmonics using the SCALES ABSPACK algorithm. The crystal structures were solved and refined using the ShelX program package [33] within the Olex2 software shell [34]. Crystal data, data-collection information, and structure-refinement details are given in Table 2 (for temperature of −173 and 52 °C); atom coordinates and displacement parameters are in Supplementary Tables S1 and S2, and selected interatomic distances are in Supplementary Table S3. The positions of H atoms were derived from the analysis of Fourier difference electron–density maps and refined freely in an isotropic approximation. The parameters of the hydrogen bonds of boussingaultite (for temperature of −173 °C) are given in Supplementary Table S4. Vesta software was used to visualize the structural data [35].
Additionally, the crystal structure of boussingaultite from Kladno was refined by X-ray diffraction analysis at 20 °C, which was performed using a Rigaku XtaLAB Synergy–S diffractometer (MoKα radiation, 50 kV and 1.0 mA, frame widths 0.5° in ω, and 10 s counting time for each frame) with high-stability, sharp-focus X-ray source PhotonJet–S and a high-speed, direct-action detector HyPix–6000HE. The data were processed using the same software packages as for low-temperature single-crystal X-ray diffraction analysis. All crystal structure data of boussingaultite from Kladno are given in Supplementary Tables S5–S9.

3.3. High-Temperature In Situ Powder X-Ray Diffraction Analysis (HTXRD)

Anisotropy of thermal expansion and high-temperature transformations of boussingaultite were investigated by high-temperature in situ powder X-ray diffraction (HTXRD) analysis, which was carried out using a Rigaku Ultima IV diffractometer (CuKα radiation, 40 kV/30 mA). The sample was studied in the temperature range from 30 up to 700 °C; and the temperature step was 10 °C in the range from 30 to 240 °C and 20 °C in the range from 240 to 700 °C. Unit-cell parameters of boussingaultite were refined using the Rietveld method implemented in the Topas software package [36]. The crystal-structure data obtained at 20 °C were used for unit-cell refinement of HTXRD data. The main thermal-expansion coefficients were calculated using the TTT program package [37].

3.4. Infrared Spectroscopy

The infrared spectra were registered using a Ge–ATR probe of a Micran–3 infrared microscope with a liquid nitrogen cooled MCT detector connected to a Fourier Transform Infrared (FTIR) spectrometer FT–801, made by Simex. The spectra were measured using 32 scans, with a resolution of 1 cm−1, and processed by the Simex software Zair v.3.5. The dehydration was studied on material subsequently heated in a muffle furnace in air for about 10 min at each temperature in the interval between 20 and 400 °C. Attenuated total internal reflection (ATR) spectroscopy in the infrared region was carried out for samples of boussingaultite annealed at temperatures of 150, 200, 250, 300, and 400 °C, as well as for the untreated (initial) sample.

3.5. Raman Spectroscopy

Raman spectra were obtained using the Horiba LabRam HR800 Evolution spectrometer equipped with the confocal Olympus BX–FM microscope, Ar laser (radiation wavelength of 488 nm), and the Linkam TSM 600 heating/cooling stage. A diffraction grating of 1800 gr/mm and an electrically cooled CCD detector were used for recording. The spectrometer calibration was guided along the Rayleigh line and the emission lines of a neon lamp. The spectral resolution was about 1 cm−1. The spatial lateral resolution was about 1 μm. The temperature-dependent spectra were collected during the heating of the sample in the temperature range from −190 to 500 °C, with a step of 1–10 °C (depends on temperature range); the exposure time of the sample to stabilize the temperature was 20 s. The heating rate was 20 °C/min. The accuracy of maintaining the temperature was 0.1 °C. The temperature-measurement range is limited due to the increasing photoluminescent background from the sample. The peak fitting procedure was carried out using LabSpec6 software, ArDI web-application [38], and Origin 2020 by approximating spectra with Voigt contours. The background was subtracted by a third-order polynomial, each spectrum in the equation has its own characteristic coefficients.

4. Results

4.1. Chemical Composition

In the initial boussingaultite sample, the following elements were detected: N, O, Mg, S, and minor Mn and Fe. The data were used for the calculation of empirical chemical formula of boussingaultite on the basis of S = 2 atoms per formula unit (apfu). The measurement gave a somewhat underestimated value for N (Table 1). Therefore, it was corrected in the chemical formula based on charge-balancing requirements. The amount of H2O was calculated based on the crystal-structure data. The obtained formula is (NH4)1.92(Mg1.02Mn0.01Fe0.01)∑1.04(SO4)2·6H2O, and it is close to the ideal one, (NH4)2Mg(SO4)2·6H2O.

4.2. Crystal Structure

The structure refinement based on single-crystal data was performed for six temperatures (at −173, −148, −123, −98, −73, and 52 °C). All refinement results are available as CIF–files from the CCDC database (CSD # 2372190, 2372191, 2372192, 2372193, 2372194, and 2372198, respectively). The crystal structure of boussingaultite was refined in the space group P21/a; the unit-cell parameters (at –173 °C) are a = 9.1992(5), b = 12.4049(7), c = 6.2590(3) Å, β = 107.026(6) °, V = 682.94(7) Å3 to R1 = 0.0392 (Table 2). The calculated powder X-ray diffraction patterns are given in Supplementary Figure S1. The structure refinements performed at other temperatures are available as supplementary data (see data-availability paragraph). In the text, we provide data for T = −173 °C since this is the best refinement with minimized atomic displacement parameters, whereas data at T = 52 °C correspond to the highest temperature and show absence of structure transformation.
The structure of boussingaultite is built from regular Mg(H2O)6 octahedra, with the average <Mg–O> bond length of 2.068 Å; SO4 tetrahedra (<S–O> = 1.479 Å); and NH4 tetrahedra (<N–H> = 0.89 Å) (Table 3 and Supplementary Tables S3 and S8) interconnected by a system of hydrogen bonds (Supplementary Tables S4 and S9). The structure can be considered as consisting of pseudolayers (within the ab plane) of Mg(H2O)6 octahedra and SO4 tetrahedra, while the NH4 groups are located in the interlayer space. In general, the crystal chemical characteristics of boussingaultite studied in this work are similar to those revealed for boussingaultite from Pecs–Vasas, Mecsek Mountains, South Hungary, by neutron diffraction [5], except for the occupancy of monovalent cation that is fully populated by N in our sample and has ~10% of K, in addition to N, in the sample studied by Gatta and coauthors [5]. The hydrogen positions and hydrogen bonding scheme are similar for both cases despite the different techniques used for the localization of H atoms.
No principal structure changes occur in the temperature range from −173 to 52 °C; the main deformations are expansion and contraction in different directions (see below). Through our analysis of geometric parameters of boussingaultite based on atomic displacement parameters and residual electron density, we did not find any indications of disordering of sulfate tetrahedra at 52 °C (or any other polyhedra) that have been suggested previously [26]. The decomposition of boussingaultite was registered at 77 °C by SCXRD. The crystal structure of boussingaultite consists of isolated fragments linked by hydrogen bonds; thus, the thermal behavior can be determined by both the expansion of structural fragments and the distances between them, as shown by the data in Table 3; the second factor is the main one.

4.3. Phase Evolution upon Temperature

The boussingaultite maxima are observed in the HTXRD pattern up to 90 °C (Figure 1 and Supplementary Figure S2). Further heating leads to the formation of an X-ray amorphous phase in the temperature range of 100–150 °C. The efremovite reflections are registered in the range of 160–340 °C, two modifications of MgSO4 (Cmcm and Pbnm) are identified between 340 and 580 °C, and the only Pbnm modification of MgSO4 is observed above 580 °C (Figure 1). Thus, upon boussingaultite heating, the dehydration and deammonization processes were registered at 100–150 °C and ~340 °C, respectively.

4.4. Thermal Expansion

The unit-cell parameters of boussingaultite were refined for seven temperatures in the high-temperature region from 25 to 90 °C, in addition to the low-temperature SCXRD region, with five values in the temperature range from −173 to −73 °C. As shown by Figure 2 for the unit-cell parameters a and b and the unit-cell volume, V, the dependencies of lattice parameters versus temperatures show linear approximation for both low- and high-temperature regions. The c lattice parameter undergoes contraction in the low-temperature region and expansion in the high-temperature region (approximated by two linear dependencies).
Thermal expansion of boussingaultite is anisotropic (Figure 3; Table 4) in low- and high-temperature regions. A nearly isotropic section is obtained for the ab plane within the octahedral–tetrahedral pseudolayer (with α11 and α22). Maximal thermal expansion is observed parallel to the b-axis due to the straightening of corrugated angles in pseudolayer (Figure 3). Minimal thermal expansion (α33) is observed perpendicular to the pseudolayer plane that is 15(1) × 10−6 °C−1 at high temperature and −89(3) × 10−6 °C−1 at low temperature (Table 4). Thus, the structure mainly expands within the plane of pseudolayers, with much less expansion or even contraction in the low-temperature region in perpendicular direction.

4.5. Infrared Spectroscopy

The infrared spectra of boussingaultite (initial sample) contains bands overlapping O–H and N–H stretching bands in the range of 3464–2839 cm−1 and H–O–H, H–N–H bending in the range of 1660–1423 cm−1 (Figure 4) [20,21,22,23,24,25,39,40]; the bands associated with SO4 tetrahedra are found in the range of 1167–1012 cm−1 and 700–657 cm−1, and H–O–H libration mode is registered between 880 and 892 cm−1; the detailed band assignment is given in Table 5. The sample annealed at 150 °C corresponds to partly dehydrated boussingaultite and possibly amorphous phase characterized by band broadening. The samples annealed at 200, 250, and 300 °C correspond to efremovite (the sample treated at 200 °C is partly hydrated). The sample annealed at 400 °C contains MgSO4 [41] (Table 5). In general, the phase-evolution sequence determined by infrared spectroscopy and diffraction methodologies coincides, but the transition temperatures are shifted to higher values due to the difference in the heating kinetics.
The transformation of boussingaultite to efremovite results in the decrease in intensities associated with O–H vibrations and disappearance of such bands at temperatures higher than 200 °C. At the same time, deammonization is complete at 400 °C; thus, for boussingaultite annealed at 250 and 300 °C, the O–H related bands are no longer observed, while the N–H bands remain. This allows for the separation of the N–H bonds from the O–H ones. The dehydration process is accompanied by the amorphization of boussingaultite, clearly seen by the broadening of the infrared absorption bands in the sample heated at 150 °C (Figure 4). The bands at 3464, 3350, 3233, 3155, 3040, and 3012 cm−1 are associated with the O–H stretching modes. The bands corresponding to N–H stretching modes are observed at 3272, 3189, 3124, 3077, 2980, 2920, and 2839 cm−1. The H–O–H bending mode is found at 1660 cm−1 that is rather typical for that vibration. The band at ~1450–1460 cm−1 is split into two components peaked at 1463 and 1468 cm−1. The 1468 cm−1 component disappears at temperatures higher than 150 °C that may be due to either a change in the geometry of NH4 cation or the loss of perturbed H2O. The component at 1463 cm−1 remains above 150 °C and attributed to the H–N–H bending vibration. The bands in the region 1200–600 cm−1 correspond to the different vibration modes of SO4 tetrahedra [20,21,22,23,24,25,39,40]. In the samples heated at 150 and 200 °C, the bands in the interval of 880–890 cm−1 could be related to H–O–H libration modes (Table 5).

4.6. Raman Spectroscopy

The 45 of the 114 Raman-active vibration modes were recorded in the spectra of boussingaultite (Figure 5), all predicted via theoretical group analysis: Γ = 57Ag + 59Au + 57Bg + 58Bu (Bilbao Crystallographic Server [42]), where Ag and Bg are Raman-active modes, while Au and Bu are IR-active modes. The detection of weak modes can be hampered by spectrally broad photoluminescence emission that is possibly due to the presence of Fe3+. The Raman spectra of the boussingaultite crystal are characterized by narrow bands, which reflect a high degree of structural perfection. The band assignment, in general, agrees with those reported previously [20]; however, the 3500–2800 cm−1 region was analyzed in more detail with the separation of O–H and N–H vibrations. The band assignment is listed in Table 6.
Table 6. Raman bands of boussingaultite and its high-temperature derivatives, efremovite, magnesium sulfate, and possibly ammonium leonite.
Table 6. Raman bands of boussingaultite and its high-temperature derivatives, efremovite, magnesium sulfate, and possibly ammonium leonite.
BoussingaultiteAmmonium
Leonite
Amor-
Phous
EfremoviteMgSO4Phase
Raman Band Maxima (cm−1) at Temperatures (°C)
−190−110108095110120155200Band Assignment
343234443436343834393439ν(H2O)
339033883389339133963405
336433633364336433633372
3339333833393339n.i.*n.i.
33093308330933093319n.i.
327932783273327632793268
32393239323632373241n.i.ν(NH4)
32023199320031993204n.i.
315231503151315231553166ν(H2O)
31303128312731273131n.i.ν(NH4)
31093110311131103118n.i.ν(H2O)
309130913089309030853076
3063306830673066n.i.n.i.ν(NH4)
3014302430223021n.i.n.i.
2925n.i.n.i.n.i.n.i.n.i.
2872n.i.n.i.n.i.n.i.n.i.
2844n.i.n.i.n.i.n.i.n.i.
17661766176917701765n.i.ν4(NH4) + δ(HOH)
171417121713170616981701
1678167616791679n.i.n.i.
1423142914311433n.i.n.i.n.i.ν4(NH4)
11441139113911361200n.i.n.i.n.i.ν3(SO4)
109110911093109411021100n.i.n.i.
1070106910711070n.i.n.i.n.i.n.i.
1059106210621060n.i.n.i.n.i.n.i.
982982983982981996101010451053ν1(SO4)
803797800798n.i.ν(H2O)
722719717716n.i.
630627627626639n.i.642n.i.n.i.ν4(SO4)
617617617618612610617n.i.n.i.
610610610610n.i.n.i.n.i.n.i.
592583583583n.i.n.i.n.i.n.i.
552557558556n.i.n.i.ν(H2O)
465463461459464463479n.i.n.i.ν2(SO4)
451451451450446446448n.i.n.i.
391387n.i.n.i.n.i.n.i.ν(H2O)
368367365361n.i.n.i.
322320n.i.n.i.n.i.n.i.
301306306306n.i.n.i.
256255254253n.i.n.i.n.i.n.i.Lattice mode
227226224221n.i.n.i.n.i.n.i.
201195191189n.i.n.i.n.i.n.i.
191191185183n.i.n.i.n.i.n.i.
172174172173n.i.n.i.n.i.n.i.
164163159158n.i.n.i.n.i.n.i.
137133128127n.i.n.i.n.i.n.i.
129129125122n.i.n.i.n.i.n.i.
n.i.—the exact wavenumber is not identified due to low signal-to-noise ratio; however, the band is visible in Figure 6 and Figure 7; a dash means the absence of a corresponding vibration in the spectrum. Interpretation of vibrations according to [20,24,39,43].
Figure 5. Raman spectra of boussingaultite at different temperatures by in situ measurements (a) and detailed 3600–2800 cm−1 region at different temperatures (b).
Figure 5. Raman spectra of boussingaultite at different temperatures by in situ measurements (a) and detailed 3600–2800 cm−1 region at different temperatures (b).
Minerals 14 01052 g005
Figure 6. Raman spectra of boussingaultite and its high-temperature derivatives: partly dehydrated modification interpreted as ammonium leonite at 95, 110 °C, amorphous phase at 120 °C, and efremovite at 155 °C) in 2000–100 cm−1 (a), 1100–900 cm−1 (b), and 3600–3000 cm−1 (d) regions. Frame (c) shows the dependence of the ν1(SO4) mode position on the temperature of the boussingaultite (black dots) and ammonioleonite (red dots) in the temperature range from −190 to 90 °C.
Figure 6. Raman spectra of boussingaultite and its high-temperature derivatives: partly dehydrated modification interpreted as ammonium leonite at 95, 110 °C, amorphous phase at 120 °C, and efremovite at 155 °C) in 2000–100 cm−1 (a), 1100–900 cm−1 (b), and 3600–3000 cm−1 (d) regions. Frame (c) shows the dependence of the ν1(SO4) mode position on the temperature of the boussingaultite (black dots) and ammonioleonite (red dots) in the temperature range from −190 to 90 °C.
Minerals 14 01052 g006
Figure 7. The dependence of corrugated angle of pseudolayer from temperature in boussingaultite structure.
Figure 7. The dependence of corrugated angle of pseudolayer from temperature in boussingaultite structure.
Minerals 14 01052 g007
The Raman bands in the 3500–2800 cm−1 region correspond to an overlap of N–H and O–H vibrations [20,21,22,23,24,25,39,40] that have been assigned in detail in accord with infrared spectroscopy (Figure 5b; Supplementary Figure S3 and the procedure described as note to it). The O–H stretching is revealed by the following band components (at T = 10 °C): 3436, 3389, 3364, 3339, 3309, 3273, 3151, 3111, and 3089 cm−1. The N–H vibrations are observed at 3236, 3200, 3127, 3067, and 3022 cm−1.
The bands at 1769, 1713, and 1679 cm−1 in the Raman spectrum are assigned to the combination of the (ν2) NH4 and H–O–H deformation vibration of H2O molecules. The (ν4) NH4 stretching vibration is demonstrated by the 1431 cm−1 band. The antisymmetric stretching vibration, ν3, of the SO4 group is outlined by the 1139, 1093, 1071, and 1062 cm−1 bands. The strongest band in the Raman spectrum at 983 cm−1 is attributed to the ν1 symmetric stretching vibration of the SO4 group, and the band at 981 cm−1 at the infrared spectrum is also intensive, suggesting the symmetry reduction of sulfate group that may be considered as distorted tetrahedra. Medium-intensity bands at wavenumbers 625 and 615 cm−1 (Raman) and 626 and 616 cm−1 (infrared) are attributed to the four bending vibrations of the SO4 group. A medium-intensity band at 454 cm−1 in the Raman spectrum is attributed to the ν2 bending vibration of the SO4 group. The broad bands at wavenumbers 787, 724, 676, and 564 cm−1 can be attributed to the torsion vibrations of H2O molecules. The Raman bands at 360 and 310 cm−1 can be attributed to the vibrations of H2O molecules, while the 222 cm−1 band and lower wavenumbers are attributed to the lattice modes. In general, the band assignment is in agreement with that of [20,24]; however, the detailed assignment of the 3600–2500 cm−1 region is carried out for the first time.
Temperature-dependent Raman Spectroscopy. Three heating cycles from −190 to 500 °C were performed using three different boussingaultite crystals. Two of the three cycles are characterized by the following sequence of transformations (Figure 5):
(NH4)2Mg(SO4)2·6H2O (−190–85 °C) →
(NH4)2Mg(SO4)2·6H2O + amorphous phase (85 °C) →
amorphous phase (86–150 °C) →
(NH4)2Mg(SO4)2 (from ~160 °C).
It should be noted that MgSO4 is also likely observed, the band of which is clearly visible at ~1050 cm−1 (assigned to ν1(SO4) in accord with [43]). But the observation of the band at higher temperatures is problematic due to the strong photoluminescent background.
The third temperature cycle of measurements showed some differences in the Raman spectra (Figure 6). The intensive Raman band at 981 cm−1 is observed for boussingaultite and assigned to sulfate anion vibrations (Table 6). In the third heating cycle at temperature 90 °C, an additional band occurs at 996 cm−1, and only this band is presented at the temperatures 95 and 110 °C, while the 981 cm−1 band disappeared at these temperatures (Figure 6b,c). However, the position of the boussingaultite mode ν1(SO4) is practically independent of temperature and varies from ~981 to ~983 cm−1 in the temperature range from −190 to 90 °C (Figure 6c), which also shows the absence of SO4-tetrahedra ordering in this temperature range, in accordance with X-ray diffraction single-crystal analyses data. Another interesting observation is that, in the third cycle, the “intermediate” (partly dehydrated) phase was detected that was not revealed by other methods and two other temperature-dependent Raman cycles. Partial dehydration is proven by the presence of H2O vibrations in the region of 3600–3000 cm−1 at 95 °C. The phase was identified only by Raman spectroscopy in heating cycles with steps of 1–2 °C; thus, this phase has a very narrow temperature-stability field. This “intermediate” phase was identified as the ammonium analogue of leonite, i.e., phase with the composition (NH4)2Mg(SO4)2·4H2O. It is suggested that thermal decomposition of boussingaultite may pass through an intermediate stage, accompanied by the removal of a certain number of H2O molecules, like in gypsum [44]; however, it has not been detected by other methods.

5. Discussion

Boussingaultite is stable up to 90 °C. In the temperature range of 100–150 °C, it transforms into the X-ray amorphous phase due to the dehydration process. Efremovite crystallizes from the X-ray amorphous phase at ~160 °C; the compound is stable up to 340 °C, and, after that, two MgSO4 modifications (Cmcm and Pbnm) form due to the deammonization process in the temperature range from 340 to 580 °C, and the single Pbnm modification of MgSO4 is stable above 580 °C. A partly dehydrated form of boussingaultite, possibly “ammonioleonite”, was detected by Raman spectroscopy only in a narrow temperature range and slow heating rate. It is noteworthy that the technogenic “ammonioleonite”, (NH4)2Mg(SO4)2·4H2O (i.e., the chemical analogue of boussingaultite with 4 H2O pfu instead of 6), has been reported from CCB in association with boussingaultite and efremovite [31]. Based on our experimental data and conditions of technogenic phase formation, we may suggest that “ammonioleonite” may indeed form easily via the hydration of efremovite because it was almost undetectable in heating experiments of boussingaultite. The phase transformation of boussingaultite → ammonioleonite → efremovite → MgSO4 is followed by the following crystal chemical changes (accompanying the processes of dehydration and deammonization until their completion): isolated structure → layers from Mg(SO4)2(H2O)4 clusters interconnected by the NH4+ → tetrahedral–octahedral framework with NH4+ in the framework voids → dense structure of edge-sharing MgO6-octahedra chains interconnected via SO4 tetrahedra. The formation of a dehydrated chemical analogue of a phase may pass through intermediate forms. As an example, novograblenovite from CCB, (NH4)MgCl3·6H2O, which, at a temperature of about 90 °C, transforms to the crystalline partly dehydrated phase, (NH4)MgCl3·2H2O that, in turn, is stable up to 150 °C [18]. Meanwhile, contrasting behavior is observed for another ammonium sulfate, tschermigite, (NH4)Al(SO4)2·12H2O, which melts (at 60–70 °C) with the formation of amorphous phase and further crystallization (at 210 °C) of its anhydrous counterpart godovikovite, (NH4)Al(SO4)2, with its later transformation (at 380–390 °C) to millosevichite, Al2(SO4)3 [40].
Previously, the phase transition of boussingaultite has been suggested based on differential scanning calorimetry (DSC), dielectric permittivity, and heat-capacity curves that depended on crystal size [26]. The crystal chemical proposal was based on the assumption that the phase transition is associated with the disorder of sulfate tetrahedra. Neither the phase transition nor disordering of sulfate tetrahedra is observed in our study. However, we, for the first time, obtained signatures of existence of a crystalline partly dehydrated difficult-to-detect phase identified as “ammonioleonite”.
Thermal expansion of boussingaultite is strongly anisotropic and changes dramatically upon heating, which is reflected by the variation in character of thermal expansion of c and β unit-cell parameters (Figure 2). Thus, the crystal structure of boussingaultite experiences compression along the c-axis in the temperature range from −173 to −73 °C, while the a, b, and V parameters are increasing. However, in the temperature range from 25 to 90 °C, a, b, c, and V increase, while the monoclinic angle, β, slightly decreases. At the same time, expansion along the c-axis is weak (minimal).
It is worth noting that the c-axis direction is perpendicular to the plane of the pseudolayers, allowing us to assume that the thermal-expansion anisotropy of boussingaultite is controlled by its structural motif. Moreover, the direction of maximal thermal expansion (the b-axis) is parallel to the direction of the corrugation of pseudolayers, leading us to analyze the change in corrugation angle with temperature (Figure 7). Thus, the corrugation angle was calculated for the LT SCXRD data, using Olex2 software [34], as 180 − n, where n = normal to the angle between two symmetrical S–Mg–S planes. As a result, it was revealed that the corrugated layer straightened (with increasing temperature) from 92.391(5) to 98.240(9)°, causing the maximal thermal expansion along the b-axis.
The vibrational spectroscopy data allowed for the distinction of N–H bands to be carried out for the first time (given in detail in Table 5 and Table 6 and visualized in Figure 5 and Figure 8). This can be helpful for the identification of ammonium minerals. Also, according to vibrational spectroscopy, the heating leads to the oxidation of Fe2+ impurity to Fe3+. It is registered by the increasing luminescence signal in the Raman spectra of the heated samples. The mechanism of Fe oxidation is accompanied by the deprotonation of H2O molecules (Figure 8). Figure 8 also demonstrates dehydration and deammonization steps of boussingaultite based on the N–H and O–H bands’ intensity dependences on temperature: dehydration occurs in the temperature range from 100 to 200 °C, while deammonization occurs from 200 to 400 °C, which is, generally, in accordance with HTXRD data.

6. Conclusions

The high-temperature behavior of boussingaultite has been studied on the sample of technogenic origin from the burnt-coal dumps. Upon heating, boussingaultite transforms to efremovite, which, at a higher temperature, transforms to MgSO4, according to X-ray diffraction and vibrational spectroscopy data. The shift of ν1(SO4) mode from 981 to 996 cm−1 was observed at the temperature range of 95–110 °C in one of three Raman heating cycles, where the temperature step was 1–2 °C. This has been interpreted as the formation of a crystalline partly dehydrated modification that may correspond to “ammonioleonite”. The thermal expansion of boussingaultite is different for low (from −173 to −73 °C) and high (from 25 to 90 °C) regions. In both regions, the thermal behavior of boussingaultite is anisotropic, with the maximal thermal expansion parallel to the b-axis. The structure analysis shows that the strongest expansion is due to the straightening of corrugated pseudolayers with increasing temperature, while the minimal expansion (or even contraction in the low-temperature interval) is perpendicular to pseudolayers. Thus, anisotropy of thermal expansion of boussingaultite is mostly controlled by its layered structural character. We did not detect either the phase transition or disordering of sulfate tetrahedra suggested for synthetic analogue of boussingaultite previously. The study of boussingaultite and its dehydrated counterpart efremovite by infrared and Raman spectroscopy allowed for the separation of N–H and O–H vibrations that could be helpful for the diagnostic of ammonium minerals, including those studied by rovers.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/min14101052/s1, Table S1. Atomic coordinates and equivalent isotropic displacement parameters (Å2) of boussingaultite at T = −173 °C; Table S2. Anisotropic displacement parameters (Å2) of boussingaultite at T = −173 °C; Table S3. Selected bond lengths (Å) in the crystal structure of boussingaultite collected at T = −173 °C; Table S4. Hydrogen bonding scheme for boussingaultite at T = −173 °C; Table S5. Crystal data and structure refinement of boussingaultite from Kladno at T = 20 °C. Table S6. Atomic coordinates and equivalent isotropic displacement parameters (Å2) of boussingaultite from Kladno at T = 20 °C; Table S7. Anisotropic displacement parameters (Å2) of boussingaultite from Kladno at T = 20 °C; Table S8. Selected bond lengths (Å) in the crystal structure of boussingaultite from Kladno at T = 20 °C; Table S9. Hydrogen bonds of boussingaultite from Kladno at T = 20 °C; Table S10. Unit-cell parameters of boussingaultite in a wide range of temperatures: LT SCXRD data from −173 to −73 °C, SCXRD data for boussingaultite at 20 °C, and HTXRD data from 30 to 90 °C; Figure S1. The powder X-ray diffraction pattern calculated from single-crystal X-ray diffraction data (CuKα radiation) collected at different temperatures; Figure S2. The experimental powder X-ray diffraction pattern of boussingaultite in the temperature range from 25 to 90 °C; Figure S3. The Raman spectrum of boussingaultite measured at a temperature of −190 °C (1), the total curve of decomposition into individual peaks of the boussingaultite spectrum (2), bands associated with N–H vibrations (3), and bands associated with O–H vibrations (4).

Author Contributions

Conceptualization, E.S.Z. and A.A.Z. (Andrey A. Zolotarev); methodology, E.S.Z., R.M.S., A.A.Z. (Andrey A. Zolotarev), R.Y.S. and E.A.P.; software, R.M.S., K.A.T. and N.S.V.; validation, M.S.A., M.G.K. and A.A.Z. (Anatoly A. Zolotarev); formal analysis, R.M.S., R.Y.S., E.A.P., K.A.T., M.S.A. and N.S.V.; investigation, E.S.Z., R.M.S., A.A.Z. (Andrey A. Zolotarev), E.A.P., R.Y.S., K.A.T., M.S.A., M.G.K., N.S.V., A.A.Z. (Anatoly A. Zolotarev), M.A.R. and S.V.K.; resources, A.A.Z. (Andrey A. Zolotarev), M.A.R. and S.V.K.; data curation, E.S.Z., A.A.Z.( Andrey A. Zolotarev) and S.V.K.; writing—original draft preparation, E.S.Z., R.M.S., A.A.Z. (Andrey A. Zolotarev), E.A.P. and R.Y.S.; writing—review and editing, E.S.Z., R.M.S., A.A.Z. (Andrey A. Zolotarev), E.A.P., R.Y.S., K.A.T., M.S.A., M.G.K., N.S.V., A.A.Z. (Anatoly A. Zolotarev), M.A.R. and S.V.K.; visualization, E.S.Z., R.M.S., R.Y.S. and E.A.P.; supervision, A.A.Z. (Andrey A. Zolotarev) and S.V.K.; project administration, A.A.Z. (Andrey A. Zolotarev); funding acquisition, A.A.Z. (Andrey A. Zolotarev). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, grant number 23–27–00147, https://rscf.ru/en/project/23-27-00147/ (accessed on 2 September 2024).

Data Availability Statement

The Raman spectra of boussingaultite and intermediate ammonium leonite phases were deposited to raman-fmm database located at https://ardi.fmm.ru (accessed on 2 September 2024). The crystal-structure data for boussingaultite at T = −173, −148, −123, −98, −73, and 52 °C are available as CIF–files from the CCDC/FIZ Karlsruhe database as CSD No. 2372190, 2372191, 2372192, 2372193, 2372194, and 2372198, respectively, and CIF-file for sample from Kladno is registered as CSD No. 2376873, all at https://www.ccdc.cam.ac.uk (accessed on 2 September 2024).

Acknowledgments

The X-ray diffraction studies were performed in the X-ray Diffraction Resource Centre of St. Petersburg State University. The chemical analytical studies were performed in the “Geomodel” Resource Centre of St. Petersburg State University. Raman spectroscopy was carried out at the UB RAS “Geoanalitik” Centre for Collective Use. The infrared spectroscopic measurements were performed in the Isotope-geochemical research center for Collective Use (A. P. Vinogradov Institute of Geochemistry of the Siberian Branch of the Russian Academy of Sciences).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bosi, F.; Belardi, G.; Ballirano, P. Structural features in Tutton’s salts K2[M2+(H2O)6](SO4)2, with M2+ = Mg, Fe, Co, Ni, Cu, and Zn. Am. Mineral. 2009, 94, 74–82. [Google Scholar] [CrossRef]
  2. Nalbandyan, V.B. Thallium manganese sulfate hexahydrate, a missing Tutton’s salt, and a brief review of the entire family. Powder Diffr. 2008, 23, 52–55. [Google Scholar] [CrossRef]
  3. Shimobayashi, N.; Ohnishi, M.; Miura, H. Ammonium sulfate minerals from Mikasa, Hokkaido, Japan: Boussingaultite, godovikovite, efremovite and tschermigite. J. Mineral. Petrol. Sci. 2011, 106, 158–163. [Google Scholar] [CrossRef]
  4. Bechi, E. Lavori eseguiti nel laboratorio di Chimica del R. Istituto Tecnico di Forenze. Continuazione degli Atti Della R. Accad. Econ.–Agrar. Georg. Firenze 1863, 10, 201–206. [Google Scholar]
  5. Gatta, G.D.; Guastella, G.; Guastoni, A.; Gagliardi, V.; Cañadillas–Delgado, L.; Fernandez-Diaz, M.T. A neutron diffraction study of boussingaultite, (NH4)2Mg(H2O)6(SO4)2. Am. Mineral. 2023, 108, 354–361. [Google Scholar] [CrossRef]
  6. Garavelli, A.; Grasso, M.F.; Vurro, F. Mineral occurrence and depositional processes at Baia di Levante area (Vulcano Island, Italy). Mineral. Petrogr. Acta 1996, 39, 251–261. [Google Scholar]
  7. Ciesielczuk, J.; Żaba, J.; Bzowska, G.; Gaidzik, K.; Głogowska, M. Sulphate efflorescences at the geyser near Pinchollo, southern Peru. J. S. Am. Earth Sci. 2013, 42, 186–193. [Google Scholar] [CrossRef]
  8. Balić–Žunić, T.; Garavelli, A.; Jakobsson, S.P.; Jonasson, K.; Katerinopoulos, A.; Kyriakopoulos, K.; Acquafredda, P. Fumarolic minerals: An overview of active European volcanoes. In Updates in Volcanology—From Volcano Modelling to Volcano Geology; InTech: London, UK, 2016; pp. 267–322. [Google Scholar]
  9. Zhitova, E.S.; Siidra, O.I.; Shilovskikh, V.V.; Belakovsky, D.I.; Nuzhdaev, A.A.; Ismagilova, R.M. Ammoniovoltaite, (NH4)2Fe2+5Fe3+3Al(SO4)12(H2O)18, a new mineral from the Severo–Kambalny geothermal field, Kamchatka, Russia. Mineral. Mag. 2018, 82, 1057–1077. [Google Scholar] [CrossRef]
  10. Chesnokov, B.; Kotrly, M.; Nisanbajev, T. Brennende Abraumhalden und Aufschlüsse im Tscheljabinsker Kohlenbecken-eine reiche Mineralienküche. Mineralien-Welt 1998, 9, 54–63. [Google Scholar]
  11. Parafiniuk, J.; Kruszewski, Ł. Ammonium minerals from burning coal–dumps of the Upper Silesian Coal Basin (Poland). Geol. Quart. 2009, 53, 341–356. [Google Scholar]
  12. Kruszewski, Ł. Secondary sulphate minerals from Bhanine Valley coals (South Lebanon): A crystallochemical and geochemical study. Geol. Quart. 2019, 63, 65–87. [Google Scholar] [CrossRef]
  13. Mysen, B. Nitrogen in the Earth: Abundance and transport. Prog. Earth Planet. Sci. 2019, 6, 38. [Google Scholar] [CrossRef]
  14. Nachlas, W.O.; Bushmaker, S.; Sasson, E. X-ray Spectroscopy of Nitrogen in Jarosite, Ammoniojarosite, and other NH4-Bearing Sulfate Minerals. Microsc. Microanal. 2023, 29, 862–864. [Google Scholar] [CrossRef]
  15. Britvin, S.N.; Vereshchagin, O.; Vlasenko, N.; Krzhizhanovskaya, M.; Ivanova, M. Meteoritic Tutton salt, a naturally inspired reservoir of cometary and asteroidal ammonium. arXiv 2024, arXiv:2407.20997. [Google Scholar]
  16. Todd, Z.R. Sources of Nitrogen–, Sulfur–, and Phosphorus–containing feedstocks for prebiotic chemistry in the planetary environment. Life 2022, 12, 1268. [Google Scholar] [CrossRef]
  17. Johnson, B.; Goldblatt, C. The nitrogen budget of Earth. Earth Sci. Res. 2015, 148, 150–173. [Google Scholar] [CrossRef]
  18. Zolotarev, A.A.; Zhitova, E.S.; Krzhizhanovskaya, M.G.; Rassomakhin, M.A.; Shilovskikh, V.V.; Krivovichev, S.V. Crystal chemistry and high–temperature behaviour of ammonium phases NH4MgCl3·6H2O and (NH4)2Fe3+Cl5·H2O from the burned dumps of the Chelyabinsk Coal Basin. Minerals 2019, 9, 486. [Google Scholar] [CrossRef]
  19. Stern, J.C.; Sutter, B.; Freissinet, C.; Navarro–González, R.; McKay, C.P.; Archer, P.D., Jr.; Buch, A.; Brunner, A.E.; Coll, P.; Eigenbrode, J.L.; et al. Evidence for indigenous nitrogen in sedimentary and aeolian deposits from the Curiosity rover investigations at Gale crater, Mars. Proc. Natl. Acad. Sci. USA 2015, 112, 4245–4250. [Google Scholar] [CrossRef]
  20. Culka, A.; Jehlička, J.; Němec, I. Raman and infrared spectroscopic study of boussingaultite and nickelboussingaultite. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2009, 73, 420–423. [Google Scholar] [CrossRef]
  21. Sergeeva, A.V.; Zhitova, E.S.; Bocharov, V.N. Infrared and Raman spectroscopy of tschermigite, (NH4)Al(SO4)2·12H2O. Vib. Spectrosc. 2019, 105, 102983. [Google Scholar] [CrossRef]
  22. Sergeeva, A.V.; Zhitova, E.S.; Nuzhdaev, A.A.; Zolotarev, A.A.; Bocharov, V.N.; Ismagilova, R.M. Infrared and Raman spectroscopy of ammoniovoltaite, (NH4)2Fe2+5Fe3+3Al(SO4)12·(H2O)18. Minerals 2020, 10, 781. [Google Scholar] [CrossRef]
  23. Culka, A.; Košek, F.; Drahota, P.; Jehlička, J. Use of miniaturized Raman spectrometer for detection of sulfates of different hydration states–Significance for Mars studies. Icarus 2014, 243, 440–453. [Google Scholar] [CrossRef]
  24. Košek, F.; Culka, A.; Jehlička, J. Field identification of minerals at burning coal dumps using miniature Raman spectrometers. J. Raman Spectrosc. 2017, 48, 1494–1502. [Google Scholar]
  25. Košek, F.; Culka, A.; Rousaki, A.; Vandenabeele, P.; Jehlička, J. Raman spectroscopy of anhydrous and hydrated aluminum sulfates: Experience from burning coal heaps. J. Raman Spectrosc. 2022, 53, 1959–1973. [Google Scholar] [CrossRef]
  26. Kosova, D.A.; Druzhinina, A.I.; Tiflova, L.A.; Monayenkova, A.S.; Uspenskaya, I.A. Thermodynamic properties of ammonium magnesium sulfate hexahydrate (NH4)2Mg(SO4)2·6H2O. J. Chem. Thermodyn. 2018, 118, 206–214. [Google Scholar] [CrossRef]
  27. Shcherbakova, Y.P.; Bazhenova, L.F.; Chesnokov, B.V. Godovikovite—NH4(Al,Fe)(SO4)2 a new ammonium–bearing sulfate. Zap. RMO (Proc. Russ. Mineral. Soc.) 1988, 117, 208–211. (In Russian) [Google Scholar]
  28. Shcherbakova, Y.P.; Bazhenova, L.F. Efremovite (NH4)2Mg2(SO4)3—Ammonium analogue of langbeinite—A new mineral. Zap. RMO (Proc. Russ. Mineral. Soc.) 1989, 118, 84–87. (In Russian) [Google Scholar]
  29. Chesnokov, B.V.; Shcherbakova, E.P. Mineralogy of the Burnt Dumps of the Chelyabinsk Coal Basin (an Experience in Technogenic Mineralogy); Nauka: Moscow, Russia, 1991; pp. 1–152. (In Russian) [Google Scholar]
  30. Shcherbakova, Y.P. Ammonium-Containing Sulfates from the Kopeisk Coal Basin. New and Poorly Studied Minerals and Mineral Associations of the Urals; UC USSR Academy of Sciences: Sverdlovsk, Russia, 1986; pp. 209–211. (In Russian) [Google Scholar]
  31. Chesnokov, B.V.; Shcherbakova, E.P.; Nishanbaev, T.P. Minerals of Burnt Dumps of the Chelyabinsk Coal Basin; Ural Branch of RAS: Miass, Russia, 2008; pp. 1–139. (In Russian) [Google Scholar]
  32. CrysAlisPro Software System, version 1.171.39.44; Rigaku Oxford Diffraction: Oxford, UK, 2015.
  33. Sheldrick, G.M. Crystal Structure Refinement with It SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  34. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. It OLEX2: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  35. Momma, K.; Izumi, F. VESTA 3 for three–dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  36. BRUKER-AXS Topas V4.2: General Profile and Structure Analysis Software for Powder Diffraction Data; Bruker-AXS: Karlsruhe, Germany, 2009.
  37. Bubnova, R.S.; Firsova, V.A.; Filatov, S.K. Software for determining the thermal expansion tensor and the graphic representation of its characteristic surface (theta to tensor-TTT). Glass Phys. Chem. 2013, 39, 347–350. [Google Scholar] [CrossRef]
  38. Shendrik, R.Y.; Plechov, P.Y.; Smirnov, S.Z. ArDI—The system of mineral vibrational spectroscopy data processing and analysis. New Data Miner. 2024, 58, 26–35. (In Russian) [Google Scholar]
  39. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds: Part A: Theory and Applications in Inorganic Chemistry; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2009; pp. 1–16. [Google Scholar]
  40. Zhitova, E.S.; Sergeeva, A.V.; Nuzhdaev, A.A.; Krzhizhanovskaya, M.G.; Chubarov, V.M. Tschermigite from thermal fields of Southern Kamchatka: High-temperature transformation and peculiarities of IR-spectrum. Zap. RMO (Proc. Russ. Mineral. Soc.) 2019, 148, 100–116. [Google Scholar]
  41. Smith, D.H.; Seshadri, K.S. Infrared spectra of Mg2Ca(SO4)3, MgSO4, hexagonal CaSO4, and orthorhombic CaSO4. Spectrochim. Acta A Mol. Biomol. Spectrosc. 1999, 55, 795–805. [Google Scholar] [CrossRef]
  42. Kroumova, E.; Aroyo, M.I.; Perez-Mato, J.M.; Kirov, A.; Capillas, C.; Ivantchev, S.; Wondratschek, H. Bilbao crystallographic server: Useful databases and tools for phase-transition studies. Phase Transit. 2003, 76, 155–170. [Google Scholar] [CrossRef]
  43. Wang, F.; Shou, J.; Zhang, Y. Micro–Raman investigations on the structures of the surface and the inner of MgSO4 droplets. Chin. Sci. Bull. 2008, 53, 2414–2416. [Google Scholar] [CrossRef]
  44. Pankrushina, E.A.; Votyakov, S.L.; Shchapova, Y.V. Statistical approaches in the analysis of in situ thermo-Raman spectroscopic data for gypsum as a basis for studying dehydration and phase transformations in crystalline hydrates. J. Raman Spectrosc. 2021, 52, 877–889. [Google Scholar] [CrossRef]
Figure 1. High-temperature X-ray diffraction (HTXRD) patterns of boussingaultite and its high-temperature derivatives: (I) boussingaultite (NH4)2Mg(SO4)2·6H2O, 25–90 °C; (II) X-ray amorphous phase, 100–150 °C; (III) efremovite (NH4)2Mg2(SO4)3, 160–340 °C; (IV) MgSO4 (Cmcm), 340–580 °C; and (V) MgSO4 (Pbnm), 340–700 °C.
Figure 1. High-temperature X-ray diffraction (HTXRD) patterns of boussingaultite and its high-temperature derivatives: (I) boussingaultite (NH4)2Mg(SO4)2·6H2O, 25–90 °C; (II) X-ray amorphous phase, 100–150 °C; (III) efremovite (NH4)2Mg2(SO4)3, 160–340 °C; (IV) MgSO4 (Cmcm), 340–580 °C; and (V) MgSO4 (Pbnm), 340–700 °C.
Minerals 14 01052 g001
Figure 2. The temperature dependencies of the unit cell parameters of boussingaultite. The approximation equations are a = 9.298 + 0.47 × T × 10−3 (from −173 to 90 °C); b = 12.570 + 0.85 × T × 10−3 (from −173 to 90 °C); c = 6.172 − 0.51 × T × 10−3 (from −173 to −73 °C); c = 6.203 + 0.10 × T × 10−3 (from 25 to 90 °C); β = 107.22 + 1.2 × T × 10−3 (from −173 to −73 °C); β = 107.120 − 0.47 × T × 10−3 (from 25 to 90 °C); and V = 694.6 + 62 × T × 10−3 (from −173 to 90 °C) (for details, see Supplementary Table S10).
Figure 2. The temperature dependencies of the unit cell parameters of boussingaultite. The approximation equations are a = 9.298 + 0.47 × T × 10−3 (from −173 to 90 °C); b = 12.570 + 0.85 × T × 10−3 (from −173 to 90 °C); c = 6.172 − 0.51 × T × 10−3 (from −173 to −73 °C); c = 6.203 + 0.10 × T × 10−3 (from 25 to 90 °C); β = 107.22 + 1.2 × T × 10−3 (from −173 to −73 °C); β = 107.120 − 0.47 × T × 10−3 (from 25 to 90 °C); and V = 694.6 + 62 × T × 10−3 (from −173 to 90 °C) (for details, see Supplementary Table S10).
Minerals 14 01052 g002
Figure 3. The crystal structure of boussingaultite (a) and the figures of thermal-expansion coefficients at 60 °C (b) and −123 °C (c). * Angle corresponds to corrugated angle (see text).
Figure 3. The crystal structure of boussingaultite (a) and the figures of thermal-expansion coefficients at 60 °C (b) and −123 °C (c). * Angle corresponds to corrugated angle (see text).
Minerals 14 01052 g003
Figure 4. Infrared absorption spectra of boussingaultite and its high-temperature derivatives (efremovite and MgSO4) obtained by ex situ heat treatment. (a) The whole spectra and (b) the stretching-mode region of NH4- and H2O-related vibrations are shown. Blue bands are related to O–H stretching, and green bands are related to N–H stretching modes.
Figure 4. Infrared absorption spectra of boussingaultite and its high-temperature derivatives (efremovite and MgSO4) obtained by ex situ heat treatment. (a) The whole spectra and (b) the stretching-mode region of NH4- and H2O-related vibrations are shown. Blue bands are related to O–H stretching, and green bands are related to N–H stretching modes.
Minerals 14 01052 g004
Figure 8. The temperature dependences of integral intensities of bands relayed to NH4 bending (curve 1) and O–H stretching bands (curve 2). Curve 3 is the intensity of luminescence registered in Raman spectra of the heated sample.
Figure 8. The temperature dependences of integral intensities of bands relayed to NH4 bending (curve 1) and O–H stretching bands (curve 2). Curve 3 is the intensity of luminescence registered in Raman spectra of the heated sample.
Minerals 14 01052 g008
Table 1. Chemical composition of boussingaultite (in wt. %).
Table 1. Chemical composition of boussingaultite (in wt. %).
ConstituentMeanRangeAtoms per Formula Unit for S = 2Probe Standard
(NH4)2Omeas (1)12.9011.97–15.391.77BN (N)
[(NH4)2Ocalc] (2)14.552.00
MgO11.4711.29–12.641.02MgO (Mg)
MnO0.260.19–0.370.01Mn (Mn)
FeO0.300.27–0.340.01FeS2 (Fe)
SO344.8544.64–45.952.00FeS2 (S)
H2O (3)30.276.00
Total100.05
[101.70]
(1) Nitrogen (ammonium) content can be taken as approximate due to the semi-quantitative element content with Z < 8, using energy-dispersive spectroscopy; (2) calculated from the crystal-structure data; (3) calculated by stoichiometry.
Table 2. Crystal data and structure refinement of boussingaultite collected at T = −173 °C (or 100 K) and 52 °C (or 325 K).
Table 2. Crystal data and structure refinement of boussingaultite collected at T = −173 °C (or 100 K) and 52 °C (or 325 K).
Crystal SystemMonoclinic
Space groupP21/a
Temperature, °C–17352
a, Å9.1992(5)9.2964(12)
b, Å12.4049(7)12.5825(9)
c, Å6.2590(3)6.1942(7)
β, °107.026(6)107.106(13)
Volume, Å3682.94(7)692.50(14)
Z2
ρcalc, g/cm31.7541.883
μ, mm−10.5080.520
F(000)380.0412.0
RadiationMoKα (λ = 0.71073)
2Θ range for data collection, °from 6.57 to 65.228from 7.608 to 65.046
Index ranges−13 ≤ h ≤ 8,
−16 ≤ k ≤ 18,
−19 ≤ l ≤ 9
−13 ≤ h ≤ 11,
−18 ≤ k ≤ 17,
−8 ≤ l ≤ 8
Reflections collected47724880
Independent reflections [Rint, Rsigma]2227
[Rint = 0.0362, Rsigma = 0.0548]
2162
[Rint = 0.0700, Rsigma = 0.1102]
Data/restraints/parameters2227/0/1282162/0/128
Goodness-of-fit on F21.0821.053
Final R indexes [I ≥ 2σ(I)]R1 = 0.0392, wR2 = 0.0824R1 = 0.0626, wR2 = 0.1232
Final R indexes [all data]R1 = 0.0556, wR2 = 0.0920R1 = 0.1378, wR2 = 0.1794
Largest diff. peak/hole/e Å−30.55/−0.610.48/−0.69
Table 3. Geometric parameters of boussingaultite structure at T = −173 (or 100 K) and 52 °C (or 325 K).
Table 3. Geometric parameters of boussingaultite structure at T = −173 (or 100 K) and 52 °C (or 325 K).
Parameter−173 °C52 °C
<Mg–O>, Å2.0682.063
<S–O>, Å1.4791.468
<N–H>, Å0.8870.852
VMg, Å311.7811.70
Vsulf, Å31.661.62
Vamm, Å30.360.31
Distortion index for Mg(H2O)60.00570.0066
Distortion index for SO40.00590.0055
Distortion index for NH40.03570.0640
Table 4. Thermal-expansion coefficients, ×106 °C−1 for boussingaultite.
Table 4. Thermal-expansion coefficients, ×106 °C−1 for boussingaultite.
T, °Cα11α22α33αV
–12352(2)68(2)–89(3)31(3)
6053(2)67(2)15(1)136(3)
Table 5. IR absorption bands of boussingaultite and its high-temperature derivatives, efremovite and magnesium sulfate.
Table 5. IR absorption bands of boussingaultite and its high-temperature derivatives, efremovite and magnesium sulfate.
IR Band Maxima (cm−1) at Temperatures (°C)Band Assignment
BoussingaultiteEfremoviteMgSO4
RT150200250300400
34643464 ↓----O–H stretching
33503350 ↓----O–H stretching
32723271326732663265 ↓-N–H stretching
32333233 ↓----O–H stretching
31893189318931883187 ↓-N–H stretching
31553155 ↓----O–H stretching
31243125312531273130 ↓-N–H stretching
30773075307530743071 ↓-N–H stretching
30403040 ↓----O–H stretching
30123012 ↓----O–H stretching
29802980298029802980 ↓-N–H stretching
29202922292029202920 ↓-N–H stretching
28392840284228442846 ↓-N–H stretching
16601660 ↓----H–O–H bending
14681468 ↓----H–N–H bending *
142314201418 ↓1418 ↓1414 ↓-H–N–H bending
----1167 w1167v3 (SO4)
114011401140114011401140v3 (SO4)
10601083
1125 w
112511251125-v3 (SO4)
981981
1040 w
1040104010401062v1 (SO4)
-----1012v1 (SO4)
-880880---H–O–H libration
--892---H–O–H libration
--657657657680
700
v4 (SO4)
↓ intensity reduced. Note: w—weak. * The band is splinted into two components (1468 and 1463 cm−1), one of the components disappears upon heating up to 150 °C. This may be due to either a change in the geometry of NH4 cation or due to the loss of perturbed H2O.
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

Zhitova, E.S.; Sheveleva, R.M.; Zolotarev, A.A.; Shendrik, R.Y.; Pankrushina, E.A.; Turovsky, K.A.; Avdontceva, M.S.; Krzhizhanovskaya, M.G.; Vlasenko, N.S.; Zolotarev, A.A.; et al. The Crystal Chemistry of Boussingaultite, (NH4)2Mg(SO4)2·6H2O, and Its Derivatives in a Wide Temperature Range. Minerals 2024, 14, 1052. https://doi.org/10.3390/min14101052

AMA Style

Zhitova ES, Sheveleva RM, Zolotarev AA, Shendrik RY, Pankrushina EA, Turovsky KA, Avdontceva MS, Krzhizhanovskaya MG, Vlasenko NS, Zolotarev AA, et al. The Crystal Chemistry of Boussingaultite, (NH4)2Mg(SO4)2·6H2O, and Its Derivatives in a Wide Temperature Range. Minerals. 2024; 14(10):1052. https://doi.org/10.3390/min14101052

Chicago/Turabian Style

Zhitova, Elena S., Rezeda M. Sheveleva, Andrey A. Zolotarev, Roman Yu. Shendrik, Elizaveta A. Pankrushina, Konstantin A. Turovsky, Margarita S. Avdontceva, Maria G. Krzhizhanovskaya, Natalia S. Vlasenko, Anatoly A. Zolotarev, and et al. 2024. "The Crystal Chemistry of Boussingaultite, (NH4)2Mg(SO4)2·6H2O, and Its Derivatives in a Wide Temperature Range" Minerals 14, no. 10: 1052. https://doi.org/10.3390/min14101052

APA Style

Zhitova, E. S., Sheveleva, R. M., Zolotarev, A. A., Shendrik, R. Y., Pankrushina, E. A., Turovsky, K. A., Avdontceva, M. S., Krzhizhanovskaya, M. G., Vlasenko, N. S., Zolotarev, A. A., Rassomakhin, M. A., & Krivovichev, S. V. (2024). The Crystal Chemistry of Boussingaultite, (NH4)2Mg(SO4)2·6H2O, and Its Derivatives in a Wide Temperature Range. Minerals, 14(10), 1052. https://doi.org/10.3390/min14101052

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