Next Article in Journal
Syn-Sedimentary Exhalative or Diagenetic Replacement? Multi-Proxy Evidence for Origin of Metamorphosed Stratiform Barite–Sulfide Deposits near Aberfeldy, Scottish Highlands
Previous Article in Journal
Source Rock Assessment of the Permian to Jurassic Strata in the Northern Highlands, Northwestern Jordan: Insights from Organic Geochemistry and 1D Basin Modeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Crystal Chemistry of Synthetic Mg(Si1−xGex)O3 Pyroxenes: A Single-Crystal X-ray Diffraction Study

by
Günther J. Redhammer
* and
Gerold Tippelt
Department Chemistry & Physics of Materials, University of Salzburg, Jakob Haringer Str. 2A, 5020 Salzburg, Austria
*
Author to whom correspondence should be addressed.
Minerals 2024, 14(9), 864; https://doi.org/10.3390/min14090864
Submission received: 18 July 2024 / Revised: 12 August 2024 / Accepted: 17 August 2024 / Published: 25 August 2024

Abstract

:
Germanate-pyroxenes often are used as model systems to study the stability and phase relationships of analog silicate systems. Based on such analyses, it is assumed that silicates and germanates behave ideally in terms of mixing. A systematic study was performed to monitor in detail the changes introduced by a Si4+ through Ge4+ replacement in the important rock-forming pyroxene enstatite MgSiO3. Well-shaped, idiomorphic singe crystals of a MgSi1−xGexO3 pyroxene solid solution were grown at ambient pressure from a high-temperature flux-assisted synthesis. Structural analysis using single-crystal X-ray diffraction methods revealed orthorhombic symmetry, Pbca, Z = 8, for the complete solid-solution series. Long-term storage over a period of 8 years at ambient conditions or annealing at 525 °C over a period of 10 weeks did not change the symmetry of the proposed thermodynamically stable monoclinic polymorph. Within the solid-solution series, lattice parameters increased almost linearly with increasing Si4+ by Ge4+ substitution. The main changes occurred on the tetrahedral sites, which showed an almost linear increase in individual and average bond lengths but also in distortion parameters. The refined site occupancy of Si4+ and Ge4+ showed a distinct preference of Ge4+ for the TB site. The altered topology and kinking state in the tetrahedral chains also imposed significant changes to the bonding topology and geometry of the neighboring M1 and M2 sites.

1. Introduction

Pyroxenes are very abundant and important rock-forming silicates and, thus, are widely studied in geoscience. Pyroxenes of the (Mg,Fe)SiO3 series make up approximately 40% of the Earth’s upper mantle [1,2]. An understanding of the phase relationships and physical properties under the geological conditions of these phases is thus of general interest in geoscience. However, there is also a vivid interest in pyroxenes in crystal structure research because of their rich crystal chemistry with triclinic ( P 1 ¯ ), monoclinic (high-temperature and high-pressure C2/c structures, namely, P21/c, P21/n and P2/n) and orthorhombic (Pbca) polymorphs as a function of temperature (T), pressure (P) and chemical composition (X). The crystal structures of these polymorphs are closely related to each other. The main characteristics of the pyroxene structure are infinite zig-zag chains of octahedrally coordinated M1 sites with laterally attached five- to eight-fold coordinated M2 sites. Both sites form bands parallel to the c-axis, which themselves are stacked along the a-axis alternately with layer-like units built up by infinite single chains of tetrahedral sites. The tetrahedral chains are kinked to maintain the connection between M1 and octahedral sites and also run parallel to the c-axis. The stacking and orientation of the M1 site octahedra with respect to the silicate chain along the a-axis and the topology of the tetrahedral chains (one equivalent T or two symmetry non-equivalent TA and TB chains) yield different polymorphs with different space group symmetries. A detailed review of pyroxene topology and chemistry with the introduction of the often-used I-beam scheme was carried out by Cameron and Papike [2], and more information on the behavior of pyroxenes at high T and P levels can be found in [3]. Generalized phase relationships in the P-T field were proposed by Angel and Arlt [4]. The rarely found P 1 ¯ polymorph is the low-temperature form of synthetic NaTiSi2O6 [5], while the unusual P21/n structure is found so far only in synthetic LiAlGe2O6 [6].
Enstatite MgSiO3 is a typical example of a pyroxene, showing both monoclinic and orthorhombic polymorphs depending on formation (synthesis) conditions, with a total of five different polymorphs. In its low-temperature, low-pressure form, MgSiO3 adopts a monoclinic P21/c structure (clino-enstatite). At elevated temperatures, it transforms to the orthorhombic Pbca structure, while increasing pressure stabilizes the high-pressure C2/c structure for P > 6 GPa [7,8,9]. A similar effect was found by Ulmer and Stalder from Raman spectroscopic data on quenched samples [10]. Extrapolating the phase boundary between P21/c clino- and Pbca orthopyroxene to atmospheric pressure gives a transition temperature of ~550 °C [10]. At a high temperature of 1170 °C, ortho-enstatite transforms to proto-enstatite (high-temperature orthopyroxene) [11,12]. This high-temperature orthorhombic phase is non-quenchable and is only observable in in situ experiments. The equation of state and the phase relationships of the MgSiO3 pyroxene system at pressures up to 12 GPa were recently summarized by Sokolova et al. [13], confirming the generally accepted P-T phase diagram, with P21/c clino-enstatite being the thermodynamically stable phase at ambient pressure and temperature conditions.
Several studies, however, have shown that during single-crystal growth of MgSiO3 and using solvent-assisted synthesis with a lithiumvanadomolybdate as flux at ambient pressures, Pbca ortho-enstatite is the phase obtained instead of clino-enstatite, as would be expected from the phase relationships [9,11,14]. This indicates that the Pbca phase is quenchable and preserved during the (rapid) cooling from high temperatures after synthesis. Using this synthesis method, large, euhedral single crystals up to several millimeters in size were grown [9].
Many studies have also been performed on MgGeO3, as can be seen in a low-pressure model of the (Mg,Fe)SiO3 Earth crust- and mantle-forming system [15,16]. Like in MgSiO3, MgGeO3 pyroxene-type structures transform to ilmenite, to perovskite, and to post-perovskite structures but at lower and thus more accessible pressures [15]. Several studies were performed to determine the phase stability of different pyroxene modifications of MgGeO3 at pressures < 8 GPa [14,17,18], yielding partially contradicting results. Recently Hunt et al. [15] reinvestigated the phase relationships and confirmed the results of [17], with orthorhombic MgGeO3 being the stable phase at ambient pressures up to at least 1200 °C. Applying low pressures < 1 GPa, however, quickly stabilizes the clinopyroxene structure. This contradicts some findings of [19] that state that the stable, low-temperature form of MgGeO3 at ambient pressure is also monoclinic: space group C2/c. Their monoclinic MgGeO3 material was synthesized from flux-assisted crystal growth, quenched from 800 °C. Ozima and Akimoto [14] observed dominating orthopyroxene in their flux growth experiments but also, simultaneously, some monoclinic-phase MgGeO3 in a single run. Their cut-off temperature of synthesis was 650 °C. Welch and Pawley [20] performed synthesis of MgGeO3 using hydrothermal techniques at 650 °C and low pressures of 0.2 GPa, also yielding monoclinic C2/c clinopyroxenes.
Many silicate pyroxenes can only be synthesized by applying non-ambient pressures (>1 kbar), which makes experimental set-up more advanced. For such synthesis techniques, special equipment (hydrothermal synthesis in cold-sealed autoclaves or piston–cylinder apparatus for pressures < 4 kbar) is needed. Besides more sophisticated synthesis techniques, an additional disadvantage is the often very limited sample amount. This may hinder analyses requiring sample masses in the gram range, such as neutron diffraction, or multi-methodical approaches. Alternatively, it has been shown that Ge analogs of silicate-based minerals can often be used as good model systems to understand the mechanisms of structural phase transitions in minerals as a function of T and P [20,21,22,23,24]. As summarized in [20], the main assumption in investigations of Si4+ with Ge4+ substituted synthetic minerals is that they are thermodynamically and structurally representatives of their silicate counterparts. This holds true in the general sense and adds some additional benefits: (i) the synthesis of Ge analogs often can be performed at ambient pressures and (ii) phase transitions are shifted to experimentally more accessible (lower) temperature and pressure regions [25,26,27,28]. Studying binary Si-Ge series can offer reliable information on the ideality or non-ideality of such substitutions. This can finally prove the utility of Ge analogs as model systems for silicate mineral behavior.
Herein, the geoscientific, important solid-solution series MgSiO3-MgGeO3 was investigated structurally at room temperature to clarify the ideal (linear) or non-ideal behavior of mixing. Further aims are mainly the question of symmetry of the MgSiO3 end-member, how the substitution of the larger Ge4+ cation in place of Si4+ alters the structure and if there is random distribution of Ge4+ onto the two possible tetrahedral positions.

2. Materials and Methods

2.1. Synthesis

Single crystals of MgSi1−xGexO3 with 0.0 ≤ x ≤ 1.0 were synthesized using a high-temperature, flux-assisted growth technique, which was first applied by Ito [9] for the synthesis of ortho-enstatite MgSiO3 and later by Ozima [14] for the orthopyroxene MgGeO3. In the first step, reagent-grade MgO (98%, Merck, Darmstadt, Germany), quartz SiO2 (99.8%, Merck, Darmstadt, Germany) and GeO2 (99.998% metal basis, Aldrich, St.Louise, MO, USA) were weighed to the stoichiometry of the desired solid-solution composition and carefully mixed by grinding them in an agate mortar under alcohol for at least 20 min. The flux was the same as used by [9] and consisted of a mixture of 34.9 weight % MoO3 (99.5%, Sigma-Aldrich), 6.3 weight % V2O5 (>99.6%, Aldrich) and 21.9 weight % Li2CO3 (99.9%, Merck). The flux was also mixed carefully by grinding for 20 min. One large batch of flux was prepared and the same flux was used for all experiments. For the synthesis itself, the flux-to-nutrient ratio was 14:1 by weight. The mixture was put into a platinum crucible, covered with a platinum lid, heated to 960 °C at a rate of 5°/minute in a chamber furnace under ambient conditions (air, atmospheric pressure) and maintained at this temperature for 7 d for soaking. Then, the melt was cooled to 700 °C at a rate of 0.03°/min. After reaching the final temperature, the furnace was shut down and the synthesis batch furnace cooled.
After synthesis, the melt cake had a yellowish color. The flux was removed completely by soaking the synthesis batch in hot distilled water for several hours with intermediate washing cycles until only small, transparent, colorless pyroxene crystals were recovered. The final product was filtered, air dried and stored for further analysis.

2.2. Single-Crystal X-ray Diffraction (SCXRD)

For single-crystal X-ray diffraction, suitable crystals, selected on the basis of their optical properties (regular shape and homogeneity in color), were glued on top of glass capillaries (0.1 mm Ø). Single-crystal X-ray diffraction data were collected on a Bruker SMART APEX CCD-diffractometer. Intensity data were collected with graphite-monochromatized Mo Kα X-radiation (50 kV, 20 mA); the crystal-to-detector distance was 60 mm and the detector was positioned at −30° and −50° 2Θ using an ω-scan mode strategy at four different ϕ positions (0°, 90°, 180° and 270°) for each 2Θ position. In total, 630 frames with Δω = 0.3° were acquired for each run.
Three-dimensional data were integrated and corrected for Lorentz, polarization and background effects using the APEX3 software [29]. Structure solution (using direct methods) and subsequent weighted full-matrix least-squares refinements on F2 were performed with SHELX-2012 [30] as implemented in the program suite WinGX 2023.1 [31]. Structural drawings and calculations of polyhedral volume and distortion parameters were carried out using VESTA [32]. The definition of the distortion parameters can be found in Appendix A.

3. Results and Discussion

3.1. Synthesis and Morphology

It was possible to synthesize well-shaped, colorless crystals, partly reaching up several mm in size for all compositions. This especially holds true for Si-rich compositions. The obtained crystals are highly transparent without obvious defects inside, except for some cracks. Optical microscopic photographs of typical crystals of MgSiO3 and MgGeO3 crystals are displayed in Figure 1. For MgSiO3, the crystals are needle-like, columnar and they are elongated along the c-axis with well-developed faces. The largest needles reach up 5 mm in length, while most of the crystals are shorter and prismatic with sizes up to 0.5 mm. The crystals have an octagonal cross-section with well-developed {1 0 0}, {0 1 0} and {2 1 0} faces, and these columns are capped by {1 1 1} and {2 1 1} faces. Some large platelet-shaped crystals were also observed (up to 3 × 5 × 0.5 mm), with {1 0 0} and {0 1 0} faces dominating. In MgGeO3, crystals tend to be smaller and exhibit less perfect idiomorphic shapes; however, the habit is exactly the same as for the MgSiO3 end-member described above. No obvious coloring of the crystals was observed.

3.2. Single-Crystal X-ray Diffraction

A total of 26 datasets for crystals with 11 different nominal compositions were collected. Most of the data recordings were conducted directly after the synthesis experiments in early 2016. Additionally, some new datasets were collected recently for the Si and Ge end-member compositions to check for any changes in symmetry due to long storage times at ambient pressure and temperature over a period of ~8 years. Analysis of systematic absences of observed Bragg peaks and intensity statistics yields orthorhombic symmetry with space group Pbca for all crystals along the solid solution and for those in long-term storage. No change to the proposed thermodynamic stable monoclinic P21/c polymorphs was observed with time and the orthorhombic forms remained metastable when exposed to ambient pressure/temperature conditions.
Single crystals of pure MgSiO3 and MgGeO3 were also annealed at 400 °C over a period of 14 days. Quick data collections to check the unit cell parameters and symmetry of these annealed crystals showed them to have retained their original structure. Lattice parameters are identical within standard deviations to full data collections. In addition, the same material was further annealed at 525 °C for another 56 days and no change from orthorhombic to monoclinic symmetry was observed; diffraction peaks remained remarkably sharp with no indication of streaks or any hints of transformation.
Details on data collection and structure refinement for selected crystals are given in Table 1. Fractional atomic coordinates and equivalent isotropic atomic displacement parameters are given in Table 2. Table 3 reports selected bond lengths for the refinements given in Table 1. The full data for 10 representative compositions in this study, including anisotropic atomic displacement parameters, were deposited as a CIF (crystallographic information file) to the Cambridge Crystallographic Data Centre and can be retrieved under CSD Numbers # 2371508–# 2371517.
Data for pure enstatite MgSiO3 were used for model building. Structure solution and subsequent refinements yield a model with two positions for Mg and Si each and six positions for the oxygen atoms in agreement with published data. All atoms reside on general position 8c with site symmetry 1, so there are no symmetry-equivalent coordinating anions to the cations, nor are there symmetry-equivalent bonds. In the final model, fractional atomic coordinates were listed in accordance with the established convention [33,34,35]. Test refinements on end-member compositions with unconstrained site occupation factors for the M1, M2, TA and TB sites with the site occupation of the oxygen atoms were fixed to l (full occupation) yield values of ~1, indicating that all the sites show full occupation. Thus, in subsequent refinements on end-member and solid-solution compounds, the occupation of the tetrahedral sites TA and TB was constrained to a value of Si + Ge = 1, but the fraction and distribution of Si + Ge over the two T sites was allowed to refine freely. The scattering contrast is significantly different between these two elements (14 e vs. 32 e), so a refinement of the chemical composition of the investigated crystals is enabled from the structure refinement. During these refinements, the occupation factors of the M1 and M2 sites were kept unconstrained and refined to values between 0.97 and 0.99, indicating full occupation of M sites. To reduce refinement parameters, in the final refinements the site occupation of M1 and M2 was also fixed to 1.
Exceptions to these observations were crystals from the first batch of compositions with x Ge = 0.5. Modeling the electron density at the M1 and M2 sites, respectively, with Mg only, it turned out that the M1 site showed site occupations larger than 1, i.e., heavier elements than Mg were present to some extent. At the M2 site, however, the reverse was observed with occupation factors of ~0.8 only, meaning that some lighter elements were present. As the flux contained Li, Mo and V, the M1 site finally was modeled with a mixed occupation of M1 = Mo + Mg = 1. For the M2 site, Li was also assumed to be present here for the charge balance and M2 = Li + Mg = 1. With this model, charge-balanced structural formulas were almost obtained, yielding Mg0.84Mo0.05Li0.12Si0.56Ge0.44O3 as the chemical composition from SCXRD. A second crystal gave Mg0.85Mo0.03Li0.12Si0.60Ge0.40O3. The incorporation of Mo3+ into the structure of orthopyroxene was facilitated by the coupled substitution Mg1−3yMo3+yLi+3y(Si1−xGex)O3. More experiments would be necessary to clarify the maximum solubility of Mo3+ and Li+ in the pyroxene and the effect of Ge4+ by Si4+ substitution on it. The reason for the exceptional behavior of this run is unknown. It may be speculated that some problems occurred during furnace cooling as, in total, the number of single crystals in this run was low.
Because of the above behavior, new syntheses were performed with compositions of x = 0.50 to check for any unusual effects at the intermediate states of the solid-solution series. The obtained single crystals here fit perfectly to trends observed for all other compositions with no Li+ and Mo3+ incorporation.

3.2.1. Unit Cell Parameters

The variation in unit cell parameters and unit cell volume as a function of the Ge4+ content (x Ge) is shown in Figure 2. Data from the literature are included for comparison. It is evident that the increase in lattice parameters is smooth and follows a slightly non-linear trend for all three main axes. The substitution of the smaller Si4+ (Shannon ionic radius 0.26 Å) with the larger Ge4+ (Shannon radius = 0.40 Å) expands the a and c unit-cell dimension by ~3%, while the expansion along the b-axis is ~½ the value (Figure 2d). This smaller value can be related to changes in the kinking state of the tetrahedral chains, which mainly affects the b-axis.

3.2.2. Crystal Structure of MgSiO3-MgGeO3 Compounds

As mentioned in the Introduction, the structure of the orthopyroxenes is built up by three main building blocks: (i) octahedrally coordinated M1, (ii) octahedral M2 sites and (iii) two symmetrically distinct tetrahedral sites, TA and TB. A representative drawing of the structure of MgSiO3, based on data from this study, is given in Figure 3. The M1 octahedra share two common edges with neighboring M1 sites, forming the infinite zig-zag chain parallel to the c-axis (Figure 3c). The M2 sites are alternatively attached left and right of the M1 chain forming M-site bands within the (1 0 0)/bc plane. The orientation of the triangular faces of the M1 site with respect to the a-axis is of special interest. The tetrahedral sites share two common corners with each other (the O3 bridging oxygen atom O3br), forming a distinctly kinked chain running parallel the c-axis (Figure 3b). These tetrahedral chains also form layer-like units within the (1 0 0) plane. These M-site and T-site “layers” alternate along the a-axis. It is to be noted that these—of course—are not classical layers as individual chains are not interconnected with each other.
In the monoclinic polymorphs, the orientation of the M1 sites along the a-axis (expressed by the orientation of the triangular faces of the octahedra) is the same throughout the whole structure, while in the orthorhombic form, a stagging with different orientations (expressed as + and − signs) is found as displayed in Figure 3a. This doubles the a unit-cell parameter with respect to the monoclinic polymorphs and also allows the choice of an orthogonal cell. The infinite tetrahedral zig-zag chains TA and TB also alternate in their stagging along the a-axis (Figure 3a).
As can be seen from Figure 4, the cations exhibit small and almost isotropic atomic displacement parameters. The smallest values are found for the atoms at the tetrahedral site, indicating strong bonding here. Within the solid-solution series, there is no evident increase in the atomic displacement parameters, which indicates some positional disorder at the TA and TB sites. For the cations, the largest atomic displacement is found for the Mg2 site with its more distant and highly distorted oxygen atom environment. Chemical bonding is obviously slightly weaker here than for the Mg1 and TA and TB sites. Oxygen atoms are generally larger and also display partly more anisotropic atomic displacement parameters. The O1A and O1B oxygen atoms bonded to the M1 and the tetrahedral sites (via the apex of the tetrahedron) exhibit the smallest values among the O sites, while the O3A and O3B exhibit largest atomic displacements. The latter are more anisotropic with the longest elongation perpendicular to the O3-O3-O3 bonds within the tetrahedral chains.
Looking at bond valences, the Mg1 cation in both MgSiO3 and MgGeO3 is over-bonded with valence sums of 2.149(2) and 2.099(4) charges, respectively, while the Mg2 cation is under-bonded with 1.885(3) and 1.919(2) charges, respectively. For the tetrahedral sites, the TA appears to be close to the ideal value of 4, while the TB site is distinctly under-bonded with 3.831(5) and 3.877(7) charges for MgSiO3 and MgGeO3, respectively. Generally, it can be observed that bond valences in MgGeO3 are closer to the ideal value than in the Si4+ end-member. This also is expressed by a smaller global instability index GII in MgGeO3.

3.2.3. The M1 Site

The Mg2+ ion at the M1 site in MgSiO3 is six-fold coordinated to the O1Aiv,vi, the O1Bv,vii and the O2Aviii and the O2B oxygen atom, there is no link to the O3 tetrahedral bridging oxygen atoms (Figure 4). Individual M1 sites are connected to each other via the common O1A-O1B edges. The M1 site appears to be regular with Mg-O bond lengths in the range from 2.0049(9) to 2.1717(9) Å with an average value of <Mg1-O> = 2.0781 Å. Even if there is no change in chemistry at the M1 site, increasing Ge4+ substitution affects the M1 site as there is a direct connection of tetrahedral chains with the octahedral ones via common oxygen atoms O1A and O1B (apex oxygen atoms of the tetrahedral chain). The average <Mg1-O> bond length increases non-linearly by ~0.45% with the main increase at Si4+-rich compositions, while above x ~0.5 <Mg-O> remains almost constant. Data from the literature [20,35] are included for comparison and show the very same trend as that observed in this study; however, the trend is supported by more data with less scatter. The non-linear variation in <Mg1-O> is mainly due to the distinct anisotropic and also non-linear variation in individual M1-O bond lengths with increasing Ge4+ content (Table 3).
The changes in the bonds involving the O1A and O1B oxygen atoms can be directly related to the increasing kinking of the tetrahedral chains. Also, the O1Aiv-O1Avi and the O1Bv-O1Bvii octahedral edge become stretched (Figure 5b). The oxygen atoms involved here are not only bonded to the M1 site but also are the apex oxygen atoms of the two opposing tetrahedral chains TA and TB.
The bonds from Mg1 to the O2 oxygen atoms remain constant within 0.3%. These bonds involve the connection of the M1 sites with the tetrahedral sites within the b,c plane and appear to be less affected by the Si4+ with Ge4+ replacement on the neighboring tetrahedral sites. The volume of the M1 site increases slightly towards intermediate compositions and remains constant afterwards. The M1 site in pure MgSiO3 is remarkably regular, expressed by small numbers of the octahedral quadratic elongation (OQE) and octahedral angle variance (OAV). The Si4+ with Ge4+ substitution introduces increasing distortion both in the OQE as well as in the OAV, which increases by more than 60%.

3.2.4. The M2 site

The Mg2 cation in MgSiO3 is six-fold coordinated to the O1Aiv, O1Biii, O2A, O2B, O3Ai and O3Bii oxygen atoms; the resulting octahedral site, however, is highly distorted, especially in the MgSiO3 end-member. Four Mg-O bond lengths lie within 1.9952(8) and 2.0912(7) Å, while the bonds to the O3 bridging oxygen atoms O3Ai and O3Bii of the tetrahedral chain are distinctly stretched with 2.2876(8) and 2.4218(8) Å, respectively. The large distortion of the M2 site is mainly due to the bonding to these two oxygen atoms. With increasing substitution of Si4+ by Ge4+, marked changes can be observed especially for the Mg2-O3Ai and Mg2-O3Bii bond lengths, which distinctly decrease non-linearly (Figure 6a) by −2.6 and as much as −6.9%, respectively. The average M2-O bond length in sum decreases almost linearly up to x ~0.6 and then starts to flatten out (Figure 6b). Generally, <Mg2-O> is distinctly longer than <Mg1-O> by ~0.08 Å in pure MgSiO3. This difference is halved towards MgGeO3, but still amounts to ~0.04 Å.
The volume of the M2 octahedron decreases by ~2% toward MgGeO3 and approaches the value of the M1 octahedron. Due to the more regular bond length distribution, all distortion parameters decrease and the M2 site in MgGeO3 is much more regular than that in MgSiO3 (Figure 6c,d). It is worth recalling that the OAV increases by 60% for the M1 site, while it decreases by 17% with increasing Ge4+ content for the M2 site. Still, the M2 octahedron is also distinctly more distorted than the corresponding M1 octahedron in MgGeO3. Similar is observed for the other distortion parameters like the distortion index or the quadratic octahedral elongation. The observations of more regular M2 sites and a closer similarity between M1 and M2 octahedra in MgGeO3 is also depicted by the more ideal bond valence sums and smaller global instability index (GII) values as reported in Section 3.2.2.

3.2.5. The Tetrahedral Sites

Pyroxenes with primitive unit cells (P21/c, Pbca) show two symmetrically distinct tetrahedral chains, TA and TB, also allowing different kinking states. This is the main difference to C-centered pyroxenes, where the two chains become symmetric-equivalent. The tetrahedral cation is four-fold coordinated to the O1A, O2A, O3A and O3Ai oxygen atoms for the TA and to O1B, O2B, O3B and O3Bi for the TB site. In MgSiO3, the SiA-O bond lengths vary between 1.5909(9) Å and 1.6655(9) Å; the average bond length is <SiA-O> = 1.6293(9) Å. Generally, the distances to the bridging oxygen atoms, O3, are longer than those to O1 and O2. This is related to the stronger covalent character of the T-O3 bridging bond [35]. The difference between bridging <SiA-Obr> and non-bridging <SiA-Onbr> bond lengths amounts to 0.055 Å.
The SiB-O bonds in MgSiO3 vary between 1.5892(9) Å and 1.6793(10) Å, with an average value of <SiB-O> = 1.6420(10) Å. The difference between SiB-Obr and SiB-Onbr is 0.075 Å and thus larger than for the A-chain. Because of the larger distribution of Si-O bonds, the distortion index is somewhat larger for the SiB tetrahedron, while in terms of the bond angle variance and the tetrahedral quadratic elongation, the SiB tetrahedron is more regular. The SiB tetrahedron exhibits a somewhat larger volume (2.19 Å3 and 2.26 Å3 for SiA and SiB, respectively). The A-chain is more stretched than the B-chain, expressed by larger ∠ O3A-O3A-O3A = 158.96(6)°. For the B-chain, the ∠ O3B-O3B-O3B = 138.98(7)°, showing this chain to be distinctly more kinked.
An important finding of this study is how Ge4+ distributes over the two possible sites, TA and TB. The high-quality data of this study allow for the refinement of Si4+ and Ge4+ distribution over the two possible TA and TB sites. In agreement with results of Welch and Pawley [20], there is a distinct ordering of the Ge4+ onto the TB site, as displayed in Figure 7a. Data from this study and those from [20] perfectly plot onto each other. For low overall Ge4+ concentrations, almost three times more Ge4+ enters the B site than the A site. With increasing overall Ge4+ content, the ratio between Ge4+ on TA/TB sites approaches a value of 1:2 for ~MgSi0.65Ge0.35O3 and a ratio 1:1.5 for intermediate compositions. The TA/TB ratio becomes close to equal distribution only for very high overall Ge4+ contents. The variation in the Ge4+ distribution on TA and TB sites is non-linear. Due to different concentrations of Ge4+ on the two possible tetrahedral sites, considerations of variations in structural parameters are performed as a function of the Ge4+ content on the specific site.
Increasing the replacement of Si4+ by Ge4+ yields almost linear variations in average bond length with the Ge4+ content at the specific site. They increase by ~7.3% on average from MgSiO3 to MgGeO3 (Figure 7b). The difference between T-Obr and T-Onbr bonds is unaffected by the Ge4+ substitution at the A-chain, while for the B-chain it increases slightly. This can be interpreted as an increasing covalency of the bonding within the bridging bonds towards the MgGeO3 end-member in the B-chain. The volume of the TA and TB tetrahedra increases linearly with the Ge4+ content (Figure 7c) by 21.5% and 23.3%, respectively. There is a somewhat steeper increase in the volume of the B-chain tetrahedron, so the volume difference between the two tetrahedral sites also increases towards the MgGeO3 end-member.
The substitution of Ge4+ influences the geometry and the distortion state of the tetrahedra (Figure 7d). The distortion of the A-chain tetrahedra thereby increases in a more pronounced manner as compared to the B-chain ones. For the B-chain, the tetrahedral angle variance (TAV) as well as the tetrahedral quadratic elongation (TQE) increase by 26% and 0.1%, respectively. This is distinctly smaller than the changes in TAV and TQE for the A-chain tetrahedra, with an increase in TAV of 102% and in TQE of 1.1% from MgSiO3 to MgGeO3. It is to be noted that the slopes and variations in TAV and TQE are almost identical; thus, only the TAV is shown in Figure 7d. For the distortion index, which is based on the deviations in individual bond length from their mean value, the situation is reversed. It slightly decreases for the A-chain tetrahedra, while it tends to increase for the B-chain. This means that even if the T-O bonds at the A-tetrahedron become more similar, the bond angles deviate more from their ideal value towards MgGeO3.
As the size of the tetrahedra increases in course of the Si4+ by Ge4+ replacement (bond lengths increase by ~7%, volume increase by ~23%), however, the size of the M1 and M2 sites exhibit only small changes in the range of 1%; some effective mechanisms need to be at work to maintain connection, especially between the tetrahedral chains and the M1 octahedral zig-zag chains, which is sandwiched between a TA and TB chain. This mechanism is found in a rotation of the tetrahedra within the tetrahedral chains. With increasing Ge4+ content, the ∠ O3-O3-O3 bridging angles, which are the indicators for tetrahedral rotation in the pyroxenes, decrease for both sites, reflecting an increase in the tetrahedral kinking (Figure 8a). Different to [20], a somewhat non-parallel variation in angle reduction with increasing Ge4+ content is observed. Recalling the distinct changes in the Mg2-O3A and Mg2-O3B bond lengths, they can be explained as a direct consequence of this tetrahedral rotation. The smaller change in the Mg2-O3A bond lengths correlates with the smaller decrease in kinking of the A-chain. Also, the smaller change in the O1Aiv-O1Avi octahedral edge compared to the O1Bv-O1Bvii edge is related to the smaller rotation of the A-chain with Ge4+.
It is pointed out by [20] that the reduced kinking of the A-chain is due to a constraining effect of the O2A-O3A edge, which is shared by the TA and the M2 polyhedra. When looking at the variations in tetrahedral O-O edge lengths with the Ge4+ content, it becomes evident that by far the shortest among all the tetrahedral edges is indeed the shared O2A-O3A edge (Figure 8b). Also, the rate of lengthening with increasing Ge4+ content is smallest and amounts to 3.7% only, followed by 4.8% for the O3-O3i edge. Also, for this edge, some constraining effects of the sharing of the O3 oxygen atoms with the M2 site are observed. The remaining four edges of the TA tetrahedron expand by 7.9%–8.7% with increasing Ge4+ content. The TB tetrahedron shares only corners with the neighboring Mg2 polyhedron and no obvious hindering effects on edge length expansion is found.

4. Conclusions

In this study, the structural variations within a series of synthetic MgSi1−xGexO3 pyroxenes have been studied. Different to [20], where the synthesis was performed under low-pressure conditions and the orthorhombic structure was found only for 0.2 ≤ x ≤ 0.65, the orthorhombic Pbca symmetry is observed throughout the whole solid-solution series in this study. Even long storage times of crystals and annealing at 525 °C preserves orthorhombic symmetry. Most structural parameters vary linearly as a function of Si4+ with Ge4+ substitution. This especially accounts for the unit cell parameters and the bond lengths on the tetrahedral sites. So, the exchange of Si4+ with Ge4+ seems to be ideal and using the Ge analog of silicates is a justified approach to study phase relations and phase transitions. The changed topology of the tetrahedral sites also introduces distinct changes at the M1 and M2 sites, respectively. In particular, the highly distorted Mg2 is subjected to distinct changes in bond lengths to the O3 oxygen atoms. These changes are a direct consequence of the rotation of the tetrahedral chains to maintain connection between octahedral M1 and tetrahedral chains.
The preferential ordering of Ge4+ onto the TB tetrahedron can be explained by the fact that the A-chain shares a polyhedral edge with the distorted M2 octahedron. This edge forces a significant limitation in the expansion of the tetrahedral TA site. In contrast, there is only corner sharing of the TB tetrahedron with the M2 octahedron and thus can accommodate the expansion of the tetrahedron in due course of Si4+ simple with Ge4+ substitution by a rotation of the base plane of the tetrahedron within the infinite chain. What is more, the TB site is already somewhat larger in MgSiO3 and probably is more favorable for chemical substitution.

Author Contributions

Conceptualization, G.J.R.; methodology, G.J.R.; software, G.J.R.; investigation and synthesis, G.J.R. and G.T.; data curation, G.J.R. and G.T.; writing—original draft preparation, G.J.R. and G.T.; writing—review and editing, G.J.R. and G.T.; visualization, G.J.R. and G.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

A set of 10 representative compositions of this study, including anisotropic atomic displacement parameters, has been made available to the public by depositing them as a CIF (crystallographic information file) to the CCSD (Cambridge Crystallographic Data Centre) (https://www.ccdc.cam.ac.uk/) and can be retrieved under CSD Numbers 2371508–2371517. The full set of data may be requested from the first author as CIFs.

Acknowledgments

We would like to thank Gregor Zickler for help with optical microscopy imaging.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

The calculations of polyhedral volume and distortion parameters were implemented in VESTA [32]. This is easily achieved with the creation of a polyhedral view in VESTA and then clicking on the polyhedron under Investigation. After this, all the important information is displayed in the data window of VESTA. For more details, see the VESTA manual.
Distortion Index [32]
The distortion index is based on the variation in bond lengths, defined as
D = 1 n i = 1 n l i l a v l a v
where li is the distance from the central atom to the ith coordination atom, and lav is the average bond length.
Quadratic Elongation [32] (OQE and TQE)
The quadratic elongation, λ , was defined first by [36] and is calculated for platonic polyhedra (tetrahedra, octahedra, cubes, dodecahedra and icosahedra) only. It is given by
λ = 1 n i = 1 n l i l o 2
where lo is the center-to-vertex distance of a regular polyhedron of the same volume. λ is dimensionless, giving a quantitative measure of the polyhedral distortion, thereby being independent of the effective size of the polyhedron [32].
Bond Angle Variance [32] (OAV and TAV)
The bond angle variance σ 2 is again calculated only for platonic polyhedra and is given by [36]
σ 2 = 1 m 1 i = 1 m ϕ i ϕ o 2
where m is the number of faces in the polyhedron times 3/2 (i.e., number of bond angles), ϕ i is the ith bond angle and ϕ o is the ideal bond angle for a regular polyhedron (90° for an octahedron, 109.47° for a tetrahedron).

References

  1. Ringwood, A.E. Composition and Petrology of the Earth’s Mantle; Mac Graw-Hill Book Company: New York, NY, USA, 1975. [Google Scholar]
  2. Cameron, M.; Papike, J.J. Structural and chemical variations in pyroxenes. Am. Mineral. 1981, 66, 1–50. [Google Scholar]
  3. Yang, H.X.; Prewitt, C.T. Chain and layer silicates at high temperatures and pressures. High-Temp. High-Press. Cryst. Chem. 2000, 41, 211–255. [Google Scholar]
  4. Arlt, T.; Angel, R.J. Displacive phase transitions in C-centred clinopyroxenes: Spodumene, LiScSi2O6 and ZnSiO3. Phys. Chem. Miner. 2000, 27, 719–731. [Google Scholar] [CrossRef]
  5. Redhammer, G.J.; Ohashi, H.; Roth, G. Single-crystal structure refinement of NaTiSi2O6 clinopyroxene at low temperatures (298 < T < 100 K). Acta Crystallogr. Sect. B-Struct. Sci. 2003, 59, 730–746. [Google Scholar]
  6. Redhammer, G.; Nestola, F.; Miletich, R. Synthetic LiAlGe2O6: The first pyroxene with P21/n symmetry. Am. Mineral. 2012, 97, 1213–1218. [Google Scholar] [CrossRef]
  7. Shinmei, T.; Tomioka, N.; Fujino, K.; Kuroda, K.; Irifune, T. In situ X-ray diffraction study of enstatite up to 12 GPa and 1473 K and equations of state. Am. Mineral. 1999, 84, 1588–1594. [Google Scholar] [CrossRef]
  8. Hugh-Jones, D.A.; Woodland, A.B.; Angel, R.J. The structure of high-pressure C2/c ferrosilite and crystal chemistry of high-pressure C2/c pyroxenes. Am. Mineral. 1994, 79, 1032–1041. [Google Scholar]
  9. Ito, J. High temperature solvent growth of orthoenstatite, MgSiO3, in air. Geophys. Res. Lett. 1975, 2, 533–536. [Google Scholar] [CrossRef]
  10. Ulmer, P.; Stalder, R. The Mg(Fe)SiO3 orthoenstatite-clinoenstatite transitions at high pressures and temperatures determined by Raman-spectroscopy on quenched samples. Am. Mineral. 2001, 86, 1267–1274. [Google Scholar] [CrossRef]
  11. Ohi, S.; Miyake, A.; Yashima, M. Stability field of the high-temperature orthorhombic phase in the enstatite-diopside system. Am. Mineral. 2010, 95, 1267–1275. [Google Scholar] [CrossRef]
  12. Ohi, S.; Miyake, A. Phase transitions between high- and low-temperature orthopyroxene in the Mg2Si2O6-Fe2Si2O6 system. Am. Mineral. 2016, 101, 1414–1422. [Google Scholar] [CrossRef]
  13. Sokolova, T.S.; Dorogokupets, P.I.; Filippova, A.I. Equations of state of clino- and orthoenstatite and phase relations in the MgSiO3 system at pressures up to 12 GPa and high temperatures. Phys. Chem. Miner. 2022, 49, 37. [Google Scholar] [CrossRef]
  14. Ozima, M.; Akimoto, S.-I. Flux growth of single crystals of MgGeO3 polymorphs (orthopyroxene, clinopyroxene, and ilmenite) and their phase relations and crystal structures. Am. Mineral. 1983, 68, 1199–1205. [Google Scholar]
  15. Hunt, S.A.; Santangeli, J.R.; Dobson, D.P.; Wood, I.G. Phase diagram and thermal expansion of orthopyroxene-, clinopyroxene-, and ilmenite-structured MgGeO3. Am. Mineral. 2021, 106, 1113–1127. [Google Scholar] [CrossRef]
  16. Leinenweber, K.; Wang, Y.; Yagi, T.; Yusa, H. An unquenchable perovskite phase of MgGeO3 and comparison with MgSiO3 perovskite. Am. Mineral. 1994, 79, 197–199. [Google Scholar]
  17. Kirfel, A.; Neuhaus, A. Zustandsverhalten und elektrische Leitfähigkeit von MgGeO3 bei Drücken bis 65 kbar und Temperaturen bis 1300 °C (mit Folgerungen für das Druckverhalten von MgSiO3). Z. Phys. Chem. 1974, 91, 121–152. [Google Scholar] [CrossRef]
  18. Ross, N.L.; Navrotsky, A. Study of the MgGeO3 polymorphs (orthopyroxene, clinopyroxene, and ilmenite structures) by calorimetry, spectroscopy, and phase equilibria. Am. Mineral. 1988, 73, 1355–1365. [Google Scholar]
  19. Yamanaka, T.; Hirano, M.; Takeuchi, Y. A high temperature transition in MgGeO3 from clinopyroxene (C2/c) type to orthopyroxene (Pbca) type. Am. Mineral. 1985, 70, 365–374. [Google Scholar]
  20. Welch, M.D.; Pawley, A.R. Structural systematics of Ge substitution in primitive pyroxenes. Phys. Chem. Miner. 2016, 43, 161–169. [Google Scholar] [CrossRef]
  21. Kerschhofer, L.; Dupas, C.; Liu, M.; Sharp, T.G.; Durham, W.B.; Rubie, D.C. Polymorphic transformations between olivine, wadsleyite and ringwoodite: Mechanisms of intracrystalline nucleation and the role of elastic strain. Mineral. Mag. 1998, 62, 617–638. [Google Scholar] [CrossRef]
  22. Burnley, P.C. Investigation of the martensitic-like transformation from Mg2GeO4 olivine to its spinel structure polymorph. Am. Mineral. 2005, 90, 1315–1324. [Google Scholar] [CrossRef]
  23. Redhammer, G.J.; Roth, G.; Amthauer, G.; Lottermoser, W. On the crystal chemistry of olivine-type germanate compounds, Ca1+xM1−xGeO4 (M2+ = Ca, Mg, Co, Mn). Acta Crystallogr. Sect. B 2008, 64, 261–271. [Google Scholar] [CrossRef] [PubMed]
  24. Lambruschi, E.; Aliatis, I.; Mantovani, L.; Tribaudino, M.; Bersani, D.; Redhammer, G.J.; Lottici, P.P. Raman spectroscopy of CaM2+Ge2O6 (M2+=Mg, Mn, Fe, Co, Ni, Zn) clinopyroxenes. J. Raman Spectrosc. 2015, 46, 586–590. [Google Scholar] [CrossRef]
  25. Redhammer, G.J.; Roth, G.; Treutmann, W.; Paulus, W.; André, G.; Pietzonka, C.; Amthauer, G. Magnetic ordering and spin structure in Ca-bearing clinopyroxenes CaM2+(Si,Ge)2O6, M = Fe, Ni, Co, Mn. J. Solid State Chem. 2008, 181, 3163–3176. [Google Scholar] [CrossRef]
  26. Redhammer, G.J.; Roth, G.; Amthauer, G. Chromium-based clinopyroxene-type germanates NaCrGe2O6 and LiCrGe2O6 at 298 K. Acta Crystallogr. Sect. C-Cryst. Struct. Commun. 2008, 64, I97–I102. [Google Scholar] [CrossRef]
  27. Redhammer, G.J.; Roth, G.; Treutmann, W.; Hoelzel, M.; Paulus, W.; André, G.; Pietzonka, C.; Amthauer, G. The magnetic structure of clinopyroxene-type LiFeGe2O6 and revised data on multiferroic LiFeSi2O6. J. Solid State Chem. 2009, 182, 2374–2384. [Google Scholar] [CrossRef]
  28. Redhammer, G.J.; Senyshyn, A.; Meven, M.; Roth, G.; Prinz, S.; Pachler, A.; Tippelt, G.; Pietzonka, C.; Treutmann, W.; Hoelzel, M.; et al. Nuclear and incommensurate magnetic structure of NaFeGe2O6 between 5 K and 298 K and new data on multiferroic NaFeSi2O6. Phys. Chem. Miner. 2011, 38, 139–157. [Google Scholar] [CrossRef]
  29. Bruker. APEX3 (Version 2015. 10-0); Bruker AXS Inc: Madison, WI, USA, 2015. [Google Scholar]
  30. Sheldrick, G. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  31. Farrugia, L. WinGX and ORTEP for Windows: An update. J. Appl. Crystallogr. 2012, 45, 849–854. [Google Scholar] [CrossRef]
  32. 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]
  33. Nestola, F.; Gatta, G.D.; Ballaran, T.B. The effect of Ca substitution on the elastic and structural behavior of orthoenstatite. Am. Mineral. 2006, 91, 809–815. [Google Scholar] [CrossRef]
  34. Gatta, G.D.; Rinaldi, R.; Knight, K.S.; Molin, G.; Artioli, G. High temperature structural and thermoelastic behaviour of mantle orthopyroxene: An in situ neutron powder diffraction study. Phys. Chem. Miner. 2007, 34, 185–200. [Google Scholar] [CrossRef]
  35. Ozima, M. Structure of orthopyroxene-type and clinopyroxene-type magnesium germanium oxide MgGeO3. Acta Crystallogr. Sect. C 1983, 39, 1169–1172. [Google Scholar] [CrossRef]
  36. Robinson, K.; Gibbs, G.V.; Ribbe, P.H. Quadratic elongation—Quantitative measure of distortion in coordination polyhedra. Science 1971, 172, 567–570. [Google Scholar] [CrossRef]
Figure 1. Microscopy images of typical crystals from the flux-assisted synthesis of orthopyroxenes showing well-shaped, idiomorphic crystals of (a) MgSiO3 and (b) MgGeO3. The scale is 200 μm.
Figure 1. Microscopy images of typical crystals from the flux-assisted synthesis of orthopyroxenes showing well-shaped, idiomorphic crystals of (a) MgSiO3 and (b) MgGeO3. The scale is 200 μm.
Minerals 14 00864 g001
Figure 2. Variation in the unit cell parameters of synthetic orthopyroxenes along the MgSi1−xGexO3 solid-solution series. The gray diamonds correspond to data from the Li-Mo-containing orthopyroxenes from run MgGe50#1. Data from [20,35] are included as orange triangles. (ac) Variations in the a, b and c lattice parameters, respectively; (d) relative changes in the lattice parameters with respect to the values of pure end-member ortho-enstatite MgSiO3.
Figure 2. Variation in the unit cell parameters of synthetic orthopyroxenes along the MgSi1−xGexO3 solid-solution series. The gray diamonds correspond to data from the Li-Mo-containing orthopyroxenes from run MgGe50#1. Data from [20,35] are included as orange triangles. (ac) Variations in the a, b and c lattice parameters, respectively; (d) relative changes in the lattice parameters with respect to the values of pure end-member ortho-enstatite MgSiO3.
Minerals 14 00864 g002
Figure 3. Drawing of the structure of MgSiO3. M1-site octahedra are displayed in light green, M2 sites are in dark green, the TA tetrahedral sites appear in blue and the TB sites are colored in turquoise: (a) structure view along the c-axis; (b) excerpt of the structure showing two TA and TB tetrahedral chains in a view along the a-axis; (c) two bands of M1- and M2-site octahedra are shown in a view along the a-axis.
Figure 3. Drawing of the structure of MgSiO3. M1-site octahedra are displayed in light green, M2 sites are in dark green, the TA tetrahedral sites appear in blue and the TB sites are colored in turquoise: (a) structure view along the c-axis; (b) excerpt of the structure showing two TA and TB tetrahedral chains in a view along the a-axis; (c) two bands of M1- and M2-site octahedra are shown in a view along the a-axis.
Minerals 14 00864 g003
Figure 4. Asymmetric unit of the structure of MgSiO3 with some symmetry-equivalent oxygen atoms, filling up the coordination spheres around the Mg1, Mg2, TA and TB sites. The atomic nomenclature used within this paper is included. Atoms drawn as thermal ellipsoids at a probability level of 95%. Symmetry codes are (i) x, ½ − y, ½ + z; (ii) x, ½ − y, −½ + z; (iii) 1 − x, 1 − y, 1 − z; (iv) ½ − x, 1 − y, ½ + z; (v) 1 − x, ½ + y, 1.5 − z; (vi) ½ − x, ½ + y, 1 + z; (vii) 1 − x, 1 − y, 2 − z; (viii) x, y, 1 + z.
Figure 4. Asymmetric unit of the structure of MgSiO3 with some symmetry-equivalent oxygen atoms, filling up the coordination spheres around the Mg1, Mg2, TA and TB sites. The atomic nomenclature used within this paper is included. Atoms drawn as thermal ellipsoids at a probability level of 95%. Symmetry codes are (i) x, ½ − y, ½ + z; (ii) x, ½ − y, −½ + z; (iii) 1 − x, 1 − y, 1 − z; (iv) ½ − x, 1 − y, ½ + z; (v) 1 − x, ½ + y, 1.5 − z; (vi) ½ − x, ½ + y, 1 + z; (vii) 1 − x, 1 − y, 2 − z; (viii) x, y, 1 + z.
Minerals 14 00864 g004
Figure 5. Variations in average Mg1-O bond length (a), selected O-O octahedral edge lengths (b) and the octahedral distortion parameters OQE and OAV (c,d) for the synthetic pyroxenes along the MgSi1−xGexO3 solid-solution series. The gray triangles correspond to data from the literature. Error bars, if not visible, are smaller than the symbol. Symmetry codes for the atoms are given in Figure 4.
Figure 5. Variations in average Mg1-O bond length (a), selected O-O octahedral edge lengths (b) and the octahedral distortion parameters OQE and OAV (c,d) for the synthetic pyroxenes along the MgSi1−xGexO3 solid-solution series. The gray triangles correspond to data from the literature. Error bars, if not visible, are smaller than the symbol. Symmetry codes for the atoms are given in Figure 4.
Minerals 14 00864 g005
Figure 6. Variations in bond lengths (a,b) and distortion parameters (c,d) for the M2 site for the synthetic pyroxenes along the MgSi1−xGexO3 solid-solution series. Symmetry codes for the atoms are given in Figure 4.
Figure 6. Variations in bond lengths (a,b) and distortion parameters (c,d) for the M2 site for the synthetic pyroxenes along the MgSi1−xGexO3 solid-solution series. Symmetry codes for the atoms are given in Figure 4.
Minerals 14 00864 g006
Figure 7. Distribution of Ge4+ over the two possible sites, TA and TB (a); variations in the average bond lengths (b), the tetrahedral volume (c) and the tetrahedral angle variance (TAV) (d) for the tetrahedral sites for the synthetic pyroxenes along the MgSi1−xGexO3 solid-solution series. Data from [20] are included as light gray squares and light orange triangles in (a,b). Symmetry codes for the atoms are given in Figure 4.
Figure 7. Distribution of Ge4+ over the two possible sites, TA and TB (a); variations in the average bond lengths (b), the tetrahedral volume (c) and the tetrahedral angle variance (TAV) (d) for the tetrahedral sites for the synthetic pyroxenes along the MgSi1−xGexO3 solid-solution series. Data from [20] are included as light gray squares and light orange triangles in (a,b). Symmetry codes for the atoms are given in Figure 4.
Minerals 14 00864 g007
Figure 8. Degree of rotation of the tetrahedral chains (a) and variations in the edge lengths of the TA tetrahedra (b) for the synthetic pyroxenes along the MgSi1−xGexO3 solid-solution series.
Figure 8. Degree of rotation of the tetrahedral chains (a) and variations in the edge lengths of the TA tetrahedra (b) for the synthetic pyroxenes along the MgSi1−xGexO3 solid-solution series.
Minerals 14 00864 g008
Table 1. Experimental details and refinement results of X-ray diffraction data of selected flux-grown MgSi1−xGexO3 single crystals. Experiments were carried out with Mo Kα-radiation, λ = 0.71073 Å, using a Bruker Smart Apex 3-circle diffractometer. Full-matrix least-squares refinement of F2 crystal system is orthorhombic; space group is Pbca with Z = 8 for all data given.
Table 1. Experimental details and refinement results of X-ray diffraction data of selected flux-grown MgSi1−xGexO3 single crystals. Experiments were carried out with Mo Kα-radiation, λ = 0.71073 Å, using a Bruker Smart Apex 3-circle diffractometer. Full-matrix least-squares refinement of F2 crystal system is orthorhombic; space group is Pbca with Z = 8 for all data given.
Composition
Sample ID
MgSiO3
MgGe00_1
MgSi0.83Ge0.17O3
MgGe20_1
MgSi0.63Ge0.37O3
MgGe40_1
MgSi0.41Ge0.59O3
MgGe60_1
MgSi0.22Ge0.78O3
MgGe80_1
MgGeO3
MgGe100_1
Crystal Data
a (Å)18.2399(12)18.3324(3)18.4792(15)18.6067(5)18.7076(3)18.8189(7)
b (Å)8.8207(6)8.84471(15)8.8717(7)8.8974(3)8.92245(14)8.9516(3)
c (Å)5.1843(4)5.21380(9)5.2519(4)5.2863(2)5.31456(8)5.3470(2)
V (Å3)834.10(10)845.39(2)861.01(12)875.15(5)887.09(2)900.75(6)
Density mg/m33.1983.3953.6033.8414.0374.274
μ (mm−1)1.0983.3805.8998.63410.89013.575
Crystal Size (mm)0.120.140.110.130.120.14
0.070.080.100.110.100.11
0.070.080.080.090.090.10
Data Collection
s i n θ / λ m a x −1)0.6980.6940.6870.6940.6910.689
Reflections coll.10,42021,158992921,90718,42111,566
Independ. Refl.115011851182122212401250
Index Ranges: h−25 … 25−25 … 25−25 … 25−25 … 25−25 … 25−25 … 25
k−12 … 11−12 … 12−12 … 11−12 … 12−12 … 12−12 … 12
l−6 … 7−7 … 7−6 … 7−7 … 7−7 … 7−7 … 7
Completeness %99.799.799.699.699.899.8
Rint 0.0180.0200.0210.03070.03020.0216
Refinement
Data115011851182122212401250
Restrains022220
Parameters929696969692
R1 (all data)1.861.782.321.631.601.52
wR2 (all data)6.114.924.593.843.493.93
Goodness of Fit1.1561.2171.1361.1851.1191.188
Δ ρ m a x , Δ ρ m i n (e/Å)0.461/−0.3780.392/−0.3000.351/−0.3820.460/−0.2750.505/−0.3090.366/−0.572
Table 2. Atomic coordinates and equivalent isotropic displacement parameters for selected single crystals of the MgSi1−xGexO3 solid-solution series. U (eq) is defined as one-third of the trace of the orthogonalized Uij tensor.
Table 2. Atomic coordinates and equivalent isotropic displacement parameters for selected single crystals of the MgSi1−xGexO3 solid-solution series. U (eq) is defined as one-third of the trace of the orthogonalized Uij tensor.
IDWykoff Sitexyzs.o.f.U (eq)
MgGe00_1
Mg(1)8c0.3758(1)0.6537(1)0.8661(1)10.00507(13)
Mg(2)8c0.3768(1)0.4869(1)0.3589(1)10.00814(13)
SiA8c0.2717(1)0.3416(1)0.0504(1)10.00406(11)
SiB8c0.4736(1)0.3374(1)0.7983(1)10.00375(11)
O(1A)8c0.1834(1)0.3400(1)0.0351(2)10.0058(19)
O(2A)8c0.3109(1)0.5026(1)0.0433(2)10.00699(1)
O(3A)8c0.3032(1)0.2227(1)−0.1685(2)10.00671(1)
O(1B)8c0.5625(1)0.3402(1)0.8000(2)10.00603(19)
O(2B)8c0.4328(1)0.4832(1)0.6892(2)10.00712(19)
O(3B)8c0.4476(1)0.1950(1)0.6036(2)10.00654(18)
MgGe20_1
Mg(1)8c0.3754(1)0.6539(1)0.8624(1)10.00563(13)
Mg(2)8c0.3767(1)0.4871(1)0.3548(1)10.00872(14)
SiA8c0.2715(1)0.3420(1)0.0471(1)0.901(2)0.00414(12)
GeA8c0.2715(1)0.3420(1)0.0471(1)0.099(2)0.00414(12)
SiB8c0.4732(1)0.3385(1)0.8019(1)0.756(3)0.00430(10)
GeB8c0.4732(1)0.3385(1)0.8019(1)0.244(3)0.00430(10)
O(1A)8c0.1832(1)0.3401(1)0.0315(2)10.0070(2)
O(2A)8c0.3109(1)0.5035(1)0.0407(2)10.0082(2)
O(3A)8c0.3034(1)0.2219(1)−0.1707(2)10.0082(2)
O(1B)8c0.5632(1)0.3398(1)0.8045(2)10.0067(2)
O(2B)8c0.4321(1)0.4844(1)0.6842(2)10.0085(2)
O(3B)8c0.4467(1)0.1908(1)0.6108(2)10.0092(2)
MgGe40_1
Mg(1)8c0.3752(1)0.6545(1)0.8582(1)10.00758(16)
Mg(2)8c0.3767(1)0.4875(1)0.3505(1)10.00965(16)
SiA8c0.2713(1)0.3425(1)0.438(1)0.763(3)0.00547(12)
GeA8c0.2713(1)0.3425(1)0.0438(1)0.237(3)0.00547(12)
SiB8c0.4730(1)0.3393(1)0.8039(1)0.502(3)0.00565(10)
GeB8c0.4730(1)0.3393(1)0.8039(1)0.498(3)0.00565(10)
O(1A)8c0.1827(1)0.3402(1)0.0275(3)10.0087(3)
O(2A)8c0.3111(1)0.5050(2)0.0375(3)10.0101(3)
O(3A)8c0.3037(1)0.2200(2)−0.1717(3)10.0109(3)
O(1B)8c0.5640(1)0.3395(1)0.8080(2)10.0083(3)
O(2B)8c0.4313(1)0.4858(2)0.6791(3)10.0103(3)
O(3B)8c0.4459(1)0.1862(2)0.6178(3)10.0112(3)
MgGe60_1
Mg(1)8c0.3753(1)0.6550(1)0.8551(1)10.00596(14)
Mg(2)8c0.3768(1)0.4879(1)0.3472(1)10.00833(15)
SiA8c0.2712(1)0.3432(1)0.0420(1)0.542(2)0.00368(9)
GeA8c0.2712(1)0.3432(1)0.0420(1)0.458(3)0.00368(9)
SiB8c0.4727(1)0.3397(1)0.8047(1)0.284(3)0.00349(8)
GeB8c0.4727(1)0.3397(1)0.8047(1)0.716(3)0.00349(8)
O(1A)8c0.1818(1)0.3402(1)0.0248(2)10.0075(2)
O(2A)8c0.3115(1)0.5080(1)0.0361(2)10.0092(3)
O(3A)8c0.3046(1)0.2174(1)−0.1703(2)10.0100(3)
O(1B)8c0.5646(1)0.3390(1)0.8114(2)10.0068(2)
O(2B)8c0.4308(1)0.4872(1)0.6746(2)10.0087(2)
O(3B)8c0.4451(1)0.1828(1)0.6226(2)10.0088(3)
MgGe80_1
Mg(1)8c0.3760(1)0.6557(1)0.8531(1)10.00674(14)
Mg(2)8c0.3773(1)0.4884(1)0.3454(1)10.00840(14)
SiA8c0.2710(1)0.3441(1)0.0416(1)0.298(3)0.00445(8)
GeA8c0.2710(1)0.3441(1)0.0416(1)0.702(3)0.00445(8)
SiB8c0.4724(1)0.3397(1)0.8052(1)0.156(3)0.00423(7)
GeB8c0.4724(1)0.3397(1)0.8052(1)0.844(3)0.00423(7)
O(1A)8c0.1806(1)0.3403(1)0.0233(2)10.0080(2)
O(2A)8c0.3120(1)0.5115(1)0.0360(2)10.0097(3)
O(3A)8c0.3058(1)0.2142(1)−0.1669(2)10.0101(3)
O(1B)8c0.5646(1)0.3387(1)0.8134(2)10.0074(2)
O(2B)8c0.4306(1)0.4878(1)0.6724(2)10.0089(2)
O(3B)8c0.4447(1)0.1810(1)0.6248(2)10.0085(3)
MgGe100_1
Mg(1)8c0.3765(1)0.6561(1)0.8516(1)10.00316(13)
Mg(2)8c0.3778(1)0.4887(1)0.3437(1)10.00705(14)
GeA8c0.2710(1)0.3450(1)0.0413(1)10.00365(7)
GeB8c0.4722(1)0.3396(1)0.8053(1)10.00332(7)
O(1A)8c0.1794(1)0.3399(2)0.0219(3)10.0048(3)
O(2A)8c0.3124(1)0.5148(2)0.0352(3)10.0073(3)
O(3A)8c0.3070(1)0.2110(2)−0.1639(2)10.0063(3)
O(1B)8c0.5646(1)0.3379(1)0.8149(3)10.0050(3)
O(2B)8c0.4305(1)0.4886(2)0.6710(2)10.0069(3)
O(3B)8c0.4444(1)0.1798(2)0.6262(3)10.0056(3)
Table 3. Selected bond length and distortion parameters for single crystals along the MgSi1−xGexO3 solid-solution series.
Table 3. Selected bond length and distortion parameters for single crystals along the MgSi1−xGexO3 solid-solution series.
Composition
Sample ID
MgSiO3
MgGe00_1
MgSi0.83Ge0.17O3
MgGe20_1
MgSi0.63Ge0.37O3
MgGe40_1
MgSi0.41Ge0.59O3
MgGe60_1
MgSi0.23Ge0.77O3
MgGe80_1
MgGeO3
MgGe100_1
M1 site
Mg1-O2Aviii (Å)2.0049(9)2.0086(12)2.0121(15)2.0095(13)2.0082(13)2.0052(13)
Mg1-O1Aiv (Å)2.0282(10)2.0329(12)2.0407(15)2.0450(13)2.0484(13)2.0533(14)
Mg1-O2B (Å)2.0454(10)2.0470(12)2.0502(14)2.0507(13)2.0520(13)2.0524(13)
Mg1-O1Bvii (Å)2.0656(9)2.0700(12)2.0829(14)2.0883(13)2.0924(13)2.0995(13)
Mg1-O1Avi (Å)2.1526(9)2.1547(11)2.1530(14)2.1566(13)2.1565(13)2.1547(13)
Mg1-O1Bv (Å)2.1717(9)2.1742(11)2.1719(14)2.1686(12)2.1646(13)2.1598(13)
<Mg1-O> (Å)2.07812.08122.08512.08652.08702.0875
Volume (Å3)11.82411.86711.93011.92411.91511.912
Dist. Index (a)0.02700.02670.02500.02470.02440.0242
OQE (b)1.00891.00951.01011.01121.01221.0133
OAV (c)26.7228.8631.2735.2139.0142.94
M2 site
Mg2-O2B (Å)1.9944(10)1.9955(12)2.0001(15)2.0019(13)2.0036(13)2.0111(13)
Mg2-O2A (Å)2.0344(9)2.0388(12)2.0481(15)2.0526(13)2.0586(13)2.0712(13)
Mg2-O1Biii (Å)2.0570(9)2.0607(11)2.0619(14)2.0650(13)2.0671(13)2.0744(13)
Mg2-O1Aiv (Å)2.0910(10)2.0951(12)2.0994(15)2.1010(13)2.0987(13)2.1019(13)
Mg2-O3Ai (Å)2.2892(9)2.2895(12)2.2838(15)2.2690(13)2.2492(13)2.2299(13)
Mg2-O3Bii (Å)2.4484(10)2.3961(12)2.3456(15)2.3087(13)2.2909(13)2.2804(13)
<Mg2-O> (Å)2.15242.14602.13982.13292.12802.1281
Volume (Å3)12.47912.39212.32312.24312.19512.244
Dist. Index (a)0.06700.06120.05450.04880.04450.0398
OQE (b)1.04911.04661.04341.04051.03791.0359
OAV (c)140.95137.523131.849126.098120.083116.143
Tetrahedral A-site
GeA-O2A (Å)1.5909(9)1.6005(11)1.6188(14)1.6472(13)1.6782(12)1.7085(12)
GeA-O1A (Å)1.6124(10)1.6218(11)1.6395(14)1.6656(13)1.6944(12)1.7281(13)
GeA-O3A (Å)1.6484(9)1.6612(11)1.6795(14)1.7032(13)1.7307(12)1.7615(12)
GeA-O3Ai (Å)1.6655(9)1.6808(11)1.7022(14)1.7292(12)1.7589(12)1.7876(12)
<GeA-O> (Å)1.62931.64111.66001.68631.71561.7464
Volume (Å3)2.1892.2342.3072.4102.5272.658
Dis. Index (a)0.01700.01830.01860.01770.01710.0161
TQE (d)1.00981.01061.01011.01421.01721.0202
TAV (e)39.22242.59148.00256.47167.85178.566
O3A-O3A-O3A (°)158.96(8)158.39(9)157.05(9)155.22(8)152.94(9)150.74(8)
Tetrahedral B site
GeB-O2B (Å)1.5892(9)1.6154(11)1.6436(15)1.6741(12)1.6907(12)1.7060(12)
GeA-O1B (Å)1.6223(9)1.6503(11)1.6827(13)1.7099(12)1.7253(12)1.7405(14)
GeA-O3B (Å)1.6771(9)1.7131(12)1.7470(14)1.7718(12)1.7871(12)1.7985(12)
GeA-O3Bi (Å)1.6793(9)1.7017(11)1.7375(14)1.7686(12)1.7857(12)1.8022(12)
<GeA-O> (Å)1.64201.67011.70271.73111.74721.7618
Volume (Å3)2.25592.37262.5162.64042.7152.787
Dis. Index (a)0.02200.02230.02280.02260.02240.0219
TQE (d)1.00521.00561.00591.00611.00611.0063
TAV (e)19.57221.39022.87424.06224.39925.177
O3A-O3A-O3A (°)138.98(8)136.21(9)133.33(10)131.30(9)130.26(8)129.68
(a) Distortion index, (b) OQE = octahedral quadratic elongation, (c) OAV = octahedral angle variance, (d) TQE = tetrahedral quadratic elongation, (e) TAV = tetrahedral angle variance. All values calculated with VESTA [32]. For definition of parameters, see Appendix A.
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

Redhammer, G.J.; Tippelt, G. Crystal Chemistry of Synthetic Mg(Si1−xGex)O3 Pyroxenes: A Single-Crystal X-ray Diffraction Study. Minerals 2024, 14, 864. https://doi.org/10.3390/min14090864

AMA Style

Redhammer GJ, Tippelt G. Crystal Chemistry of Synthetic Mg(Si1−xGex)O3 Pyroxenes: A Single-Crystal X-ray Diffraction Study. Minerals. 2024; 14(9):864. https://doi.org/10.3390/min14090864

Chicago/Turabian Style

Redhammer, Günther J., and Gerold Tippelt. 2024. "Crystal Chemistry of Synthetic Mg(Si1−xGex)O3 Pyroxenes: A Single-Crystal X-ray Diffraction Study" Minerals 14, no. 9: 864. https://doi.org/10.3390/min14090864

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