Next Article in Journal
Synthesis of Metal Organic Frameworks by Ball-Milling
Next Article in Special Issue
Photocatalytic Decolorization of Methyl Red on Nanoporous Anodic ZrO2 of Different Crystal Structures
Previous Article in Journal
A New Piano-Stool Ruthenium(II) P-Cymene-Based Complex: Crystallographic, Hirshfeld Surface, DFT, and Luminescent Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Crystal Structure and Some Thermodynamic Properties of Ca7MgSi4O16-Bredigite

1
Key Laboratory of Orogenic Belts and Crustal Evolution, Ministry of Education of China, Beijing 100871, China
2
School of Earth and Space Sciences, Peking University, Beijing 100871, China
3
School of Gemmology, China University of Geosciences (Beijing), Beijing 100083, China
4
Key Laboratory of Earth and Planetary Physics, Institute of Geology and Geophysics, Chinese Academy of Sciences, Beijing 100029, China
*
Author to whom correspondence should be addressed.
Crystals 2021, 11(1), 14; https://doi.org/10.3390/cryst11010014
Submission received: 29 November 2020 / Revised: 20 December 2020 / Accepted: 24 December 2020 / Published: 26 December 2020
(This article belongs to the Special Issue Nanostructured Crystalline Materials)

Abstract

:
Bredigite with the composition Ca7MgSi4O16 (Ca7MgSi4O16-Bre) has been synthesized by a solid-state reaction method at 1.2 GPa and 1373 K for 7 days, and its structure has been determined by single-crystal X-ray diffraction data. Following a relevant genealogy analysis in the literature, we have refined the structure into two space groups, Pnnm and Pnn2, and found that Ca7MgSi4O16-Bre belongs to the space group Pnnm, which can be essentially derived from the space group Pnn2 via an atomic coordinate transformation (with an average deviation of 0.039 Å only). Furthermore, some thermodynamic properties of the Ca7MgSi4O16-Bre have been obtained in this study. Using first-principles simulations based on density functional theory, the isothermal bulk modulus has been determined as 90.6(4) GPa with a pressure derivative of 5.7(1). Using density functional perturbation technique, the phonon dispersions and vibrational density of the states (VDoS) have been calculated. The VDoS has been combined with a quasi-harmonic approximation to compute the isobaric heat capacity (Cp) and standard vibrational entropy ( S 298 0 ), yielding Cp = 8.22(2) × 102 − 3.76(6) × 103T−0.5 − 1.384(4) × 107T−2 + 1.61(8) × 109T−3 J mol−1 K−1 for the T range of 298-1000 K and S 298 0 = 534.1 (22) J mol−1 K−1.

1. Introduction

The continental crust of the Earth is rich in CaO and SiO2, and may form some special calcium silicates while it is subducted to certain depths of the mantle. Recently larnite-structured Ca2SiO4−, walstromite-structured CaSiO3, and titanite-structured CaSi2O5 were found as mineral inclusions in diamonds from the mantle [1,2]. While residing in the mantle, the continental crust material may interact with the surrounding MgO-rich mantle [3] and produce some special calcium magnesium silicates such as merwinite (Ca3MgSi2O8), diopside (CaMgSi2O6), monticellite (CaMgSiO4), and akermanite (Ca2MgSi2O7). Among these calcium magnesium silicates, merwinite has been discovered as mineral inclusions in some diamonds from São Luiz, Brazil [4], suggesting that the interaction between the subducted CaO- and SiO2-rich continental crust material and the MgO-rich mantle material happens indeed.
Bredigite, with a typical composition Ca7MgSi4O16, is such a potential calcium magnesium orthosilicate, which may form by the interaction between the subducted Ca-rich continental material and the MgO-rich mantle material at appropriate PT conditions. The word “bredigite”, abbreviated as Bre hereafter, was first used by Tilley and Vincent [5] to describe a mineral coexisting with spurrite, larnite, and gehlenite in the contact zone of Chalk and Tertiary dolerite at Scawt Hill, Northern Ireland. It caused substantial confusion in terms of chemical composition, phase stability, and crystallographic structure [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19]. Thanks to enormous effort along all these years, it has become generally clear that the Bre from Scawt Hill is compositionally identical to "Phase T" experimentally observed by Gutt [7,8], with mutual substitution of Ca and Mg to some small extents [17,20] but without any significant Ba in the structure [11,14,16,19]. It has also become clear that Bre is stable up to ~1372 °C at ambient P [8,9,10,12,17]. Recently Xiong [20] experimentally demonstrated that Bre is stable at ~1.2 GPa.
The structure of Bre, however, remains unclear. A large number of studies provided powder X-ray diffraction patterns, but none generated single-crystal X-ray diffraction data [7,10,17,19]. It was proposed by Tilley and Vincent [5] that Bre might be structurally identical to a synthetic Ba-bearing calcium magnesium orthosilicate phase observed in some spiegeleisen slags, on the basis of the similarities of their morphologies and optical properties (compositionally approximating Ca6.36Ba0.32Mg1.24Mn0.36Si4O16). Accordingly, two single-crystal X-ray diffraction investigations were conducted with some crystals extracted from the spiegeleisen slags [6,13]. However, the synthetic crystals used in these studies contained a trigonal phase with a volume fraction up to ~20%, which substantially reduced the quality of the X-ray data. Additionally, it was rather unclear how the incorporation of Ba into the crystals might have distorted the structure, as Ba is significantly larger than Ca (e.g., the cation radius of 12-fold Ba is 1.61 Å whereas that of 12-fold Ca is 1.34 Å [21]). Indeed, Moseley and Glasser proposed different structures for them, with the Ba-bearing phase crystallizing in the space group Pnn2 and the Ba-free phase in the space group Pmnn [17].
Along with the geological significance of Bre, Bre is not only an excellent bioactive ceramic [22] but also a high-quality phosphor candidate emitting special lights when doped with some particular rare earth elements [23]. A good understanding of the detailed structure of Bre will certainly benefit all these research fields.
Here, we synthesized the Ca7MgSi4O16-Bre by a solid-state method at high-temperature and high-pressure conditions, and determined its structure using single-crystal X-ray diffraction method. In addition, we theoretically calculated some of its thermodynamic properties using first-principles simulations. Before this study, there were no thermodynamic data for the Ca7MgSi4O16-Bre [24].

2. Experimental and Theoretical Methods

2.1. Synthesis

The Ca7MgSi4O16-Bre investigated in this study was synthesized by a solid-state reaction method at high PT conditions. The starting material was made as follows: (1) 99.9% pure chemical powders CaCO3, MgO, and SiO2 from Alfa Aesar(Alfa Aesar, Haverhill, MA, USA)were pretreated at 1 atm and 723 K for 72 hours to remove any absorbed moisture, immediately weighed in a mole ratio of 7:1:4, and then ground and homogenized in an agate mortar; (2) the resulting mixture was pressed into a pellet and degassed in a Pt crucible at 1 atm and 1273 K for 48 hours; (3) the degassed pellet was crushed and ground into a fine powder, which was further stored in a drying oven at 383 K and used as the starting material for a later synthesizing experiment. The starting material was loaded in a Pt capsule (10 mm in length and 2.5 mm in diameter) with both ends sealed using an arc-welding technique. The synthesizing experiment was carried out with a non-end loaded piston–cylinder apparatus installed at the High-Pressure Laboratory of Peking University (Quickpress 3.0, Depths of The Earth Company, Cave Creek, AZ, USA) [25]. The experimental assembly and high-P experimental technique were generally identical to those reported in Liu and Fleet [26]. The experimental PT conditions were 1.2 GPa and 1373 K, and the heating time was 7 days.

2.2. Characterization

The high-P experimental product was polished with a series of diamond pastes, cleaned in alcohol using an ultrasonic washing machine, carbon-coated, and then examined for texture and composition respectively with a scanning electron microscope (Quanta 650 FEG, FEI Company, Hillsboro, OR, USA) and an electron microprobe (JXA-8100, JEOL, Tokyo, Japan). It was found out that the product contained a single crystalline phase with grain sizes ranging from 50–100 microns (Supplementary Materials Figure S1). Moreover, 13 electron microprobe analyses performed on randomly selected grains suggested the following chemical formula Ca7.01(2)Mg1.01(1)Si3.99(1)O16, closely matching the targeted formula Ca7MgSi4O16. The product was eventually slightly pulverized, and some loose grains with appropriate sizes and shapes were hand-picked for a later single-crystal X-ray diffraction experiment. The rest of the sample was ground down to a fine powder and analyzed with a powder X-ray diffractometer (D/Max 2550 V/PC with graphite-monochromated Cu Kα radiation, Rigaku Corporation, Tokyo, Japan) at ambient PT conditions. The powder XRD pattern of the Ca7MgSi4O16-Bre, as shown in Figure 1, is consistent with the simulated XRD pattern based on the single-crystal structural analysis (more later), indicating that they are isostructural.

2.3. Single-Crystal X-Ray Diffraction

The single crystals were immersed in oil and examined under an optical microscope. A suitable crystal was then selected for the single-crystal X-ray diffraction analysis. The intensity data were collected on a Bruker Smart Apex III micro-focused diffractometer using Mo Kα radiation (λ = 0.71073 nm). The raw data were processed and corrected for the absorption effects using the programs SAINT+ and SADAB. An initial structure solution was obtained via direct methods and refined by a full-matrix least-squares method using the SHELXT software included in the SHELXTL package. All heaviest atoms were first located unambiguously in the Fourier maps, and the O atoms were later found in the subsequent difference Fourier maps. All atoms were refined with the anisotropic displacement parameters. For the atoms with I > 2sigma(I), their atomic coordinates and anisotropic thermal parameters were included in the final cycles of the least-squares refinement.

2.4. Computational Method

The first-principles simulations carried out to investigate the compressibility of the Ca7MgSi4O16-Bre were deployed with the CASTEP code using density functional theory [27,28] and planewave pseudopotential technique [29]. The exchange–correlation interaction was treated by the generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE) functional [30], and a convergence criterion of 5 × 10−7 eV/atom was used in the self-consistent field calculations. We employed a planewave basis set with a cutoff of 990 eV to expand the electronic wave functions, and a norm-conserving pseudopotential to model the ion–electron interaction [31,32]. We sampled the irreducible Brillouin zone with a 2 × 6 × 4 Monkhorst-Pack grid [33]. The effects of using larger cutoff and k point mesh on the calculated properties were found to be insignificant. The computation cell contained four Ca7MgSi4O16-Bre molecules (in total 112 atoms), with the initial structure model from our single-crystal X-ray analysis. The equilibrium lattice parameters and internal coordinates at different pressures were optimized by minimizing the Hellmann–Feynman force on the atoms and simultaneously achieving the desired hydrostatic stress tensor. These theoretical techniques were used in our previous studies for the structures and thermodynamics of some silicate minerals [34,35].
Based on the optimized structure, the phonon dispersions and vibrational density of the states (VDoS) of the Ca7MgSi4O16-Bre were further calculated by diagonalizing the dynamical matrix with density functional perturbation theory [36,37]. The q-vector grid spacing for interpolation was 0.05 Å−1, which represented the average distance between the Monkhorst-Pack q-points used in the dynamical matrix calculations. The phonon dispersions were obtained at the high-symmetry points (G, Z, T, Y, S, X, U, R). The coordinates of these points on the surface of the Brillouin zone were G = (0 0 0), Z = (0 0 1/2), T = (−1/2 0 1/2), Y = (−1/2 0 0), S = (−1/2 1/2 0), X = (0 1/2 0), U = (0 1/2 1/2), R = (−1/2 1/2 1/2).

3. Results and Discussions

3.1. Structural Analysis

Moore and Araki [13] collected single-crystal X-ray diffraction data from the Ba-rich, Bre-like phase (Ca6.15Ba0.3Mg1.2Mn0.35Si4O16) extracted from the spiegeleisen slags and found its space group to be Pnn2. Taking into account the crystal chemistry of this phase, they performed a genealogy analysis of some phases in the CaO–MgO–SiO2 system and proposed that the ideal Bre with the Ca7MgSi4O16 composition might adopt the space group Pnnm.
Using our single-crystal X-ray diffraction data, we solved the structure in the space group Pnnm, with a = 18.3434(17) Å, b = 6.7313(8) Å, c = 10.8844(12) Å, α = 90°, β = 90°, γ = 90°, and V = 1344.0(3) Å3 (Table 1). In this structure, each asymmetric unit contains two distinct Mg sites, three distinct Si sites, and six distinct Ca sites (for their coordinates and equivalent isotropic displacement parameters, see Table 2). As shown in Figure 2a, both Mg atoms are coordinated to six bridging O atoms to form [MgO6] octahedra. The Mg1–O bond lengths vary from 2.003(4) to 2.163(4) Å, the Mg2–O bond lengths from 2.044(4) to 2.080(6) Å, and the O–Mg–O angles from 87.33(15)° to 180.000(1)° (Table 3). All the Si sites are in 4-fold coordination, with the Si–O bond lengths varying from 1.594(4) to 1.637(4) Å (Table 3). Further, every [SiO4] tetrahedron shares two O corners with two neighboring [MgO6] octahedra, leading to the formation of an [Mg–O–Si] chain, which runs along the c-axis (Figure 2b). The [Mg–O–Si] chains are the basic structural units of the Ca7MgSi4O16-Bre. The Ca sites are much more complicated (Table 3): Ca1, 6-coordinate (Ca1–O ranging from 2.275(5) to 2.465(4) Å); Ca2, 10-coordinate (Ca2–O ranging from 2.305(5) to 2.886(4) Å); Ca3, 7-coordinate (Ca3–O ranging from 2.212(6) to 2.989(4) Å); Ca4, 7-coordinate (Ca4–O ranging from 2.327(6] to 2.579(5) Å); Ca5, 9-coordinate (Ca5–O, ranging from 2.253(4] to 2.938(5) Å); Ca6, 8-coordinate (Ca6–O ranging from 2.308(4) to 2.701(4) Å).
We also attempted to solve the structure in the space group Pnn2, with the unit-cell parameters a′, b′, and c′, respectively, equivalent to the unit-cell parameters a, b, and c of the Pnnm space group (Table 1). The resulting atomic coordinates and equivalent isotropic displacement parameters are listed in Supplementary Table S1. In the Pnn2 model, the atoms Ca1, Ca2, Mg1, and Mg2 deviate from symmetric centers in the Pnnm model (Figure 3), which splits the mirror plane and lowers the symmetry from Pnnm to Pnn2. Nevertheless, the atom arrangement in the Pnn2 model is very similar to that in the Pnnm model. We therefore used the PLATON software to exam the relations between these two structural models. The result suggests that the Pnn2 model can transform to the Pnnm model via the following matrix operation:
( x y z ) = ( x y z ) × [ 1.0000 0.0000 0.0000 0.0000 1.0000 0.0000 0.0000 0.0000 1.0000 ] + ( 0.0000 0.0000 0.2344 )
The average deviation of the atom positions in the two Pnnm models is just 0.039 Å. This indicates that the space group of the Ca7MgSi4O16-Bre is Pnnm indeed.
It is thus clear that due to substantial amounts of Ba2+ and Mn2+ substitutions, respectively, for Ca2+ and Mg2+, the atoms on the Ca2, Mg1, and Mg2 sites of the Pnnm model can significantly depart from high-symmetry sites, leading to the space group Pnn2, as observed by Moore and Araki [13]. Recently, a natural Mn-free Bre with the composition Ca7.006Na0.015Ba0.014Mg0.938Si4.000P0.014O16 was reported from the Hatrurim Complex, Negev Desert, Israel [38], and its space group was determined as Pnnm as well. This suggests that small amounts of impurities such as Na, Ba, and P incorporated into the structure do not alter the basic crystallographic features of the Ca7MgSi4O16-Bre. As a result, it is important to investigate the physical–chemical properties of Ca7MgSi4O16-Bre.

3.2. Isothermal Bulk Modulus, Heat Capacity, and Entropy

The unit-cell parameters of the Ca7MgSi4O16-Bre at zero pressure obtained from the first-principles calculation are summarized in Table 4. They compare well with the results obtained from the single-crystal X-ray diffraction data.
The EV data from our static calculation for the Ca7MgSi4O16-Bre are shown in Figure 4. To determine the isothermal bulk modulus, the third-order EV Birch–Murnaghan equation of state [39] was fitted with the EV data by a least-squares method:
E ( V ) = E 0 + 9 B 0 V 0 16 [ ( V 0 V ) 2 3 1 ] 2 × [ ( B 0 4 ) × ( V 0 V ) 2 3 B 0 + 6 ]
where E is the free energy, B0 is the isothermal bulk modulus, B 0 is the first pressure derivative of B0, V0 is the volume at zero pressure, and V are the volumes at other pressures. When B 0 is set as 4, B0 is 93.1(22) GPa and V0 is 1394.2(16) Å3. If B 0 is not fixed, B0 is 90.6(4) GPa, B 0 is 5.7(1), and V0 is 1388.0(6) Å3.
We calculated the phonon dispersions and VDoS. The dynamical matrices were computed at 21 wavevectors in the Brillouin zone of the cell and interpolated to obtain the bulk phonon dispersions. Figure 5 shows the dispersion curves along several symmetry directions and the VDoS. Clearly, the Ca7MgSi4O16-Bre in the space group Pnnm is dynamically stable.
The phonon spectrum for the Ca7MgSi4O16-Bre has been used to compute the internal energy (E) and isochoric heat capacity (Cv) as functions of temperature. The temperature dependence of the E was obtained by the following equation,
E ( T ) = E t o t + E z p + h ω e x p ( h ω k T ) 1 F ( ω ) d ω
with Etot representing the total electronic energy at 0 K, Ezp the zero-point vibrational energy, h the Planck’s constant, k the Boltzmann constant, and F(ω) the vibrational density of states. We evaluated the Ezp term in Equation (3) using the following equation,
E z p = F ( ω ) h ω d ω
Furthermore, we approximated the lattice contribution to Cv with the following equation:
C v ( T ) = k ( h ω k T ) 2 e x p ( h ω k T ) [ ( h ω k T ) 2 1 ] 2 F ( ω ) d ω
The Cv result calculated in this way is shown in Table 5.
Cp was calculated by adding an anharmonic effect to Cv obtained from the above calculations, using the following equation:
C p = C v + α T 2 B T V T T
where αT, BT, and VT are the thermal expansion coefficient, isothermal bulk modulus, and volume at 1 atm and T (K), respectively. The thermal expansion coefficient, as shown in Figure 6 (with some values at some selected temperatures listed in Table 5), was calculated by the Gibbs software through a quasi-harmonic Debye model [40]. The temperature derivative of BT was assumed as zero in our calculation. VT at T K was calculated with the following equation:
V T = V 298 e x p ( 298 T α T d T )
where V298 = 809.1(2) cm3/mol. The calculated Cp values are listed in Table 5, and compared to the Cv values in Figure 7. The Cp values are divided into three T ranges, 10–100, 100–298, and 298–1000 K, and empirically fitted to the equation from [24], with the coefficients summarized in Table 6.
The Cp values have been applied to the calculation of the vibrational entropy at T K using the following equation,
S T 0 = 0 T C p T d T
The vibrational entropy of the Ca7MgSi4O16-Bre at 298 K ( S 298 0 ) is calculated as 534.1 (22) J mol−1 K−1.

4. Conclusions

Ca7MgSi4O16-Bre has been successfully synthesized at 1.2 GPa and 1373 K, and its structure has been determined by single-crystal X-ray diffraction data. We find that the Ca7MgSi4O16-Bre has the space group Pnnm. Although this space group may be reduced to the space group Pnn2 as significant amounts of Ba and Mn enter the structure, it tolerates low levels of chemical impurities such as Na, Ba, and P.
Some first-principles calculations have been carried out as well. The results suggest that the isothermal bulk modulus of the Ca7MgSi4O16-Bre is 90.6(4) GPa (with a pressure derivative of 5.7(1)), the isobaric heat capacity is Cp = 8.22(2) × 102 – 3.76(6) × 103T−0.5 – 1.384(4) × 107T−2 + 1.61(8) × 109T−3 J mol−1 K−1 (298–1000 K), and the standard vibrational entropy is S 298 0 = 534.1 (22) J mol−1 K−1.

Supplementary Materials

The followings are available online at https://www.mdpi.com/2073-4352/11/1/14/s1, Figure S1: Back scattered electron (BSE) image of bredigite, Table S1: atomic coordinates and U(eq) for Ca7MgSi4O16 (Pnn2), crystal reflection data files (bredigite.HKL) and crystallographic information files (bredigite.CIF).

Author Contributions

X.B., synthesizing and characterizing the material, collecting and refining the single-crystal X-ray data, performing theoretical calculations, interpreting the results, and writing the paper; M.H., designing and supervising the project and discussing and interpreting the results; Z.Z., supervising theoretical calculation; X.L., designing and supervising the project, interpreting and finalizing the results, writing and finalizing the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This study is financially supported by the Strategic Priority Research Programs (B) of the Chinese Academy of Sciences (grant no. XDB18000000 and XDB42000000) and by the Program of the National Mineral Rock and Fossil Specimens Resource Center from MOST, China.

Acknowledgments

We thank Hejin Wang for his help with the powder X-ray diffraction analysis, and Xiaoli Li for his help with the electron microprobe analyses.

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Joswig, W.; Stachel, T.; Harris, J.W.; Baur, W.H.; Brey, G.P. New Ca-silicate inclusions in diamond-stracers from the lower mantle. Earth Planet. Sci. Lett. 1999, 173, 1–6. [Google Scholar] [CrossRef]
  2. Brenker, F.E.; Vincze, I.; Vekemans, B.; Nasdala, L.; Stachel, T.; Vollmer, C.; Kersten, M.; Somogyi, A.; Adams, F.; Joswig, W.; et al. Detection of a Ca-rich lithology in the Earth’s deep (>300 km) convecting mantle. Earth Planet. Sci. Lett. 2005, 236, 579–587. [Google Scholar] [CrossRef]
  3. Ringwood, A.E. Composition and Petrology of the Earth’s Mantle; McGraw-Hill: Sydney, Austrilia, 1975; pp. 1–618. ISBN 978-0070529328. [Google Scholar]
  4. Zedgenizov, D.A.; Shatskiy, A.; Ragozin, A.L.; Kagi, H.; Shatsky, V.S. Merwinite in diamond from Sao Luiz, Brazil: A new mineral of the Ca-rich mantle environment. Am. Mineral. 2014, 99, 547–550. [Google Scholar] [CrossRef]
  5. Tilley, C.E.; Vincent, H.C.G. The occurrence of an orthorhombic high-temperature form of Ca2SiO4 (bredigite) in the contact-zone and as a constituent of slags. Mineral. Mag. 1948, 28, 255–271. [Google Scholar]
  6. Douglas, A.M.B. X-ray investigation of bredigite. Mineral. Mag. 1952, 29, 875–884. [Google Scholar] [CrossRef]
  7. Gutt, W. A new calcium magnesiosilicate. Nature 1961, 190, 339–340. [Google Scholar] [CrossRef]
  8. Gutt, W. The system dicalcium silicate-merwinite. Nature 1965, 207, 184–185. [Google Scholar] [CrossRef]
  9. Schlaudt, C.M.; Roy, D.M. The join Ca2SiO4-CaMgSiO4. J. Am. Ceram. Soc. 1966, 49, 430–432. [Google Scholar] [CrossRef]
  10. Biggar, G.M. Phase relationships of bredigite (Ca5MgSi3O12) and of the quaternary compound (Ca6MgAl8SiO21) in the system CaO-MgO-Al2O3-SiO2. Cem. Concr. Res. 1971, 1, 493–513. [Google Scholar] [CrossRef]
  11. Midgley, H.G.; Bennett, M. A microprobe analysis of larnite and bredigite from Scawt Hill, Larne, Northern Ireland. Cem. Concr. Res. 1971, 1, 413–418. [Google Scholar] [CrossRef]
  12. Lin, H.C.; Foster, W.R. Stability relations of bredigite (5CaO.MgO.3SiO2). J. Am. Ceram. Soc. 1975, 58, 73. [Google Scholar] [CrossRef]
  13. Moore, P.B.; Araki, T. The crystal structure of bredigite and the genealogy of some alkaline earth orthosilicates. Am. Mineral. 1976, 61, 74–87. [Google Scholar]
  14. Sarkar, S.L.; Jeffery, J.W. Electron microprobe analysis of Scawt Hill bredigite-larnite rock. J. Am. Ceram. Soc. 1978, 61, 177–178. [Google Scholar] [CrossRef]
  15. Essene, E. The stability of bredigite and other Ca-Mg silicates. J. Am. Ceram. Soc. 1979, 63, 464–466. [Google Scholar] [CrossRef]
  16. Bridge, T.E. Progressive metamorphism of Permian siliceous limestone and dolomite-a complete sequence around a monzonite intrusion, Marble Canyon, Diablo Plateau, West Texas. In Trans Pecos Region (West Texas); Dickerson, P.W., Hoffer, J.M., Callender, J.F., Eds.; New mexico Geological Society: Socorro, NM, USA, 1980; pp. 225–229. [Google Scholar]
  17. Moseley, D.; Glasser, F.P. Identity, composition and stability of bredigite and phase T. Cem. Concr. Res. 1981, 11, 559–565. [Google Scholar] [CrossRef]
  18. Moseley, D.; Glasser, F.P. Properties and composition of bredigite-structured phases. J. Mater. Sci. 1982, 17, 2736–2740. [Google Scholar] [CrossRef]
  19. Sabine, P.A.; Styles, M.T.; Young, B.R. The nature and paragenesis of natural bredigite and associated minerals from Carneal and Scawt Hill, Co. Antrim. Mineral. Mag. 1985, 49, 663–670. [Google Scholar] [CrossRef]
  20. Xiong, Z. 2015. High Temperature and High Pressure Study about the Ca-rich Silicates in the System Ca2SiO4-Ca3MgSi2O8. Ph.D. Thesis, Peking University, Beijing, China.
  21. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  22. Wu, C.; Chang, J.; Wang, J.; Ni, S.; Zhai, W. Preparation and characteristics of a calcium magnesium silicate (bredigite) bioactive ceramic. Biomaterials 2005, 26, 2925–2931. [Google Scholar] [CrossRef]
  23. Jia, Y.; Qiao, H.; Zheng, Y.; Guo, N.; You, H. Synthesis and photoluminescence properties of Ce3+ and Eu2+-activated Ca7Mg(SiO4)4 phosphors for solid state lighting. Phys. Chem. Chem. Phys. 2017, 14, 3537–3542. [Google Scholar] [CrossRef]
  24. Holland, T.J.B.; Powell, R. An improved and extended internally consistent thermodynamic dataset for phases of petrological interest, involving a new equation of state for solids. J. Metamorph. Geol. 2011, 29, 333–383. [Google Scholar] [CrossRef]
  25. He, Q.; Liu, X.; Li, B.; Deng, L.; Chen, Z.; Liu, X.; Wang, H. Expansivity and compressibility of strontium fluorapatite and barium fluorapatite determined by in situ X-ray diffraction at high-T/P conditions: Significance of the M-site cations. Phys. Chem. Miner. 2013, 40, 29–40. [Google Scholar] [CrossRef]
  26. Liu, X.; Fleet, M.E. Phase relations of nahcolite and trona at moderate P-T conditions. J. Miner. Petrol. Sci. 2009, 104, 25–36. [Google Scholar] [CrossRef] [Green Version]
  27. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, 864–871. [Google Scholar] [CrossRef] [Green Version]
  28. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, 1133–1138. [Google Scholar] [CrossRef] [Green Version]
  29. Payne, M.C.; Teter, M.P.; Allan, D.C.; Arias, T.A.; Joannopoulos, J.D. Iterative minimization techniques for ab initio total-energy calculations: Molecular dynamics and conjugate gradients. Rev. Mod. Phys. 1992, 64, 1045–1097. [Google Scholar] [CrossRef] [Green Version]
  30. Perdew, J.; Burke, K.; Emzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  31. Lin, J.; Qteish, A.; Payne, M.; Heine, V. Optimized and transferable nonlocal separable ab initio pseudopotentials. Phys. Rev. B 1993, 47, 4147–4180. [Google Scholar] [CrossRef]
  32. Lee, M. Advanced Pseudopotentials for large scale electronic structure calculations. Ph.D. Dissertation, University of Cambridge, Cambridge, UK, 1995. [Google Scholar]
  33. Monkhorst, H.; Pack, J. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  34. Chang, L.; Liu, X.; Liu, H.; Kojitani, H.; Wang, S. Vibrational mode analysis and heat capacity calculation of K2SiSi3O9-wadeite. Phys. Chem. Miner. 2013, 40, 563–574. [Google Scholar] [CrossRef]
  35. Xiong, Z.; Liu, X.; Shieh, S.R.; Wang, S.; Chang, L.; Tang, J.; Hong, X.; Zhang, Z.; Wang, H. Some thermodynamic properties of larnite (β-Ca2SiO4) constrained by high T/P experiment and/or theoretical simulation. Am. Mineral. 2016, 101, 277–288. [Google Scholar] [CrossRef]
  36. Baroni, S.; Gironcoli, S.; Dal Corso, A.; Giannozzi, P. Phonons and related crystal properties from density-functional perturbation theory. Rev. Mod. Phys. 2001, 73, 515–562. [Google Scholar] [CrossRef] [Green Version]
  37. Refson, K.; Tulip, P.; Clark, S. Variational density-functional perturbation theory for dielectrics and lattice dynamics. Phys. Rev. B 2006, 73, 155114–1551125. [Google Scholar] [CrossRef] [Green Version]
  38. Kathlenberg, V.; Galuskina, I.; Krüger, B.; Pauluhn, A.; Galuskin, E. Structural investigations on bredigite from the Hatrurim Complex. Miner. Petrol. 2019, 113, 261–272. [Google Scholar] [CrossRef] [Green Version]
  39. Birch, F. Finite Elastic Strain of Cubic Crystals. Phys. Rev. 1947, 71, 809–824. [Google Scholar] [CrossRef]
  40. Blanco, M.A.; Francisco, E.; Luaña, V. GIBBS: Isothermal-isobaric thermodynamics of solids from energy curves using a quasi-harmonic Debye model. Comput. Phys. Commun. 2004, 158, 57–72. [Google Scholar] [CrossRef]
Figure 1. Experimental and simulated powder X-ray diffraction patterns of Ca7MgSi4O16-Bre (bredigite). Numbers represent the hkl indexes.
Figure 1. Experimental and simulated powder X-ray diffraction patterns of Ca7MgSi4O16-Bre (bredigite). Numbers represent the hkl indexes.
Crystals 11 00014 g001
Figure 2. Polyhedral views of (a) Ca7MgSi4O16-Bre structure in space group Pnnm and (b) [Mg–O–Si] chains.
Figure 2. Polyhedral views of (a) Ca7MgSi4O16-Bre structure in space group Pnnm and (b) [Mg–O–Si] chains.
Crystals 11 00014 g002
Figure 3. Polyhedral view of Bre structure in space group Pnn2.
Figure 3. Polyhedral view of Bre structure in space group Pnn2.
Crystals 11 00014 g003
Figure 4. Total energy of Ca7MgSi4O16-Bre as a function of cell volume (at 0 K). Symbols represent simulated results while curve stands for the fitted EV third-order Birch–Murnaghan equation of state (B0 = 90.6(4) GPa, B 0 = 5.7(1), and V0 = 1388.0(6) Å3).
Figure 4. Total energy of Ca7MgSi4O16-Bre as a function of cell volume (at 0 K). Symbols represent simulated results while curve stands for the fitted EV third-order Birch–Murnaghan equation of state (B0 = 90.6(4) GPa, B 0 = 5.7(1), and V0 = 1388.0(6) Å3).
Crystals 11 00014 g004
Figure 5. Phonon dispersions and vibrational density of the states (VDoS) of Ca7MgSi4O16-Bre (V = 1382.18 Å3).
Figure 5. Phonon dispersions and vibrational density of the states (VDoS) of Ca7MgSi4O16-Bre (V = 1382.18 Å3).
Crystals 11 00014 g005
Figure 6. Calculated thermal expansion coefficient of Bre.
Figure 6. Calculated thermal expansion coefficient of Bre.
Crystals 11 00014 g006
Figure 7. Isobaric and isochoric heat capacity (Cp and Cv, respectively) of Ca7MgSi4O16-Bre. Note that the Cv values remain lower than the harmonic limit of Dulong–Petit (3R/M, with R as the gas constant and M as the molar mass) for T less than 1000 K, suggesting that either the quantum effects for the ionic vibrations remain noticeable for this T range, or our Cv values have been slightly underestimated (presumably by a few percent) because the volume used in the calculation is for zero P and zero T.
Figure 7. Isobaric and isochoric heat capacity (Cp and Cv, respectively) of Ca7MgSi4O16-Bre. Note that the Cv values remain lower than the harmonic limit of Dulong–Petit (3R/M, with R as the gas constant and M as the molar mass) for T less than 1000 K, suggesting that either the quantum effects for the ionic vibrations remain noticeable for this T range, or our Cv values have been slightly underestimated (presumably by a few percent) because the volume used in the calculation is for zero P and zero T.
Crystals 11 00014 g007
Table 1. Crystal data for bredigite structure and bredigite-like structure.
Table 1. Crystal data for bredigite structure and bredigite-like structure.
Empirical Formula Ca7MgSi4O16 aCa14Mg2Si8O32 bCa6.15Ba0.3Mg1.2Mn0.35Si4O16 c
Crystal system OrthorhombicOrthorhombicOrthorhombic
Space group PnnmPnn2Pnn2
Unit cell dimensions (Å)a = 18.3434(17) a’ = 18.343(4)10.909(9)
b = 6.7313(8) b’ = 6.7313(13)18.34(1)
c = 10.8844(12) c’ = 10.884(2)6.739(9)
Volume (Å3)1344.0(3) 1344.0(5)1348(4)
Z42/
Crystal size (mm3)0.06 × 0.05 × 0.050.06 × 0.05 × 0.05/
Theta range for data collection (°)2.18 to 28.372.18 to 28.37/
Limiting indices−24 ≤ h ≤ 24−24 ≤ h ≤ 24/
−9 ≤ k ≤ 5−9 ≤ k ≤ 5/
−13 ≤ k ≤ 14−13 ≤ k ≤ 14/
Independent reflections1756 [R(int) = 0.0483]3101 [R(int) = 0.0444]/
Completeness99.50%99.50%/
Refinement methodFull-matrix least-squares on F2Full-matrix least-squares on F2/
Data/restraints/parameters1756/0/1473101/1/255/
Goodness-of-fit on F21.0031.064/
Final R indices [I > 2sigma(I)]R1 = 0.0472, wR2 = 0.1132R1 = 0.0486, wR2 = 0.1095/
R indices (all data)R1 = 0.0675, wR2 = 0.1259R1 = 0.0753, wR2 = 0.1279/
Largest diff. peak and hole (e Å−3)0.877 and −0.720 0.758 and −0.752 /
a Data refined in space group Pnnm. b Data refined in space group Pnn2. c Moore and Araki [13], original formula Ca24.6Ba1.2Mg4.8Mn1.4[SiO4]16.
Table 2. Atomic coordinates (x, y, z) and equivalent isotropic displacement parameters (U(eq)) for Ca7MgSi4O16-Bre (Pnnm).
Table 2. Atomic coordinates (x, y, z) and equivalent isotropic displacement parameters (U(eq)) for Ca7MgSi4O16-Bre (Pnnm).
Atom LabelxyzU(eq)
Ca(1)0.50000.50001.000018(1)
Ca(2)0.50000.50000.500041(1)
Ca(3)0.3328(1)0.8028(2)1.000015(1)
Ca(4)0.6733(1)0.1718(2)0.500014(1)
Ca(5)0.4119(1)0.1582(2)0.7486(1)15(1)
Ca(6)0.7726(1)0.9989(2)0.2316(1)10(1)
Mg(1)0.50001.00001.00009(1)
Mg(2)0.50001.00000.50008(1)
O(1)0.2887(2)0.2570(6)1.1220(4)14(1)
O(2)0.6968(2)0.8473(6)0.3803(4)16(1)
O(3)0.4757(2)0.7809(6)0.8606(4)16(1)
O(4)0.3389(2)0.8100(6)0.7862(4)13(1)
O(5)0.3760(3)0.4959(8)1.000014(1)
O(6)0.4115(2)0.4895(6)0.7191(4)19(1)
O(7)0.3978(3)0.1045(8)1.000018(1)
O(8)0.5764(3)0.7718(8)0.500015(1)
O(9)0.6946(3)0.5126(8)0.500018(1)
O(10)0.4444(2)0.8409(7)0.6295(4)25(1)
Si(1)0.3382(1)0.2828(3)1.00008(1)
Si(2)0.6638(1)0.7355(3)0.50009(1)
Si(3)0.4192(1)0.7194(2)0.7518(1)9(1)
Table 3. Selected bond lengths (Å) and angles (°) for Ca7MgSi4O16 (Pnnm).
Table 3. Selected bond lengths (Å) and angles (°) for Ca7MgSi4O16 (Pnnm).
Bond aBond LengthBond aBond Angle
Ca(1)–O(5) 2.275(5)O(7)#7–Mg(1)–O(7)#1180.000(1)
Ca(1)–O(5)#1 2.275(5)O(7)#7–Mg(1)–O(3)92.67(15)
Ca(1)–O(3)#2 2.465(4)O(7)#1–Mg(1)–O(3)87.33(15)
Ca(1)–O(3)#3 2.465(4)O(7)#7–Mg(1)–O(3)#1887.33(15)
Ca(1)–O(3)#1 2.465(4)O(7)#1–Mg(1)–O(3)#1892.67(15)
Ca(1)–O(3) 2.465(4)O(3)–Mg(1)–O(3)#18180.000(1)
Ca(2)–O(8)#5 2.305(5)O(7)#7–Mg(1)–O(3)#392.67(15)
Ca(2)–O(8) 2.305(5)O(7)#1–Mg(1)–O(3)#387.33(15)
Ca(2)–O(10) 2.880(4)O(3)–Mg(1)–O(3)#389.1(2)
Ca(2)–O(10)#6 2.880(4)O(3)#18–Mg(1)–O(3)#390.9(2)
Ca(2–O(10)#5 2.880(4)O(7)#7–Mg(1)–O(3)#1987.33(15)
Ca(2)–O(10)#2 2.880(4)O(7)#1–Mg(1)–O(3)#1992.67(15)
Ca(2)–O(6)#6 2.886(4)O(3)–Mg(1)–O(3)#1990.9(2)
Ca(2)–O(6)#2 2.886(4)O(3)#18–Mg(1)–O(3)#1989.1(2)
Ca(2)–O(6) 2.886(4)O(3)#3–Mg(1)–O(3)#19180.000(1)
Ca(2)–O(6)#5 2.886(4)O(10)#6–Mg(2)–O(10)#19180.0(2)
Ca(3)–O(5) 2.212(6)O(10)#6–Mg(2)–O(10)87.3(3)
Ca(3)–O(4) 2.330(4)O(10)#19–Mg(2)–O(10)92.7(3)
Ca(3)–O(4)#3 2.330(4)O(10)#6–Mg(2)–O(10)#1592.7(3)
Ca(3)–O(7)#7 2.355(6)O(10)#19–Mg(2)–O(10)#1587.3(3)
Ca(3)–O(9)#8 2.823(6)O(10)–Mg(2)–O(10)#15180
Ca(3)–O(2)#8 2.989(4)O(10)#6–Mg(2)–O(8)87.11(16)
Ca(3)–O(2)#9 2.989(4)O(10)#19–Mg(2)–O(8)92.89(16)
Ca(4)–O(9) 2.327(6)O(10)–Mg(2)–O(8)87.11(16)
Ca(4)–O(1)#10 2.544(4)O(10)#15–Mg(2)–O(8)92.89(16)
Ca(4)–O(1)#11 2.544(4)O(10)#6–Mg(2)–O(8)#1592.89(16)
Ca(4)–O(2)#4 2.580(4)O(10)#19–Mg(2)–O(8)#1587.11(16)
Ca(4)–O(2)#12 2.580(4)O(10)–Mg(2)–O(8)#1592.89(16)
Ca(4)–O(10)#5 2.579(5)O(10)#15–Mg(2)–O(8)#1587.11(16)
Ca(4)–O(10)#2 2.579(5)O(8)–Mg(2)–O(8)#15180.000(1)
Ca(5)–O(6) 2.253(4)O(5)–Si(1)–O(1)#3109.92(18)
Ca(5)–O(3)#2 2.430(4)O(5)–Si(1)–O(1)109.92(18)
Ca(5)–O(2)#5 2.439(4)O(1)#3–Si(1)–O(1)110.3(3)
Ca(5)–O(10)#4 2.568(5)O(5)–Si(1)–O(7)111.9(3)
Ca(5)–O(4)#4 2.730(4)O(1)#3–Si(1)–O(7)107.37(19)
Ca(5)–O(1)#3 2.745(4)O(1)–Si(1)–O(7)107.37(19)
Ca(5)–O(8)#5 2.7554(16)O(9)–Si(2)–O(8)119.3(3)
Ca(5)–O(7) 2.7718(15)O(9)–Si(2)–O(2)#6107.6(2)
Ca(5)–O(10)#2 2.938(5)O(8)–Si(2)–O(2)#6107.40(19)
Ca(6)–O(1)#13 2.308(4)O(9)–Si(2)–O(2)107.6(2)
Ca(6)–O(2) 2.365(4)O(8)–Si(2)–O(2)107.40(19)
Ca(6)–O(1)#14 2.377(4)O(2)#6–Si(2)–O(2)106.9(3)
Ca(6)–O(4)#15 2.424(4)
Ca(6)–O(4)#16 2.482(4)
Ca(6)–O(6)#16 2.553(4)
Ca(6)–O(9)#17 2.5931(17)
Ca(6)–O(2)#17 2.701(4)
Mg(1)–O(7)#7 2.003(5)
Mg(1)–O(7)#1 2.003(5)
Mg(1)–O(3) 2.163(4)
Mg(1)–O(3)#18 2.163(4)
Mg(1)–O(3)#3 2.163(4)
Mg(1)–O(3)#19 2.163(4)
Mg(2)–O(10)#6 2.044(4)
Mg(2)–O(10)#19 2.044(4)
Mg(2)–O(10) 2.044(4)
Mg(2)–O(10)#15 2.044(4)
Mg(2)–O(8) 2.080(6)
Mg(2)–O(8)#15 2.080(6)
Si(1)–O(5) 1.594(6)
Si(1)–O(1)#3 1.619(4)
Si(1)–O(1) 1.619(4)
Si(1)–O(7) 1.623(6)
Si(2)–O(9) 1.603(6)
Si(2)–O(8) 1.622(6)
Si(2)–O(2)#6 1.622(4)
Si(2)–O(2) 1.622(4)
Si(3)–O(3) 1.627(4)
Si(3)–O(4)1.637(4)
Si(3)–O(6) 1.594(4)
Si(3)–O(10)1.629(4)
a Symmetry transformations used to generate equivalent atoms: #1 −x + 1, −y + 1, −z + 2; #2 −x + 1, −y + 1, z; #3 x, y, −z + 2; #4 x, y − 1, z; #5 −x + 1, −y + 1, −z + 1; #6 x, y, −z + 1; #7 x, y + 1, z; #8 x − 1/2, −y + 3/2, z + 1/2; #9 x − 1/2, −y + 3/2, −z + 3/2; #10 x + 1/2, −y + 1/2, z − 1/2; #11 x + 1/2, −y + 1/2, −z + 3/2; #12 x, y − 1, −z + 1; #13 x + 1/2, −y + 3/2, −z + 3/2; #14 −x + 1, −y + 1, z − 1; #15 −x + 1, −y + 2, −z + 1; #16 x + 1/2, −y + 3/2, z − 1/2; #17 −x + 3/2, y + 1/2, −z + 1/2; #18 −x + 1, −y + 2, −z + 2; #19 −x + 1, −y + 2, z.
Table 4. Comparison between energy-optimized (0 P and 0 K) and experimentally refined (1 atm and 298 K) unit-cell parameters of the Ca7MgSi4O16-Bre (Pnnm).
Table 4. Comparison between energy-optimized (0 P and 0 K) and experimentally refined (1 atm and 298 K) unit-cell parameters of the Ca7MgSi4O16-Bre (Pnnm).
Cell ParametersExperimentalCalculatedR.D. (%) a
a18.3434(17) Ǻ18.5125 Ǻ~0.92
b6.7313(8) Ǻ6.7833 Ǻ~0.77
c10.8844(12) Ǻ11.0067 Ǻ~1.12
V1344.0(3) Ǻ31382.18 Ǻ3~2.84
a Relative difference.
Table 5. Thermal expansion coefficient (×105 K−1), and heat capacity (J mol−1 K−1) of Ca7MgSi4O16-Bre at selected T (K).
Table 5. Thermal expansion coefficient (×105 K−1), and heat capacity (J mol−1 K−1) of Ca7MgSi4O16-Bre at selected T (K).
T (K)CV (J mol−1 K−1)CP (J mol−1 K−1)αT (×105 K−1)
5077.3678.740.1048
100221.73224.490.6847
150331.46335.611.4269
200409.8415.351.9949
250467.65474.62.3858
300511.55519.92.6588
350545.31555.072.8590
400571.5582.673.0141
450592.01604.63.1394
500608.23622.253.2483
550621.19636.643.3451
600631.67648.553.4340
650640.22658.553.5179
700647.27667.053.5991
750653.15674.373.6775
800658.08680.763.7579
850662.26686.43.8385
900665.82691.443.9203
950668.89695.974.0040
1000671.54700.14.0903
Table 6. Coefficients of the CP polynomials of Ca7MgSi4O16-Bre (Pnnm).
Table 6. Coefficients of the CP polynomials of Ca7MgSi4O16-Bre (Pnnm).
T Range: 10–100 KT Range: 100–298 KT Range: 298–1000 K
k0 = 2.8(6)k0 = −1.24(2) × 102k0 = 8.22(2) × 102
k4 = −7.0(6) × 10−1k4 = 4.50(3)k1 = −3.76(6) × 103
k5 = 5.9(2) × 10−2k5 = −1.16(2) × 10−2k2 = −1.38(4) × 107
k6 = − 3.0(1) × 10−4k6 = 1.23(3) × 10−5k3 =1.61(8) × 109
aCp = k0 + k1T−0.5 + k2T−2 + k3T−3 + k4T + k5T2 + k6T3 (J mol−1 K−1).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bao, X.; He, M.; Zhang, Z.; Liu, X. Crystal Structure and Some Thermodynamic Properties of Ca7MgSi4O16-Bredigite. Crystals 2021, 11, 14. https://doi.org/10.3390/cryst11010014

AMA Style

Bao X, He M, Zhang Z, Liu X. Crystal Structure and Some Thermodynamic Properties of Ca7MgSi4O16-Bredigite. Crystals. 2021; 11(1):14. https://doi.org/10.3390/cryst11010014

Chicago/Turabian Style

Bao, Xinjian, Mingyue He, Zhigang Zhang, and Xi Liu. 2021. "Crystal Structure and Some Thermodynamic Properties of Ca7MgSi4O16-Bredigite" Crystals 11, no. 1: 14. https://doi.org/10.3390/cryst11010014

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