Next Article in Journal
Taxifolin Adsorption on Nitrogenated Graphenes: Theoretical Insights
Next Article in Special Issue
Study of the Structure of Zn and Na Borophosphate Glasses Using X-ray and Neutron Scattering Techniques
Previous Article in Journal
Plasmon Excitation in the Interaction of Slow Singly Charged Argon Ions with Magnesium
Previous Article in Special Issue
Critical Experiments and Thermodynamic Modeling of the Li2O-SiO2 System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Pressure Synthesis of the Iodide Carbonate Na5(CO3)2I

1
Department of Physics, Chemistry and Biology (IFM), Linköping University, SE-581 83 Linköping, Sweden
2
Bayerisches Geoinstitut, University of Bayreuth, 95440 Bayreuth, Germany
3
Material Physics and Technology at Extreme Conditions, Laboratory of Crystallography, University of Bayreuth, 95440 Bayreuth, Germany
4
European Synchrotron Radiation Facility, 38000 Grenoble, France
*
Author to whom correspondence should be addressed.
Solids 2024, 5(2), 333-340; https://doi.org/10.3390/solids5020022
Submission received: 23 April 2024 / Revised: 19 May 2024 / Accepted: 30 May 2024 / Published: 18 June 2024

Abstract

:
Here, we present the synthesis of a novel quaternary compound, iodide carbonate Na5(CO3)2I, at 18(1) and 25.1(5) GPa in laser-heated diamond anvil cells. Single-crystal synchrotron X-ray diffraction provides accurate structural data for Na5(CO3)2I and shows that the structure of the material can be described as built of INa8 square prisms, distorted NaO6 octahedra, and trigonal planar CO32− units. Decompression experiments show that the novel iodide carbonate is recoverable in the N2 atmosphere to ambient conditions. Our ab initio calculations agree well with the experimental structural data, provide the equation of state, and shed light on the chemical bonding and electronic properties of the new compound.

1. Introduction

Inorganic carbonates and alkali halides are two common classes of compounds that are ubiquitous in nature. Their high-pressure (HP) behavior has been intensively studied for decades. The HP synthesis of new carbonates enriched the list of C-O anions, including now not only the planar trigonal [CO3]2− group with the sp2-hybridized carbon, common for numerous carbonates known at ambient conditions, but also “the sp3 carbonates” [1] and other anions (ex. [CO4]4− and [C2O5]2−) [2,3,4]. Recent HP studies on alkali halides (such as the synthesis of a series of novel compounds including NaCl3, Na2Cl3, and Na4Cl5 [5,6]) provide examples of how the chemistry of materials dramatically changes under HP [7,8].
The successful synthesis of alkali metal halide carbonate compounds has rarely been reported, and, so far, only two alkali metals fluoride carbonates (K3CO3F and Rb3CO3F) can be found in the Inorganic Crystal Structure Database (ICSD) [9]. The characterization of complex (quaternary) compounds in a laser-heated diamond anvil cell (LHDAC) has been extremely difficult. However, recent developments in the methodology of experiments on the single-crystal X-ray diffraction (SCXRD) from multiphase microcrystalline samples, as well as in the data analysis, especially using the DAFi program [10] for data processing, have made possible the structural characterization of individual phases in complex mixtures in DACs.
In this study, we synthesized a novel iodide carbonate Na5(CO3)2I through a reaction of sodium carbonate Na2CO3 with iodine I2 at 18(1) GPa and a reaction of sodium iodide containing hygroscopic water (NaI + H2O) with carbon tetraiodide CI4 at 25.1(5) GPa after heating the mixtures to ~2000 K in DACs. The crystal structure of Na5(CO3)2I was solved and refined. Our ab initio calculations support experimental observations and reveal the electronic properties and chemical bonding of the novel compound.

2. Materials and Methods

2.1. Sample Preparation

Two BX90-type screw-driven diamond anvil cells (DACs) [11] equipped with 250 μm culet diamond anvils were used. The sample chambers were formed by the pre-indenting of steel (DAC #1) and rhenium (DAC #2) gaskets to ~25 µm thickness and laser-drilling a hole of 120 µm. DAC #1 was loaded with a plate (~5 µm thick, filling the whole sample chamber) of well-dried sodium carbonate Na2CO3 together with a piece (~10 µm in diameter and ~5 µm thick) of solid iodine. DAC #2 was loaded with one stack of a plate of NaI (2–5 µm thick, filling the whole sample chamber) and a piece of CI4. Although Na2CO3 and NaI powders were dried on a heating table at 220 °C for 48 h before loading, NaI still contained water, being very hygroscopic. The pressure was measured in situ using the first-order Raman mode of the stressed diamond anvils [12]. Double-sided laser-heating of the samples up to ~2000 K was performed in the home laboratory at the Bayerisches Geoinstitut [13]; iodine and CI4 served as laser light absorbers.

2.2. X-ray Diffraction

Synchrotron X-ray diffraction (XRD) measurements of the compressed samples were performed at ID15b (λ = 0.4100 Å, beam size ~1.5 × 1.5 μm2) of the EBS-ESRF in Grenoble, France. In order to determine the sample position for single-crystal X-ray diffraction data acquisition, full X-ray diffraction mapping of the pressure chamber was performed. The sample positions displaying the greatest number of single-crystal reflections belonging to the phases of interest were chosen, and step-scans of 0.5° from −36° to +36° ω were performed. CrysAlisPro software (version 43.67a) [14] was utilized for single-crystal data analysis. To calibrate the instrumental model in the CrysAlisPro software, i.e., the sample-to-detector distance, detector’s origin, offsets of the goniometer angles, and inclination of both the X-ray beam and detector surface with respect to the instrument axis, we used a single crystal of orthoenstatite [(Mg1.93Fe0.06)(Si1.93Al0.06)O6, Pbca space group, a = 8.8117(2) Å, b = 5.1832(10) Å, and c = 18.2391(3) Å]. The DAFi program [10] was used for the search of reflections’ groups belonging to individual single-crystal domains. The crystal structures were then solved and refined using OLEX2 (version 1.5) [15] software. The crystallite sizes were estimated from X-ray maps using XDI software (version 1.0.0.210) [16]. The crystallographic information is available in below Table 1 of Section 3.1. Powder XRD patterns during the decompression experiments were obtained using in-house XRD with a high-brilliance Rigaku diffractometer (Ag Kα radiation) (λ = 0.5609 Å) equipped with Osmic focusing X-ray optics and a Bruker Apex-II CCD detector.

2.3. Density Functional Theory Calculations

First-principles calculations were performed using the framework of density functional theory (DFT), as implemented in the Vienna Ab initio Simulation Package (VASP.5.4.4) [17]. The Projector-Augmented-Wave (PAW) method [18,19] was used to expand the electronic wave functions in a plane wave basis. The Generalized Gradient Approximation (GGA) functional was used for calculating the exchange-correlation energies, as proposed by Perdew–Burke–Ernzerhof (PBE) [20]. The PAW potentials adapted from the VASP library with the following valence configurations of 3s for Na, 5s5p for I, 2s2p for C, and 2s2p for O were used with the Na, I, C, and O POTCARs. The plane-wave kinetic energy cutoff was set to 800 eV. We performed variable cell relaxations including lattice parameters and atomic positions on the synthesized experimental structure to optimize the atomic coordinates and the cell vectors until the total forces were smaller than 10−3 eV Å−1 per atom using the conjugate-gradient (CG) algorithm. In geometry optimization, we used a Gamma-centered k-mesh of 7x7x3. To increase the accuracy of ground-state electron density and density of states (eDOS), a denser Gamma-centered k-mesh of 11x11x11 and the tetrahedron smearing method with Blöchl corrections (ISMEAR = −5) were used. The crystal structure and the ELF visualization were made with VESTA software (version 3.5.7) [21]. The finite displacement method, as implemented in PHONOPY [22], was used to calculate harmonic phonon frequencies and phonon band structures. A supercell size of 2 × 2 × 2 was used with a k-mesh size of 2 × 2 × 2 for the harmonic phonon calculations at 0 K.

3. Results

3.1. Crystal Structure of the Iodide Carbonate Na5(CO3)2I

The structure of iodide carbonate Na5(CO3)2I synthesized at 18(1) and 25.1(5) GPa has the tetragonal space group I41/amd (#141) (Figure 1). This structure type has not been known before. Full crystallographic data and refinement details are provided in Table 1, and the CIFs are deposited at CSD 2348295 and 2348296. The lattice parameters are a = 6.4543(17) Å and c = 14.638(6) Å at 18(1) GPa. In the structure of Na5(CO3)2I, iodine atoms are in the 4b Wyckoff site, Na1 and Na2 atoms are in the 4a and 16f sites, O1 and O2 atoms are in the 8e and 16h sites, and C atoms are in the 8e site. Each I atom is surrounded by eight Na2 atoms forming an INa8 square prism with a height of ~3.66 Å and a basis length of ~3.26 Å at 18(1) GPa. (Figure 1). The prisms are linked through common edges forming the 3D framework, with rectangular channels running along the a and b directions (Figure 1a). Sodium Na1 atoms and CO3 groups are located in the channels (Figure 1b). The Na1 atoms are coordinated by six O atoms (two O1 and four O2) forming a distorted octahedron. These distorted octahedra are connected to each other through oxygen atoms of planar trigonal CO32− units (Figure 1c). At 18(1) GPa, the Na-I distance in the INa8 square prism is 2.9436(16) Å, which is compatible with the distances in the known TlI-type sodium iodide NaI (~2.82–2.95 Å) at 31 GPa [23]). As experimental structural data of Na2CO3 at high pressures are absent, one could only compare the average Na-O distances in the NaO6 octahedra of Na5(CO3)2I (2.150(8) Å at 18(1) GPa) with the γ-Na2CO3 at ambient (~2.35 Å) [24] or the P63/mcm-Na2CO3 at 20 GPa (~2.24 Å) from theoretical calculations [25].
Table 1. Crystal structure, data collection, and refinement details of Na5(CO3)2I at 18(1) and 25.1(5) GPa in comparison to the corresponding DFT-relaxed structure.
Table 1. Crystal structure, data collection, and refinement details of Na5(CO3)2I at 18(1) and 25.1(5) GPa in comparison to the corresponding DFT-relaxed structure.
Crystal DataDFT Results
Chemical formulaNa5(CO3)2INa5(CO3)2INa5(CO3)2I
Mr361.87361.87
Crystal system, space groupTetragonal, I41/amdTetragonal, I41/amdTetragonal, I41/amd
Temperature (K)298298
Pressure (GPa)18(1)25.1(5)21.5
a, c (Å)6.4543(17), 14.638(6)6.4154(9), 14.504(4)6.4162, 14.5214
V3)609.8(4)597.0(2)597.8089
Z44
Radiation typeSynchrotron, λ = 0.4100 ÅSynchrotron, λ = 0.4100 Å
μ (mm−1)1.281.31
Crystal size (mm)0.003 × 0.003 × 0.0030.003 × 0.003 × 0.003
Data Collection
Absorption correctionMulti-scanMulti-scan
No. of measured, independent, and
observed [I > 2σ(I)] reflections
757, 253, 215 458, 268, 192
Rint0.0460.033
(sin θ/λ)max−1)0.7130.885
Refinement
R[F2 > 2σ(F2)], wR(F2), S0.050, 0.130, 1.070.073, 0.195, 0.98
No. of reflections253268
No. of parameters2424
Δρmax, Δρmin (e Å−3)1.33, −1.692.98, −1.90
Crystal Structure
Wyckoff site, fractional atomic coordinates (x y z)I1: 4b, (0 1/4 3/8)
Na1: 4a, (0 3/4 1/8)
Na2: 16f, (0.2448(4) 0 0)
O1: 8e, (0 1/4 0.0213(5))
O2: 16h, (0 0.0788(10) 0.1500(4))
C1: 8e, (0 1/4 0.1092(8))
I1: 4b, (0 1/4 3/8)
Na1: 4a, (0 3/4 1/8)
Na2: 16f, (0.2454(6) 0 0)
O1: 8e, (0 1/4 0.0184(8))
O2: 16h, (0 0.0787(14) 0.1509(6))
C1: 8e, (0 1/4 0.1093(10))
I1: 4b, (0 1/4 3/8)
Na1: 4a, (0 3/4 1/8)
Na2: 16f, (0.2430 0 0)
O1: 8e, (0 1/4 0.0203)
O2: 16h, (0 0.0758 0.1517)
C1: 8e, (0 1/4 0.1093)
Uiso2)I1: 0.0189(4)
Na1: 0.0156(13)
Na2: 0.0199(9)
O1: 0.0192(16)
O2: 0.0219(11)
C1: 0.0139(19)
I1: 0.0163(5)
Na1: 0.0169(18)
Na2: 0.0172(11)
O1: 0.017(2)
O2: 0.0184(15)
C1: 0.011(2)
The C-O distances in the triangular CO32− unit are different: at 18(1) GPa, the C-O1 distance is 1.287(14) Å, whereas the two C-O2 bonds are 1.256(9) Å by, that is by ~0.03 Å shorter (black numbers in Figure 1c). At 25.1(5) Gpa, the C-O1 and C-O2 are 1.318(19) and 1.254(12) Å, respectively (red numbers in Figure 1c), giving ~0.06 Å difference due to the shortening of the C-O1 bonds oriented along the c direction. The length of the C-O2 bond is practically unaffected. Bond angles in the CO32− unit are slightly off from the ideal sp2 hybridization angle of 120 degrees (O1-C-O2 vs. O2-C-O2 are 118.4(6) vs. 123.2(11) and 118.8(8) vs. 122.5(15) at 18(1) and 25.1(5) GPa, respectively). Considering this distortion, the planar trigonal CO32− unit in Na5(CO3)2I does not possess the equally delocalized π-bond between C and three O atoms (unlike the CO32− unit in γ-Na2CO3, whose distortion at 1 bar is insignificant; average C-O1 vs. C-O2 are 1.281(3) vs. 1.283(2) Å, giving ~0.002 Å difference [24]).

3.2. Stability and Decompression Behavior of Na5(CO3)2I

The DFT calculations reproduced the crystal structure of the Na5(CO3)2I compound in good agreement with the experiment (Table 1). The formation enthalpy (∆H) was calculated via the synthesis route of NaI + 2Na2CO3 → Na5(CO3)2I. The values of ∆H are equal to −0.217 eV/f.u. at 20 GPa (B1-type NaI [23] and the theoretically predicted P63/mcm-Na2CO3 [25] were adopted) and 0.130 eV/f.u. at 1 bar (average structure of the γ-Na2CO3 [24] and B1-type NaI were adopted). The latter value corresponds to 26 meV/atom, which is well below the “standard” 70 meV/atom limit for metastability [26]. Calculated phonon modes of Na5(CO3)2I revealed that the compound is dynamically stable both at the experimental volume of ~597 Å3 (corresponding to theoretical pressure of ~21.5 GPa) (Figure 2a) and at 1 bar (Figure 2b). These results indicate that Na5(CO3)2I, though metastable at 1 bar, is expected to be quenchable to ambient conditions.
DAC #1 was fully decompressed from 18(1) GPa and opened in the glove bag in an N2 atmosphere. Then, it was closed, and the pressure after closing was 4(3) GPa. A powder XRD pattern from the recovered sample was measured in-house (see Materials and Methods). It shows that the Na5(CO3)2I was still preserved after the decompression (Figure 3a,b). The pressure dependence of the volume per atom for Na5(CO3)2I (based on the pressure–volume relations from our DFT calculations) is shown in Figure 3c along with the experimental data points. Note that the experimental data point obtained at 4(3) GPa (black circle in Figure 3c) was a result of the Le Bail fit of the powder pattern. The measured experimental volumes from SCXRD (red dots in Figure 3c) matched with the theoretical calculations, and the bulk modulus of K0 = 43.6(3) GPa (V0 = 13.793(9) Å3/atom, K’ = 5.10(2)) was determined from the third-order Birch–Murnaghan equation of state (3BM EOS).

3.3. Electronic Properties of Na5(CO3)2I

In order to analyze the electronic properties and chemical bonding of the Na5(CO3)2I compound, we performed DFT calculations of the electron localization function (ELF), band structure, and total and projected electron densities of states (TDOS and PDOS) (Figure 4).
The ELF calculated in the (100) plane of the Na5(CO3)2I compound confirms that the I-Na2 and Na1-O interactions (Figure 1c) are closed-shell interactions (ionic), and the C-O form covalent bonds by sharing electrons (Figure 4a). The calculated band structure using the PBE GGA functional shows that Na5(CO3)2I has a wide bandgap of 5.447 eV at 21.5 GPa (corresponding to the experimental volume of ~597 Å3) (Figure 4b), which means the insulator behavior of Na5(CO3)2I. The electronic states are highly localized and the valence band maximum (VBM) occurs between M and S high symmetric points, and the conducting band minimum (CBM) is located between S0 and Γ (Figure 4b). In the eDOS of the Na5(CO3)2I compound, the I-5p and O-2p orbitals contribute to the highest energy of VBM, while O-2p and C-2p contribute to the lowest energy of CBM (Figure 4c), which is similar to the case of γ-Na2CO3 [27] (O-2p contribute to the highest energy of VBM and O-2p and C-2p contribute to the lowest energy of CBM). The bandgap of Na5(CO3)2I is relatively large if one compares it with the bandgap of γ-Na2CO3 (3.69 eV using PBE GGA [28], and 3.94 eV using PW91 GGA [27]), or the B1-NaI (3.9 eV using PBE GGA [29]) at 1 bar.

4. Conclusions

To summarize, we synthesized a novel iodide carbonate Na5(CO3)2I at 18(1) and 25.1(5) GPa. The structures were solved and refined in situ using HP synchrotron SCXRD in LHDACs. On the basis of the structural data and the results of ab initio calculations, we revealed the chemical bonding, dynamical stability, equation of state, and electronic properties of the new compound.

Author Contributions

Methodology, Y.Y., L.D., I.A.A., A.A. (Andrey Aslandukov), A.A. (Alena Aslandukova), F.I.A., W.Z. and M.H.; formal analysis, Y.Y. and L.D.; resources, L.D., I.A.A. and N.D.; writing—original draft preparation, Y.Y.; writing—review and editing, L.D., N.D., I.A.A. and A.A. (Andrey Aslandukov); supervision, L.D., N.D. and I.A.A.; funding acquisition, L.D., I.A.A. and N.D. All authors have read and agreed to the published version of the manuscript.

Funding

Y.Y. and I.A.A. acknowledge support from the Knut and Alice Wallenberg Foundation (Wallenberg Scholar grant no. KAW-2018.0194) and the Swedish Research Council (VR, grant no. 2023-05358). N.D. and L.D. thank the Deutsche Forschungsgemeinschaft (DFG projects DU 954–11/1, DU 393–9/2, DU 393–13/1, and DU 945/15-1) for financial support. N.D. and I.A.A. also thank the Swedish Government Strategic Research Area in Materials Science on Functional Materials at Linköping University (faculty grant SFO-Mat-LiU no. 2009 00971). DFT calculations were enabled by resources provided by the National Academic Infrastructure for Supercomputing in Sweden (NAISS) at the National Supercomputer Center, partially funded by the Swedish Research Council through grant agreement no. 2022-06725.

Data Availability Statement

The datasets presented in this study can be found in online repositories. The names of the repository/repositories and accession number(s) can be found below (accessed on 4 June 2024): https://www.ccdc.cam.ac.uk/structures/-, 2348295 and 2348296.

Acknowledgments

The authors acknowledge the European Synchrotron Radiation Facility (ESRF) for the provision of beamtime at the ID15b beamline. For the purpose of open access, the author has applied a Creative Commons Attribution (CC BY) license to any Author Accepted Manuscript version arising from this submission.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Lobanov, S.S.; Dong, X.; Martirosyan, N.S.; Samtsevich, A.I.; Stevanovic, V.; Gavryushkin, P.N.; Litasov, K.D.; Greenberg, E.; Prakapenka, V.B.; Oganov, A.R.; et al. Raman spectroscopy and x-ray diffraction of sp3 CaCO3 at lower mantle pressures. Phys. Rev. B 2017, 96, 104101. [Google Scholar] [CrossRef]
  2. Spahr, D.; Binck, J.; Bayarjargal, L.; Luchitskaia, R.; Morgenroth, W.; Comboni, D.; Milman, V.; Winkler, B. Tetrahedrally Coordinated sp3-Hybridized Carbon in Sr2CO4 Orthocarbonate at Ambient Conditions. Inorg. Chem. 2021, 60, 5419–5422. [Google Scholar] [CrossRef] [PubMed]
  3. Spahr, D.; Konig, J.; Bayarjargal, L.; Milman, V.; Perlov, A.; Liermann, H.P.; Winkler, B. Sr[C2O5] is an Inorganic Pyrocarbonate Salt with [C2O5]2− Complex Anions. J. Am. Chem. Soc. 2022, 144, 2899–2904. [Google Scholar] [CrossRef] [PubMed]
  4. Spahr, D.; Konig, J.; Bayarjargal, L.; Luchitskaia, R.; Milman, V.; Perlov, A.; Liermann, H.P.; Winkler, B. Synthesis and Structure of Pb[C2O5]: An Inorganic Pyrocarbonate Salt. Inorg. Chem. 2022, 61, 9855–9859. [Google Scholar] [CrossRef] [PubMed]
  5. Yin, Y.; Aslandukova, A.; Jena, N.; Trybel, F.; Abrikosov, I.A.; Winkler, B.; Khandarkhaeva, S.; Fedotenko, T.; Bykova, E.; Laniel, D.; et al. Unraveling the Bonding Complexity of Polyhalogen Anions: High-Pressure Synthesis of Unpredicted Sodium Chlorides Na2Cl3 and Na4Cl5 and Bromide Na4Br5. JACS Au 2023, 3, 1634–1641. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, W.; Oganov, A.R.; Goncharov, A.F.; Zhu, Q.; Boulfelfel, S.E.; Lyakhov, A.O.; Stavrou, E.; Somayazulu, M.; Prakapenka, V.B.; Konopkova, Z. Unexpected stable stoichiometries of sodium chlorides. Science 2013, 342, 1502–1505. [Google Scholar] [CrossRef] [PubMed]
  7. Zhang, L.; Wang, Y.; Lv, J.; Ma, Y. Materials discovery at high pressures. Nat. Rev. Mater. 2017, 2, 17005. [Google Scholar] [CrossRef]
  8. Miao, M.S.; Sun, Y.H.; Zurek, E.; Lin, H.Q. Chemistry under high pressure. Nat. Rev. Chem. 2020, 4, 508–527. [Google Scholar] [CrossRef]
  9. ICSD Database. Available online: https://icsd.products.fiz-karlsruhe.de.
  10. Aslandukov, A.; Aslandukov, M.; Dubrovinskaia, N.; Dubrovinsky, L. Domain Auto Finder (DAFi) program: The analysis of single-crystal X-ray diffraction data from polycrystalline samples. J. Appl. Crystallogr. 2022, 55, 1383–1391. [Google Scholar] [CrossRef]
  11. Kantor, I.; Prakapenka, V.; Kantor, A.; Dera, P.; Kurnosov, A.; Sinogeikin, S.; Dubrovinskaia, N.; Dubrovinsky, L. BX90: A new diamond anvil cell design for X-ray diffraction and optical measurements. Rev. Sci. Instrum. 2012, 83, 125102. [Google Scholar] [CrossRef]
  12. Akahama, Y.; Kawamura, H. Pressure calibration of diamond anvil Raman gauge to 310 GPa. J. Appl. Phys. 2006, 100, 043516. [Google Scholar] [CrossRef]
  13. Fedotenko, T.; Dubrovinsky, L.; Aprilis, G.; Koemets, E.; Snigirev, A.; Snigireva, I.; Barannikov, A.; Ershov, P.; Cova, F.; Hanfland, M. Laser heating setup for diamond anvil cells for in situ synchrotron and in house high and ultra-high pressure studies. Rev. Sci. Instrum. 2019, 90, 104501. [Google Scholar] [CrossRef]
  14. Rigaku OD and CryAlis PRO; Rigaku Oxford Diffraction Ltd.: Abingdon, UK, 2018.
  15. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  16. Hrubiak, R.; Smith, J.S.; Shen, G. Multimode scanning X-ray diffraction microscopy for diamond anvil cell experiments. Rev. Sci. Instrum. 2019, 90, 025109. [Google Scholar] [CrossRef] [PubMed]
  17. Kresse, G.; Furthmuller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  18. Blochl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed]
  19. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  20. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  21. Momma, K.; Izumi, F. VESTA: A three-dimensional visualization system for electronic and structural analysis. J. Appl. Crystallogr. 2008, 41, 653–658. [Google Scholar] [CrossRef]
  22. Togo, A.; Oba, F.; Tanaka, I. First-principles calculations of the ferroelastic transition between rutile-type and CaCl2-type SiO2 at high pressures. Phys. Rev. B 2008, 78, 134106. [Google Scholar] [CrossRef]
  23. Léger, J.M.; Haines, J.; Danneels, C.; Oliveira, L.S.d. The TlI-type structure of the high-pressure phase of NaBr and NaI; pressure-volume behaviour to 40 GPa. J. Phys. Condens. Matter 1998, 10, 4201–4210. [Google Scholar] [CrossRef]
  24. Dusek, M.; Chapuis, G.; Meyer, M.; Petricek, V. Sodium carbonate revisited. Acta Crystallogr. B 2003, 59, 337–352. [Google Scholar] [CrossRef] [PubMed]
  25. Gavryushkin, P.N.; Behtenova, A.; Popov, Z.I.; Bakakin, V.V.; Likhacheva, A.Y.; Litasov, K.D.; Gavryushkin, A. Toward Analysis of Structural Changes Common for Alkaline Carbonates and Binary Compounds: Prediction of High-Pressure Structures of Li2CO3, Na2CO3, and K2CO3. Cryst. Growth Des. 2016, 16, 5612–5617. [Google Scholar] [CrossRef]
  26. Sun, W.; Dacek, S.T.; Ong, S.P.; Hautier, G.; Jain, A.; Richards, W.D.; Gamst, A.C.; Persson, K.A.; Ceder, G. The thermodynamic scale of inorganic crystalline metastability. Sci. Adv. 2016, 2, e1600225. [Google Scholar] [CrossRef] [PubMed]
  27. Duan, Y. A first-principles density functional theory study of the electronic structural and thermodynamic properties of M2ZrO3 and M2CO3 (M = Na, K) and their capabilities for CO2 capture. J. Renew. Sustain. Energy 2012, 4, 013109. [Google Scholar] [CrossRef]
  28. Cai, W.; Chen, J.; Pan, S.; Yang, Z. Enhancement of band gap and birefringence induced via π-conjugated chromophore with “tail effect”. Inorg. Chem. Front. 2022, 9, 1224–1232. [Google Scholar] [CrossRef]
  29. Kesek, M.; Kurt, K. First-principles calculations to investigate structural, electronic and phonon properties of sodium bromide (NaBr) and sodium iodide (NaI) crystals. Comput. Condens. Matter. 2022, 31, e00682. [Google Scholar] [CrossRef]
Figure 1. The structure of the novel iodide carbonate Na5(CO3)2I compound synthesized in this work: (a) 3D framework of the INa8 cubes (carbon and oxygen atoms are not shown); (b) unit cell (I, Na, O, and C atoms are shown in purple, yellow, red, and brown, respectively); (c) building blocks of the structure: INa8 and Na1O6 polyhedra and planar CO3 groups. The interatomic distances (in Å) in the trigonal CO32− at 18(1) (black numbers) and 25.1(5) GPa (red numbers) are notated.
Figure 1. The structure of the novel iodide carbonate Na5(CO3)2I compound synthesized in this work: (a) 3D framework of the INa8 cubes (carbon and oxygen atoms are not shown); (b) unit cell (I, Na, O, and C atoms are shown in purple, yellow, red, and brown, respectively); (c) building blocks of the structure: INa8 and Na1O6 polyhedra and planar CO3 groups. The interatomic distances (in Å) in the trigonal CO32− at 18(1) (black numbers) and 25.1(5) GPa (red numbers) are notated.
Solids 05 00022 g001
Figure 2. Phonon dispersion curves calculated at (a) experimental volume of ~597 Å3, which corresponds to theoretical pressure at ~21.5 GPa, and (b) 1 bar along high-symmetry directions in the Brillouin zone.
Figure 2. Phonon dispersion curves calculated at (a) experimental volume of ~597 Å3, which corresponds to theoretical pressure at ~21.5 GPa, and (b) 1 bar along high-symmetry directions in the Brillouin zone.
Solids 05 00022 g002
Figure 3. Some experimental and calculation data for Na5(CO3)2I. (a) Powder XRD pattern collected from the recovered sample of Na5(CO3)2I at 4(3) GPa compared with that at the previous pressure point of 18(1) Gpa. Miller indices for Na5(CO3)2I are notated with black numbers; “G” refers to the gasket material. (b) Le Bail fit of the powder XRD pattern of Na5(CO3)2I at 4(3) GPa; vertical ticks correspond to Bragg peaks of Na5(CO3)2I. X-ray wavelength λ = 0.5609Å. (c) The pressure dependence of the volume per atom based on the pressure–volume relations from our DFT calculations. The dashed line represents DFT-calculated volume for given pressures fitted by the 3BM EOS (V0 = 13.793(9) Å3/atom, K0 = 43.6(3) GPa, K’ = 5.10(2)). Black circles (powder XRD) and red dots (SCXRD) represent our experimental data points.
Figure 3. Some experimental and calculation data for Na5(CO3)2I. (a) Powder XRD pattern collected from the recovered sample of Na5(CO3)2I at 4(3) GPa compared with that at the previous pressure point of 18(1) Gpa. Miller indices for Na5(CO3)2I are notated with black numbers; “G” refers to the gasket material. (b) Le Bail fit of the powder XRD pattern of Na5(CO3)2I at 4(3) GPa; vertical ticks correspond to Bragg peaks of Na5(CO3)2I. X-ray wavelength λ = 0.5609Å. (c) The pressure dependence of the volume per atom based on the pressure–volume relations from our DFT calculations. The dashed line represents DFT-calculated volume for given pressures fitted by the 3BM EOS (V0 = 13.793(9) Å3/atom, K0 = 43.6(3) GPa, K’ = 5.10(2)). Black circles (powder XRD) and red dots (SCXRD) represent our experimental data points.
Solids 05 00022 g003
Figure 4. Calculated properties of Na5(CO3)2I at 21.5 GPa (experimental volume of ~597 Å3). (a) ELF calculated in the (100) plane, (b) band structure, and (c) TDOS and PDOS curves of Na5(CO3)2I. I, Na, O, and C atoms are shown in purple, yellow, red, and brown, respectively. The horizontal (in (b)) and vertical (in (c)) dashed black lines indicate the Fermi energy.
Figure 4. Calculated properties of Na5(CO3)2I at 21.5 GPa (experimental volume of ~597 Å3). (a) ELF calculated in the (100) plane, (b) band structure, and (c) TDOS and PDOS curves of Na5(CO3)2I. I, Na, O, and C atoms are shown in purple, yellow, red, and brown, respectively. The horizontal (in (b)) and vertical (in (c)) dashed black lines indicate the Fermi energy.
Solids 05 00022 g004
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

Yin, Y.; Dubrovinsky, L.; Aslandukov, A.; Aslandukova, A.; Akbar, F.I.; Zhou, W.; Hanfland, M.; Abrikosov, I.A.; Dubrovinskaia, N. High-Pressure Synthesis of the Iodide Carbonate Na5(CO3)2I. Solids 2024, 5, 333-340. https://doi.org/10.3390/solids5020022

AMA Style

Yin Y, Dubrovinsky L, Aslandukov A, Aslandukova A, Akbar FI, Zhou W, Hanfland M, Abrikosov IA, Dubrovinskaia N. High-Pressure Synthesis of the Iodide Carbonate Na5(CO3)2I. Solids. 2024; 5(2):333-340. https://doi.org/10.3390/solids5020022

Chicago/Turabian Style

Yin, Yuqing, Leonid Dubrovinsky, Andrey Aslandukov, Alena Aslandukova, Fariia Iasmin Akbar, Wenju Zhou, Michael Hanfland, Igor A. Abrikosov, and Natalia Dubrovinskaia. 2024. "High-Pressure Synthesis of the Iodide Carbonate Na5(CO3)2I" Solids 5, no. 2: 333-340. https://doi.org/10.3390/solids5020022

Article Metrics

Back to TopTop