Noncentrosymmetric Na - American Chemical Society

Oct 11, 2017 - Novosibirsk State University, 2 Pirogova Street, 630090 Novosibirsk ... Chemistry SB RAS, 3 Lavrentieva Avenue, 630090 Novosibirsk, Rus...
0 downloads 0 Views 2MB Size
Article Cite This: Cryst. Growth Des. 2017, 17, 6079-6084

pubs.acs.org/crystal

Noncentrosymmetric Na2Ca4(CO3)5 Carbonate of “M13M23XY3Z” Structural Type and Affinity between Borate and Carbonate Structures for Design of New Optical Materials Sergey V. Rashchenko,*,†,‡ Vladimir V. Bakakin,§ Anton F. Shatskiy,† Pavel N. Gavryushkin,†,‡ Yurii V. Seryotkin,†,‡ and Konstantin D. Litasov†,‡ †

Sobolev Institute of Geology and Mineralogy SB RAS, 3 Koptyuga Avenue, 630090 Novosibirsk, Russia Novosibirsk State University, 2 Pirogova Street, 630090 Novosibirsk, Russia § Nikolaev Institute of Inorganic Chemistry SB RAS, 3 Lavrentieva Avenue, 630090 Novosibirsk, Russia ‡

ABSTRACT: A number of carbonates with promising optical properties were synthesized during the past decade, proving the potential of carbonate systems as a source of novel optical materials. Here we report a new noncentrosymmetric Na2Ca4(CO3)5 carbonate (P63mc; a = 10.37402(14) Å, c = 6.25935(9) Å) with “M13M23XY3Z” structural type. This structural type, built upon an M13M23 cationic array with three types of anion-filled voids (X, Y, and Z), can accommodate a variety of cations (alkali, alkaline earth, and rare earth) and anions (carbonate, borate, and halide), as we show by extensive analysis of structures of known compounds, and appears to be very promising for the search of new optical materials.

1. INTRODUCTION An increasing demand in nonlinear optical (NLO) materials for UV applications has been stimulating an intensive search of noncentrosymmetric compounds with high coefficients of second harmonic generation (SHG) and wide transparency range. Since Chuangtian Chen’s early works on anionic group theory, borates have been regarded as the most promising class of materials for NLO applications owing to geometry of planar (BO3)3− triangles and their polyanions, favorable for SHG effect. As a result, numerous NLO borates with good performance such as β-BaB2O4 (BBO), LiB3O5 (LBO), CsB3O5 (CBO), LiCsB6O10 (CLBO), K2Al2B2O7 (KABO), KBe2BO3F2 (KBBF), and so on have been discovered−see Chen et al.1 and references therein. On the other hand, such an extensive search has significantly exhausted the potential of borate systems, so that the discovery of new NLO borates is becoming more and more difficult and requires work with complex systems and/or exotic synthesis conditions. An alternative, but much less studied class of compounds, from the perspective of NLO applications following the anionic group theory, are carbonates. Among them, alkali−alkaline earth carbonates represent a particular interest due to their transparency in the UV region.2 Recent studies revealed a number of promising carbonate materials such as AMCO3F (A = K, Rb, Cs; M = Ca, Sr, Ba), 3 Ca 2 Na 3 (CO 3 ) 3 F, 4 CsNa5Ca5(CO3)8, Na4La2(CO3)5,5 Na3RE(CO3)3 (RE = Y, Gd),6 Na8Lu2(CO3)6F2, and Na3Lu(CO3)2F2.7 However, the Na2CO3−CaCO3 system, important both in materials science © 2017 American Chemical Society

as a source of new NLO materials and in Earth and environmental science as a potential reservoir of carbon dioxide, has been remaining far from completely understood. The two best known intermediate compounds of the Na2CO3−CaCO3 system are Na2Ca(CO3)2 and Na2Ca2(CO3)3, first described half a century ago as nyerereite8 and shortite9 minerals, respectively. The syntheses of Na2Ca(CO3) 2 and Na 2Ca 2(CO3)3 were recently reported by Gavryushkin et al.10 and Song et al.,2 who also pointed out perfect optical properties of these materials (both carbonates are UV transparent, and noncentrosymmetric Na2Ca2(CO3)3 has 3×KDP SHG coefficient). In the same work2 authors for the first time reported the third intermediate compound: noncentrosymmetric Na6Ca5(CO3)8 carbonate with UV cutoff edge shorter than 200 nm and SHG coefficient of 1×KDP. The recent high-pressure reconnaissance investigations of Shatskiy et al.11 have revealed three more intermediate compounds of the Na2CO3−CaCO3 system: Na4Ca(CO3)3, Na2Ca3(CO3)4, and Na2Ca4(CO3)5 (Table 1). Surprisingly, the crystal structure of Na2Ca3(CO3)4 solved then by Gavryushkin et al.12 was found to be homeotypic with M3RE2(BO3)4 borates (M = Ca, Sr, Ba; RE = La, Pr, Nd, Sm, Eu, Gd, Dy, Ho, Er, Y, Bi), known as promising optical materials. The latter demonstrates that the search for NLO materials in carbonate Received: August 19, 2017 Revised: September 18, 2017 Published: October 11, 2017 6079

DOI: 10.1021/acs.cgd.7b01161 Cryst. Growth Des. 2017, 17, 6079−6084

Crystal Growth & Design

Article

Table 1. Alkali−Alkaline Earth Carbonates in the Na2CO3− CaCO3 System phase Na4Ca(CO3)311 Na2Ca(CO3)22,10

Na6Ca5(CO3)82

Na2Ca2(CO3)32

Na2Ca3(CO3)411,12

Na2Ca4(CO3)5, this work

structure not solved Pbca (or P21ca) a = 10.0713(5) Å b = 8.7220(2) Å c = 12.2460(4) Å P63mc a = 10.0708(19) Å c = 12.604(4) Å Amm2 a = 4.9720(9) Å b = 11.068(3) Å c = 7.1271(14) Å Pn a = 31.4421(8) Å b = 8.1960(2) Å c = 7.4360(2) Å β = 89.923(2)° P63mc a = 10.37402(14) Å c = 6.25935(9) Å

natural analogue − nyerereite13

zemkorite (?);14 structure not solved

Figure 1. BSE images of the Na2Ca4(CO3)5 sample cross section. shortite9

Olympus BX41 microscope with backscattering geometry was used with a long working distance 50× objective (numerical aperture of 0.50). The spectral resolution was set to ∼3.0 cm−1 at a Raman shift of 1300 cm−1. This resolution was achieved by using a grating with 1800 grooves/mm and equivalent 150 μm slits and pinhole. Such resolution provides peak positions accurate to about 0.5 cm−1. 2.4. Structure Solution. A high-quality single crystal was selected from the crushed piece of synthesized sample using a polarizing microscope for single-crystal X-ray diffraction. The data collection was performed at STOE IPDS-2T diffractometer, equipped with a Mo source (graphite monochromator) and IP detector. The collected data were handled in CrysAlisPro software using the ESPERANTO protocol.17 SHELX software18 in WinGX suite19 was used for the structure solution and refinement. In the solved structure, four Na+ and eight Ca2+ cations of the unit cell occupy two positions, M1 (6 per unit cell) and M2 (6 per unit cell), which implies mixed occupation for at least one of them. To reveal the disordering scheme, we used a linear restraint (SUMP SHELXL instruction) in the first step of structure refinement to distribute a stoichiometric amount of Na+ cations between M1 and M2 sites. As a result, the refinement converged with the M1 site occupied by (2/3)Na+ and (1/3)Ca2+, and M2 occupied by Ca2+ only. Details of the data collection and structure refinement are summarized in Table 2. Atomic coordinates and equivalent isotropic displacement parameters are listed in Table 3 (see also the supplementary crystallographic data).



burbankite15

systems may be guided by the principles of comparative crystal chemistry instead of “blind search”, and a variety of known borate materials may be used as a basis for structural design of new carbonate ones. To confirm this thesis we report here the high-pressure growth and structural investigation of a new noncentrosymmetric Na2Ca4(CO3)5 carbonate, whose structure links together a number of alkali−alkaline earth−rare earth carbonate minerals and recently described carbonate and borate optical materials.

2. EXPERIMENTAL SECTION 2.1. Synthesis. The Na2Ca4(CO3)5 sample was synthesized from a stoichiometric mixture of pure Na2CO3 and CaCO3 (Wako) in a Kawai-type multianvil apparatus at Tohoku University (Sendai, Japan). The high-pressure cell assemblage and experimental details are described in detail in Shatskiy et al.16 The mixture was ground in an agate mortar under acetone and loaded into a graphite capsule, which then was dried at 300 °C for 3 h and stored in a drying oven prior to experiment. During the synthesis, the sample was compressed to a pressure of 6 GPa at room temperature, heated to 1050 °C for 19 h, and then quenched by cutting off the electrical power of the heater, followed by slow decompression. The recovered sample was mounted into epoxy and polished using anhydrous lubricants for chemical analysis and Raman spectroscopy; a piece of the sample was also used for single-crystal X-ray diffraction (see below). 2.2. Chemical Analysis. The chemical composition of the sample was checked using a JSM 5410 scanning electron microscope equipped with Oxford Instruments Link ISIS series 300 energy-dispersive X-ray spectrometer (EDS). EDS spectra were collected by rastering the electron beam over a surface area available for the analysis with linear dimensions from 10 to 300 μm at 15 kV accelerating voltage and 1 nA load current. The obtained data correspond to 20.0(2) mol % of Na2CO3 and 80.0(2) mol % CaCO3 in the sample, proving the Na2Ca4(CO3)5 stoichiometry. The rastering of the beam to wide area allows careful analysis of alkalis, which can be lost by applying an electron beam; the technique of analysis is described in detail in Shatsky et al.16 A BSE image of the studied sample cross section is shown in Figure 1 to prove the sample homogeneity. 2.3. Raman Spectroscopy. Raman spectra were collected on a LabRAM HR800 spectrometer (HORIBA Jobin Yvon) with a 1024 pixel CCD detector using the 514.5 nm argon laser (Melles Griot). An

3. RESULTS AND DISCUSSION 3.1. Crystal Structure of Na2Ca4(CO3)5. The crystal structure of Na2Ca4(CO3)5 contains two independent cation Table 2. Crystal Data and Structure Refinement for Na2Ca4(CO3)5 formula wt wavelength space group unit cell dimens vol density (calcd) abs coeffic cryst size theta range for data collection index ranges reflns collected indep reflns refinement meth data/restraints/params goodness-of-fit on F2 final R indices (I > 2σ) R indices (all data) absolute structure param largest diff peak and hole 6080

506.35 (Z = 2) 0.71073 Å P63mc a = 10.37402(14) Å; c = 6.25935(9) Å 583.383(17) Å3 2.883 g/cm3 2.035 mm−1 0.10 × 0.10 × 0.01 mm3 2.267 to 29.527° −14 ≤ h ≤ 14, −14 ≤ k ≤ 14, −8 ≤ l ≤ 8 28295 636 (Rint = 0.0788) full-matrix least-squares on F2 636/1/52 1.155 R1 = 0.0199, wR2 = 0.0499 R1 = 0.0205, wR2 = 0.0505 0.006(18) 0.245 and −0.282 e/Å3 DOI: 10.1021/acs.cgd.7b01161 Cryst. Growth Des. 2017, 17, 6079−6084

Crystal Growth & Design

Article

Table 3. Atomic Coordinates (×104) and Equivalent Isotropic Displacement Parameters (Å2 × 103) for Na2Ca4(CO3)5a M1

Na1 M1 Ca1 M2 Ca2 X C1 X O1 Y C2 Y O2 Y O3 Z C3 Z O4

Wyckoff

occupancy

x

y

z

Ueq

bond-valence sum

6c 6c 6c 2a 6c 6c 6c 12d 2b 6c

2/3 1/3 1 1 1 1 1 1 1 1

4770(1) 4770(1) 6920(1) 0 8579(3) 6045(3) 5511(3) 7085(2) 3333 4044(2)

5230(1) 5230(1) 8460(1) 0 9290(1) 8022(2) 7756(1) 771(2) 6667 5956(2)

8420(2) 8420(2) 5314(1) 3623(9) 8661(4) 674(5) 8792(4) 6632(3) 5079(9) 5101(5)

14(1) 14(1) 16(1) 11(1) 21(1) 10(1) 16(1) 19(1) 12(1) 30(1)

1.459 2.291 1.848 4.073 4.021

4.069

a

M1, M2, X, Y, and Z superscripts are explained in the text. Bond-valence sums for cations were calculated using R parameters from Brese and O’Keeffe20 (1.80 for Na−O, 1.967 for Ca−O, and 1.39 for C−O) by VESTA software.21

Figure 2. Raman spectrum of Na2Ca4(CO3)5 (decomposition of ν1 carbonate band into three Gaussian components is shown in the inset).

sites: M1 statistically occupied by Na+ and Ca2+ in 2:1 ratio, and M2 occupied solely by Ca 2+. Following classical representation, the structure should be described as a complex ensemble of M1O10 and M2O8 polyhedra linked by multiple carbonate anions. However, a much more clear and meaningful description of the Na2Ca4(CO3)5 structure is possible, using a concept of cationic array with anion-filled voids. The latter approach, developed by A. Vegas and V. A. Blatov for inorganic oxoacid salts,22 was recently effectively used for description and comparative analysis of alkaline earth borates,23−25 alkali carbonates,26 and alkali−alkaline earth carbonates.12 The cationic array of Na2Ca4(CO3)5 is built of corrugated semiregular27 nets perpendicular to the c axis (Figure 3a,b). Each net is built of M13 and M23 triangular loops centered on 3-fold axes and 63 screw axes, respectively; rectangular and hexagonal loops appear between them (Figure 3b). The stacking of the nets is guided by the 63 screw axes, so that M23 triangular loops form columns of face-shared M26 octahedra, whereas M13 triangular loops alternate with hexagonal loops forming trigonal channels (Figures 3c and 4). Three types of voids (X, Y, and Z) filled by (CO3)2− anions exist in the cationic array of Na2Ca4(CO3)5. The first type (X) is the M26 octahedra (Figure 4a), and the second type (Y) is

two-capped trigonal prisms that appear on the joints of neighbor M26 octahedra (Figure 4a). The remaining (CO3)2− anions occupy the trigonal channels (Z), as shown in Figure 4b. The presence of three types of (CO3)2− anions in the Na2Ca4(CO3)5 structure is also confirmed by Raman spectrum, where a ν1 carbonate band near 1080 cm−1 decomposes into three Gaussian components (Figure 2). The structural formula of the compound herewith can be written as M1(Na2Ca) M2Ca3 (CO3)5 or M1(Na2Ca) M2Ca3 X (CO3) Y(CO3)3 Z(CO3), where superscripts indicate the three types of (CO3)2−-filled voids as listed above. 3.2. Natural Analogues of Na2Ca4(CO3)5. The studied high-pressure Na2Ca4(CO3)5 compound has never been reported among known high-pressure minerals. However, it has long been known that many high-pressure structures may be stabilized at moderate pressures by substitution of smaller atoms by larger ones (classic examples are a CsCl-type structure of high-pressure modification of NaCl,28 or germanate analogues of high-pressure silicates29). Taking the latter into account, we also examined Sr- and Ba-containing carbonate minerals, and found that natural analogues of Na2Ca4(CO3)5 compound are minerals of the burbankite group with general formula A3B3(CO3)5, where A = Na, Ca, RE, □ and B = Sr, Ca, 6081

DOI: 10.1021/acs.cgd.7b01161 Cryst. Growth Des. 2017, 17, 6079−6084

Crystal Growth & Design

Article

Na2Ca4(CO3)5 structure. First, pressure required for the material synthesis may be significantly reduced by Ca2+ ← (Sr,Ba)2+ or 2Ca2+ ← (Na+ + RE3+) substitution. Second, a high isomorphic capacity of the structure may be used for tuning of optical properties by wide variations in the material composition and use of various dopants. 3.3. Synthetic Compounds Related to the Na2Ca4(CO3)5. The idea that Sr, Ba, and RE analogues of Na2Ca4(CO3)5 can be synthesized at moderate pressures was confirmed by the recent autoclave synthesis of Na4La2(CO3)55: a promising NLO material with UV cutoff edge of 235 nm and 3×KDP SHG coefficient, which actually is an end member corresponding to complete 2Ca2+ ← (Na+ + RE3+) substitution in the Na2Ca4(CO3)5 structure. The M1 cation sites in the Na4La2(CO3)5 structure are occupied solely by Na+, whereas M2 sites are statistically populated by Na+ and La3+ in 1:2 ratio (see Table 4). The further analysis has shown that the capacity of Na 2 Ca 4 (CO 3 ) 5 structural type is not limited by the burbankite-like compounds. The use of cationic array formalism has clearly demonstrated that Cs3Ba4(CO3)3F5 NLO material3 shares the same structural type, where M1 and M2 cation sites are occupied by Ba and Cs, respectively. However, the M26 cation octahedra (X voids) in this structure are populated by single F− anions instead of (CO3)2− groups, and, even more interestingly, the spacious trigonal channels in the structure (Z voids) accommodate tetrahedral anionic clusters [F4] 4− accompanied by additional Ba2+ cations (Figure 5, compare with Figure 4b). Following the ideas of structural affinity between alkali− alkaline earth carbonates and alkaline earth−rare earth borates,12 we found that the discussed M13M23XY3Z structural type also underlies a number of promising borate materials. The closest borate analogues of the studied Na2Ca4(CO3)5 compound are Ca3La3(BO3)5, a VUV−UV phosphor used for enhancement of solar cells’ photovoltaic performance,30 and Ca3Nd3(BO3)5, a potential self-activated microchip laser medium31 (see Table 4). The M13M23XY3Z structural type can be recognized as well in the structures of recently reported fluoride borates with optical memory Ba7(BO3)4−xF2+3x and Ba4−xSr3+x(BO3)4−yF2+3y,32 and deep UV NLO borosilicate halides Ba7(BO3)3(SiO4)X (X = Cl, Br),33 which share structural features with Cs3Ba4(CO3)3F5 fluoride carbonate discussed above (see Table 4). All given examples, systematized in the Table 4, convincingly demonstrate that the M13M23XY3Z structural type is a promising source of new optical materials in borate, carbonate, and complex systems.

Figure 3. Crystal structure of Na2Ca4(CO3)5: (a) cationic array (neighbor nets are separated with red dashed lines); (b) single cationic net; (c) whole structure. Anion-filled octahedra (X voids) and twocapped trigonal prisms (Y voids) are colored in blue and yellow, respectively. Figures 3−5 were prepared using VESTA software.21

Figure 4. (CO3)2−-filled cation voids in the Na2Ca4(CO3)5 structure: (a) M26 octahedra (X voids, blue) and two-capped trigonal prism (Y void, yellow); (b) trigonal channel (Z void). The legend is the same as in Figure 3.

4. CONCLUSIONS 15

Ba, RE. A and B correspond to the M1 and M2 sites, respectively, in the Na2Ca4(CO3)5 structure. There are four end members in the burbankite group: three RE-free (Na2Ca)Ca3(CO3)5, (Na2Ca)Sr3(CO3)5, (Na2Ca)Ba3(CO3)5, and REcontaining Na 3 (NaRE 2 )(CO 3 ) 5 . Although the studied Na2Ca4(CO3)5 compound actually represents an end member of the burbankite group, natural samples with corresponding composition have not been found. The evident reason is that a significant admixture of Sr or Ba is necessary to stabilize the structure at moderate pressures of the burbankite mineral formation. The mineralogical observations given above provide essential clues to the design of optical materials on the basis of

• A new high-pressure intermediate compound of the Na2CO3−CaCO3 systemNa2Ca4(CO3)5crystallizes in a noncentrosymmetric M13M23XY3Z structural type (P63mc). • Existence of natural analogues of Na 2 Ca 4 (CO 3 ) 5 enriched in Sr, Ba, and RE among minerals of burbankite group demonstrates high isomorphic capacity of the Na2Ca4(CO3)5 structure and suggests a way of its synthesis at moderate pressures, as well as a way to control its optical properties. • Abundance of optical materials with the same M13M23XY3Z structural type among borates suggests a 6082

DOI: 10.1021/acs.cgd.7b01161 Cryst. Growth Des. 2017, 17, 6079−6084

Crystal Growth & Design

Article

Table 4. Carbonates and Borates with M13M23XY3Z-Type Structure compound

structure

M1 site

M2 site

X void (M26 octahedron)

Y void (two-capped trigonal prism)

Z void (trigonal channel)

Carbonates Na2Ca4(CO3)5, this work

burbankite group minerals15 A3B3(CO3)5 Na4La2(CO3)55 Cs3Ba4(CO3)3F53

P63mc a = 10.37402(14) Å; c = 6.25935(9) Å P63mca a = 10.4889(1)−10.5790(1) Å; c = 6.381(3)(9)−6.5446(1) Å P63mc a = 10.55(2) Å; c = 6.47(2) Å P63mc a = 11.5158(9) Å; c = 7.6132(12) Å

Na2Ca

Ca3

(CO3)

(CO3)3

□(CO3)

(Na,Ca,RE,□)3 (A)

(Sr,Ca,Ba,RE)3 (B)

(CO3)

(CO3)3

□(CO3)

Na3

NaLa2

(CO3)

(CO3)3

□(CO3)

Ba3

Cs3

F

(CO3)3

Ba[F4]

P63mc a = 10.530(3) Å; c = 6.398(2) Å P63mc a = 10.4864(2) Å; c = 6.2665(1) Å P63 a = 11.18241(11) Å; c = 7.23720(8) Å P63mc a = 10.87255(18) Å; c = 6.94718(11) Å P63mc a = 11.195(4) Å; c = 7.263(6) Å P63mc a = 11.279(3) Å; c = 7.324(4) Å P63mc a = 11.299(1) Å; c = 7.334(2) Å P63mc a = 10.813(1) Å; c = 6.952(1) Å

Ca3

La3

(BO3)

(BO3)3

□(BO3)

Ca3

Nd3

(BO3)

(BO3)3

□(BO3)

Ba3

Ba3

F

(BO3)3

Ba[Z4]b

Sr3

Ba3

F

(BO3)3

(Ba,Sr)[Z4]b

Ba3

Ba3

Cl

(BO3)3

Ba[SiO4]

Ba3

Ba3

Br

(BO3)3

Ba[SiO4]

Ba3

Ba3

(CN)

(BO3)3

Ba[SiO4]

Sr3

Sr3

(CN)

(BO3)3

Sr[SiO4]

Borates Ca3La3(BO3)534 Ca3Nd3(BO3)531 Ba7(BO3)4−xF2+3x24 Ba4−xSr3+x(BO3)4−yF2+3y25 Ba7(BO3)3(SiO4)Cl33 Ba7(BO3)3(SiO4)Br33 Ba7(BO3)3(SiO4)(CN)35 Sr7(BO3)3(SiO4)(CN)35

Symmetry decreases to the pseudohexagonal P21 in RE-rich samples (not considered here). b[Z4] represents isomorphic substitution [BO3F]4− ↔ [F4]4− in tetrahedral anionic clusters

a



ASSOCIATED CONTENT

Accession Codes

CCDC 1569455 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_ [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Figure 5. Trigonal cationic channels in the Cs3Ba4(CO3)3F5 structure (Z voids) filled by [F4]4− anionic clusters and Ba2+ cations (compare with Figure 4b).

Sergey V. Rashchenko: 0000-0003-2936-0694 Pavel N. Gavryushkin: 0000-0002-9419-2167 Notes

The authors declare no competing financial interest.

promising direction for design of optical materials in carbonate systems. • Generally, the search for optical materials in carbonate systems may be effectively guided by the principles of comparative crystal chemistry; a variety of known borate materials may be used as a basis for structural design of new carbonate ones.



ACKNOWLEDGMENTS

This work was supported by the Russian Science Foundation (Project No. 14-17-00609-Π) and performed under the project of the Ministry of Education and Science of Russian Federation (#No. 14.B25.31.0032). 6083

DOI: 10.1021/acs.cgd.7b01161 Cryst. Growth Des. 2017, 17, 6079−6084

Crystal Growth & Design



Article

(31) Wang, X.; Xia, M.; Li, R. K. New J. Chem. 2016, 40 (3), 2057− 2062. (32) Bekker, T. B.; Solntsev, V. P.; Yelisseyev, A. P.; Rashchenko, S. V. Cryst. Growth Des. 2016, 16 (8), 4493−4499. (33) Lin, X.; Zhang, F.; Pan, S.; Yu, H.; Zhang, F.; Dong, X.; Han, S.; Dong, L.; Bai, C.; Wang, Z. J. Mater. Chem. C 2014, 2 (21), 4257− 4264. (34) Zhou, T.; Ye, N. Acta Crystallogr., Sect. E: Struct. Rep. Online 2008, 64 (6), i37. (35) Schmid, S.; Senker, J.; Schnick, W. J. Solid State Chem. 2003, 174, 221−228.

REFERENCES

(1) Chen, C.; Sasaki, T.; Li, R.; Wu, Y.; Lin, Z.; Mori, Y.; Hu, Z.; Wang, J.; Aka, G.; Yoshimura, M.; Kaneda, Y. Nonlinear Optical Borate Crystals: Principals and Applications; John Wiley & Sons: 2012. (2) Song, Y.; Luo, M.; Zhao, D.; Peng, G.; Lin, C.; Ye, N. J. Mater. Chem. C 2017, 5 (34), 8758−8764. (3) Zou, G.; Ye, N.; Huang, L.; Lin, X. J. Am. Chem. Soc. 2011, 133 (49), 20001−20007. (4) Luo, M.; Song, Y.; Lin, C.; Ye, N.; Cheng, W.; Long, X. Chem. Mater. 2016, 28 (7), 2301−2307. (5) Luo, M.; Wang, G.; Lin, C.; Ye, N.; Zhou, Y.; Cheng, W. Inorg. Chem. 2014, 53 (15), 8098−8104. (6) Luo, M.; Lin, C.; Zou, G.; Ye, N.; Cheng, W. CrystEngComm 2014, 16 (21), 4414−4421. (7) Luo, M.; Ye, N.; Zou, G.; Lin, C.; Cheng, W. Chem. Mater. 2013, 25 (15), 3147−3153. (8) McKie, D.; Frankis, E. J. Z. Fuer Krist. 1977, 145 (1−6), 73−95. (9) Dickens, B.; Hyman, A.; Brown, W. J. Res. Natl. Bur. Stand., Sect. A 1971, 75A (2), 129−140. (10) Gavryushkin, P. N.; Thomas, V. G.; Bolotina, N. B.; Bakakin, V. V.; Golovin, A. V.; Seryotkin, Y. V.; Fursenko, D. A.; Litasov, K. D. Cryst. Growth Des. 2016, 16 (4), 1893−1902. (11) Shatskiy, A.; Gavryushkin, P. N.; Litasov, K. D.; Koroleva, O. N.; Kupriyanov, I. N.; Borzdov, Y. M.; Sharygin, I. S.; Funakoshi, K.; Palyanov, Y. N.; Ijiohtani, E. Eur. J. Mineral. 2015, 27 (2), 175−184. (12) Gavryushkin, P. N.; Bakakin, V. V.; Bolotina, N. B.; Shatskiy, A. F.; Seryotkin, Y. V.; Litasov, K. D. Cryst. Growth Des. 2014, 14 (9), 4610−4616. (13) Bolotina, N. B.; Gavryushkin, P. N.; Korsakov, A. V.; Rashchenko, S. V.; Seryotkin, Y. V.; Golovin, A. V.; Moine, B. N.; Zaitsev, A. N.; Litasov, K. D. Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater. 2017, 73 (2), 276−284. (14) Egorov, K.; Ushchapovskaia, Z.; Kashaev, A.; Bogdanov, G.; Sizykh, I. Dokl. Akad. Nauk SSSR 1988, 301 (1), 188−192. (15) Belovitskaya, Y. V.; Pekov, I. V. New Data Miner. 2004, 39, 50− 64. (16) Shatskiy, A.; Sharygin, I. S.; Gavryushkin, P. N.; Litasov, K. D.; Borzdov, Y. M.; Shcherbakova, A. V.; Higo, Y.; Funakoshi, K. -i.; Palyanov, Y. N.; Ohtani, E. Am. Mineral. 2013, 98 (8−9), 1593−1603. (17) Rothkirch, A.; Gatta, G. D.; Meyer, M.; Merkel, S.; Merlini, M.; Liermann, H.-P. J. Synchrotron Radiat. 2013, 20 (5), 711−720. (18) Sheldrick, G. M. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71 (1), 3−8. (19) Farrugia, L. J. J. Appl. Crystallogr. 2012, 45 (4), 849−854. (20) Brese, N. E.; O’Keeffe, M. Acta Crystallogr., Sect. B: Struct. Sci. 1991, 47, 192−197. (21) Momma, K.; Izumi, F. J. Appl. Crystallogr. 2011, 44 (6), 1272− 1276. (22) Blatov, V. A. In Inorganic 3D Structures; Structure and Bonding; Springer: Berlin, Heidelberg, 2011; pp 31−66. (23) Rashchenko, S. V.; Bekker, T. B.; Bakakin, V. V.; Seryotkin, Y. V.; Simonova, E. A.; Goryainov, S. V. J. Alloys Compd. 2017, 694, 1196−1200. (24) Bekker, T. B.; Rashchenko, S. V.; Bakakin, V. V.; Seryotkin, Y. V.; Fedorov, P. P.; Kokh, A. E.; Stonoga, S. Y. CrystEngComm 2012, 14 (20), 6910−6915. (25) Rashchenko, S. V.; Bekker, T. B.; Bakakin, V. V.; Seryotkin, Y. V.; Shevchenko, V. S.; Kokh, A. E.; Stonoga, S. Y. Cryst. Growth Des. 2012, 12 (6), 2955−2960. (26) Gavryushkin, P. N.; Behtenova, A.; Popov, Z. I.; Bakakin, V. V.; Likhacheva, A. Y.; Litasov, K. D.; Gavryushkin, A. Cryst. Growth Des. 2016, 16 (10), 5612−5617. (27) Wells, A. F. Structural Inorganic Chemistry; Clarendon Press: 1984. (28) Sato-Sorensen, Y. J. Geophys. Res. 1983, 88 (B4), 3543−3548. (29) Dachille, F.; Roy, R. Nature 1959, 183 (4670), 1257−1257. (30) Dai, W. B.; Lei, Y. F.; Li, P.; Xu, L. F. J. Mater. Chem. A 2015, 3 (9), 4875−4883. 6084

DOI: 10.1021/acs.cgd.7b01161 Cryst. Growth Des. 2017, 17, 6079−6084