Chemical Cosubstitution-Oriented Design of Rare-Earth Borates as

Nov 21, 2017 - Among borates, alkali and alkaline earth metals are the preferred alternative elements for designing borate-based NLO materials, mainly...
0 downloads 8 Views 5MB Size
Article Cite This: J. Am. Chem. Soc. 2017, 139, 18397−18405

pubs.acs.org/JACS

Chemical Cosubstitution-Oriented Design of Rare-Earth Borates as Potential Ultraviolet Nonlinear Optical Materials Miriding Mutailipu,†,‡,# Zhiqing Xie,†,# Xin Su,§ Min Zhang,*,† Ying Wang,*,† Zhihua Yang,† Muhammad Ramzan Saeed Ashraf Janjua,∥ and Shilie Pan*,† †

Key Laboratory of Functional Materials and Devices for Special Environments, Xinjiang Technical Institute of Physics & Chemistry, Chinese Academy of Sciences; Xinjiang Key Laboratory of Electronic Information Materials and Devices, 40-1 South Beijing Road, Urumqi 830011, China ‡ University of the Chinese Academy of Sciences, Beijing 100049, China § Xinjiang Laboratory of Phase Transitions and Microstructures in Condensed Matter Physics, College of Physical Science and Technology, Yili Normal University, Yining, Xinjiang 835000, China ∥ Chemistry Department, King Fahd University of Petroleum and Minerals (KFUPM), Dhahran 31261, Kingdom of Saudi Arabia S Supporting Information *

ABSTRACT: A chemical cosubstitution strategy was implemented to design potential ultraviolet (UV) and deep-UV nonlinear optical (NLO) materials. Taking the classic βBaB2O4 as a maternal structure, by simultaneously replacing the Ba2+ and [B3O6]3− units with monovalant (K+), divalent (alkaline earth metal), trivalent (rare-earth metal, Bi3+) ions, and the [B5O10]5− clusters through two different practical routes, 12 new mixed-metal noncentrosymmetric borates K7MIIRE2(B5O10)3 (MII = Ca, Sr, Ba, K/RE0.5; RE = Y, Lu, Gd) as well as K7MIIBi2(B5O10)3 (MII = Pb, Sr) were successfully designed and synthesized as high-quality single crystals. The selected K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3 compounds were subjected to experimental and theoretical characterizations. They all exhibit suitable second-harmonic generation (SHG) responses, as large as that of commercial KH2PO4 (KDP), and also exhibit short UV cutoff edges. These results confirm the feasibility of this chemical cosubstitution strategy to design NLO materials and that the three selected crystals may have potential application as UV NLO materials. UV spectral region.3,29 As a result, a variety of mixed alkali and/ or alkaline earth metal borates, BaAlBO3F2,30 Na2CsBe6B5O15,31 Li4Sr(BO3)2,32 K3B6O10X (X = Br, Cl),33,34 etc., have been frequently reported as front-line candidates for UV or deep-UV NLO materials. From the electronic configuration point of view, several trivalent rare-earth cations RE3+ (RE = Sc, Y, La, Gd, and Lu) with closed-shell electronic configurations or half-occupied 4f orbitals are also suitable, which will effectively inhibit the unfavorable d−d or f−f electronic transitions and broaden the transparency window.35−46 Furthermore, rare-earth cations coordinated into RE-based deformed polyhedra with relatively large hyperpolarizability can enhance second-harmonic generation (SHG) responses. For example, distorted YO 6 octahedra, as the dominant NLO active microscopic units in YAl3(BO3)4, improved second-order optical susceptibility.35 In view of the aforementioned advantages, rare-earth borates RECa4O(BO3)3 (RE = Y, Gd),36−38 Na3La9O3(BO3)8,39,40 and

1. INTRODUCTION The last several decades have witnessed the astounding progress in the exploration and design of ultraviolet (UV; λ < 400 nm) and deep-UV (λ < 200 nm) nonlinear optical (NLO) materials, as they can satisfy the constantly growing demands for multiple technological applications.1−6 From the sustained efforts of many chemists and materials scientists to discover NLO materials with remarkable properties, several chemical systems, such as borates,7−9 phosphates,10−12 carbonates,13−15 nitrates,16 etc.,17−20 have been considered as alternative systems in the search for UV and deep-UV NLO materials. With respect to borates, they reveal a wide optical transparency window, varied acentric structure types, high laser damage thresholds, and large polarizabilities.21−25 As such, several commercial borate-based NLO materials have been developed, including β-BaB2O4 (β-BBO),26 LiB3O5 (LBO),27 KBe2BO3F2 (KBBF),28 etc. Among borates, alkali and alkaline earth metals are the preferred alternative elements for designing borate-based NLO materials, mainly because those selected cations are free of d−d or f−f electronic transitions, which is beneficial for good transparency in the UV and even the deep© 2017 American Chemical Society

Received: October 23, 2017 Published: November 21, 2017 18397

DOI: 10.1021/jacs.7b11263 J. Am. Chem. Soc. 2017, 139, 18397−18405

Article

Journal of the American Chemical Society Table 1. Crystallographic Data for This Series of Compounds K7CaY2B15O30

K7SrY2B15O30

K7BaY2B15O30

K7CaLu2B15O30

K7SrLu2B15O30

K7BaLu2B15O30

formula weight crystal system space group a (Å) c (Å) Z volume (Å3) F(000) completeness (%) GOF on F2 absolute structure parameter final R indices [Fo2>2σ(Fo2)]a R indices (all data)a

1133.75 trigonal R32 13.2182(18) 14.949(4) 3 2262.0(8) 1638 100 0.971 −0.020(7)

1181.29 trigonal R32 13.1142(18) 15.319(4) 3 2281.6(8) 1692 99.8 0.961 −0.001(10)

1231.01 trigonal R32 12.9884(15) 15.718(4) 3 2296.4(6) 1746 99.7 1.099 0.000

1305.87 trigonal R32 13.1433(10) 14.905(2) 3 2229.9(4) 1830 100 1.096 −0.012(17)

1353.41 trigonal R32 13.0208(14) 15.297(3) 3 2246.1(6) 1884 99.5 1.095 0.017(15)

1387.02 trigonal R32 12.9705(15) 15.651(4) 3 2280.3(6) 1914 99.7 1.112 0.000

R1 = 0.0236, wR2 = 0.0477 R1 = 0.0271, wR2 = 0.0488 K7CaGd2B15O30

R1 = 0.0265, wR2 = 0.0575 R1 = 0.0323, wR2 = 0.0596 K7SrGd2B15O30

R1 = 0.0508, wR2 = 0.1367 R1 = 0.0549, wR2 = 0.1400 K7.5Y2.5B15O30

R1 = 0.0213, wR2 = 0.0569 R1 = 0.0221, wR2 = 0.0574 K7.5Gd2.5B15O30

R1 = 0.0205, wR2 = 0.0587 R1 = 0.0211, wR2 = 0.0590 K7PbBi2B15O30

R1 = 0.0351, wR2 = 0.0938 R1 = 0.0365, wR2 = 0.0947 K7SrBi2B15O30

formula weight crystal system space group a (Å) c (Å) Z volume (Å3) F(000) completeness (%) GOF on F2 absolute structure parameter final R indices [Fo2>2σ(Fo2)]a R indices (all data)a

1270.43 trigonal R32 13.310(4) 15.005(8) 3 2302.1(15) 1788 100 1.052 −0.01(2)

1317.97 trigonal R32 13.148(2) 15.417(5) 3 2308.2(9) 1842 100 1.055 0.048(15)

1160.35 trigonal R32 13.142(2) 15.099(5) 3 2258.3(9) 1119 100 1.017 −0.022(14)

1328.71 trigonal R32 13.264(6) 15.282(14) 3 2329(3) 1194 99.9 1.021 −0.004(17)

1541.00 trigonal R32 13.331(2) 14.910(5) 3 2294.6(10) 2088 99.6 1.134 0.000

1421.43 trigonal R32 13.2512(19) 15.288(4) 3 2324.8(8) 1956 100 1.068 0.001(9)

R1 = 0.0237, wR2 = 0.0607 R1 = 0.0241, wR2 = 0.0610

R1 = 0.0192, wR2 = 0.0391 R1 = 0.0206, wR2 = 0.0396

R1 = 0.0392, wR2 = 0.0855 R1 = 0.0547, wR2 = 0.0913

R1 = 0.0209, wR2 = 0.0433 R1 = 0.0221, wR2 = 0.0441

R1 = 0.0426, wR2 = 0.1182 R1 = 0.0435, wR2 = 0.1194

R1 = 0.0195, wR2 = 0.0437 R1 = 0.0217, wR2 = 0.0488

a

R1 = ∑||Fo| − |Fc||/∑|Fo| and wR2 = [∑w(Fo2 − Fc2)2/∑wFo4]1/2 for Fo2 > 2σ(Fo2).

On the basis of the ideas above, using classic β-BBO as a maternal structure, we cosubstituted the cationic parts of βBBO with multiple rare-, alkali-, and alkaline-earth metals to successfully synthesize 12 new noncentrosymmetric borates. For the anionic units, the isolated [B5O10]5− double rings in these derivatives can be considered as the recombination of two [B3O6]3− single rings in β-BaB2O4 through a shared B atom. In this paper, we select K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3 as representatives for experimental and theoretical characterizations. Results show that all three selected compounds exhibit suitable SHG responses and short UV cutoff edges, indicating that they may have potential applications as NLO materials ranging from the UV to the nearIR spectral region.

La2CaB10O1941 show short UV cutoff edges and suitable SHG coefficients. In addition to select the appropriate cations, the effective design of NLO materials with presupposed structures or properties is also critically important.47,48 Recently, the cosubstitution strategy has been proved to be an effective way to design new functional materials based on known classical compounds, which can reorganize crystal structures by simultaneously replacing two (or more) ions, fundamental building units, etc.49−52 On the basis of the chemical cosubstitution strategy, Na3Ba2(B3O6)2F (NBBF) can be structurally considered as a derivative from cosubstituting the Ba2+ atoms in α-BBO with 3Na+ and F− atoms in NBBF.50,53 Similarly, BaAlBO3F230 was derived from simultaneously replacing K+ and 2Be2+ groups in KBBF with Ba2+ and Al3+, which results in the transformation of beryllium borate-based NLO materials to beryllium-free ones. It was also discovered that [BO3]3− groups can occupy the sites of [PO4]3− units in apatite Ca5(PO4)3F, giving Ca5(BO3)3F,54 which can be considered as the result of cosubstitution behavior among different anionic building blocks and demonstrates the mutual transformation between borates and phosphates. Similar manners of cosubstitution can also occur between [PO4]3− and [VO4]3−, [BO3F]4− and [BeO3F]5−, [NbO6]7− and [TiO5F]7−, [3F−] and [BO3]3−, etc.55−58

2. EXPERIMENTAL SECTION 2.1. Reagents. All raw materials including KF, K2CO3, MIIF2 (MII = Ca, Sr, Ba), MIICO3 (MII = Ca, Sr, Ba, Pb), RE2O3 (RE = Y, Lu, Gd), Bi2O3, and B2O3 were purchased from Shanghai Aladdin BioChem Technology Co., Ltd., as K, MII, RE, Bi, and B-based reactants for the expected K7MIIRE2(B5O10)3 (MII = Ca, Sr, Ba, K/RE0.5; RE = Y, Lu, Gd) and K7MIIBi2(B5O10)3 (MII = Pb, Sr) phases, respectively. 2.2. Single-Crystal Preparation and the Synthesis of Title Compounds. Single crystals of K7MIIRE2(B5O10)3 (MII = Ca, Sr, Ba; RE = Y, Lu, Gd) and K7MIIBi2(B5O10)3 (MII = Pb, Sr) were obtained by high-temperature solution reaction by mixing KF, MIIF2, RE2O3(or Bi2O3), and B2O3 according to a molar ratio of 14−17:2−5:2−5:15− 17. For the targeted K7.5RE2.5(B5O10)3 (RE = Y, Gd), the molar ratios 18398

DOI: 10.1021/jacs.7b11263 J. Am. Chem. Soc. 2017, 139, 18397−18405

Article

Journal of the American Chemical Society

Figure 1. Powder X-ray diffraction patterns (a) and TG-DSC curves (b) of K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3. with the SAINT program.59 The structures were established by the direct method with program SHELXS and refined by the full-matrix least-squares program SHELXL.60 The final solved structures were rechecked by the program PLATON for verifying the probable missing symmetric elements,61 but no higher symmetries were recommended. During the refinement section of this series structures, we found that the K(1) and Y(2) atoms in K7.5Y2.5(B5O10)3, and K(1) and Gd(2) atoms in K7.5Gd2.5(B5O10)3, were in substitution disorder. Therefore, judging from the atomic thermal displacement parameters and charge balance principles, the K(1) and Y(2), K(1) and Gd(2) atoms in K7.5Y2.5(B5O10)3 and K7.5Gd2.5(B5O10)3 were set to share the same sites with a disorder ratio of 1:1, which can help us to get relatively better residual (R) values and reasonable equivalent isotropic displacement parameters. The crystallographic data are listed in Table 1. The related crystal data, including selected bond lengths and atomic coordinate equivalent isotropic displacement parameters, are listed in Tables S2− S25 in Supporting Information, respectively. 2.6. Vibrational Spectroscopy and Optical Properties. Infrared spectroscopy was carried out on a Shimadzu IRAffinity-1 spectrometer to specify and compare the coordination of the B atoms in K7MIIY2(B5O10)3 (MII = Ca, Sr, Ba). The samples were mixed with dried KBr, and then examined in a range from 400 to 4000 cm−1 with a resolution of 2 cm−1. The UV−vis−NIR diffuse-reflectance data for the polycrystalline powders of K7MIIY2(B5O10)3 (MII = Ca, Sr, Ba) were collected with a Shimadzu Solid Spec-3700DUV spectrophotometer with the measurement range extending from 190 to 2500 nm. 2.7. Powder Second-Harmonic Generation Measurements. The powder SHG measurements for targeted K7MIIY2(B5O10)3 (MII = Ca, Sr, Ba) were carried out on the basis of the Kurtz−Perry method at room temperature.62 A Q-switched Nd:YVO4 solid-state laser was selected as the source of radiation at λω = 1064 nm, and the radiation was doubled to second harmonics λ2ω = 532 nm. The crystal aggregates were ground and sieved into a series of distinct size ranges: 20−38, 38−55, 55−88, 88−105, 105−150, 150−200, and 200−250 μm, which were placed into a light-tight box and irradiated with the laser. The intensities of the frequency-doubled output emitted from the samples were collected by a photomultiplier tube. The commercial NLO crystal KDP was also ground and sieved into the same particle size range for the reference. 2.8. Calculation Details. The electronic structure calculations were performed using a plane-wave basis set, and pseudopotentials within density functional theory (DFT) were also implemented in the total-energy module CASTEP.63 The exchange and correlation effects were treated by the Perdew−Burke−Ernzerhof (PBE) method in generalized gradient approximation (GGA).64 The interactions between the ionic cores and electrons were described by normconserving pseudopotentials.65 The following orbital electrons were treated as valence electrons: K 3s2 3p6 4s1, Ca 3s2 3p6 4s2, Sr 4s2 4p6 5s2, Ba 5s2 5p6 6s2, Y 4d1 5s2, B 2s2 2p1, and O 2s2 2p4. The number of plane waves included in the basis was determined by a cutoff energy of

of the loaded reactants are KF: RE2O3:B2O3 = 15−17:3−5:15−17. Those raw reagents were mixed homogeneously and placed into platinum crucibles. Then the samples were heated to 850 °C at a rate of 50 °C/h and held at this selected temperature for 24 h to melt thoroughly, after that, cooled to 750 °C at a rate of 1 °C/h and then to 30 °C at a rate of 20 °C/h. Colorless, transparent single crystals formed that could be manually separated from reaction products to further determine microscopic single-crystal structures. Polycrystalline samples of K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3 were synthesized via solid-state reaction by mixing K2CO3, MIICO3 (MII = Ca, Sr, Ba), Y2O3, and B2O3 at a ratio of 3.5:0.9:0.8:7.5. The mixtures were preheated at 400 °C for 48 h. After that, the temperature was gradually raised to 820 °C, 780 °C, 760 °C for the Ca-, Sr-, Ba-based compounds, respectively. Several intermediate grindings and mixings were performed, and then the mixtures were held for 72 h. With this procedure, pure polycrystalline samples of the three targeted compounds were successfully obtained. After many experiments to search for suitable fluxes (Table S1 in Supporting Information), a single crystal of K7BaY2(B5O10)3 was finally grown by the top-seeded solution growth method with KF as the flux. Pure polycrystalline samples of K7BaY2(B5O10)3 as well as KF were mixed at a molar ratio of 1:3.2 and heated to 870 °C over the course of 24 h. A seed obtained from an earlier stage of spontaneous crystallization was introduced into the solution at a rotation rate of 5 rpm. Then the temperature was slowly decreased to 789 °C at a rate of 1 °C/h and then cooled to 786 °C at a rate of 1 °C/day. When the crystal growth process was completed, the as-grown crystal was pulled from the solution and cooled (20 °C/h) to room temperature. With this procedure, millimeter-scale single crystals of K7BaY2(B5O10)3 were obtained (Figure S1 in Supporting Information). 2.3. Thermal Behavior Analysis. Thermal gravimetric analysis (TGA) and differential scanning calorimetry (DSC) of K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3 were investigated using a NETZSCH STA 449C simultaneous thermal analyzer. The samples were placed in a platinum crucible and heated at a rate of 5 °C/min from 40 °C to 1000 °C and then cooled to 200 °C at a rate of 5 °C/min under a flow of N2. 2.4. Powder X-ray Diffraction. Powder X-ray diffraction (PXRD) measurements for the targeted K7MIIY2(B5O10)3 (MII = Ca, Sr, Ba) were carried out on a Bruker D2 PHASER diffractometer equipped with Cu Kα radiation at room temperature. The 2θ range is 10−70° with a step size of 0.02° and a fixed counting time of 1 s/step. 2.5. Structure Determination. Single crystals of K7MIIRE2(B5O10)3 (MII = Ca, Sr, Ba, K/RE0.5; RE = Y, Lu, Gd) and K7MIIBi2(B5O10)3 (MII = Pb, Sr) were mounted on a glass fiber for data collection by single-crystal XRD on an APEX II CCD diffractometer equipped with monochromatic Mo Kα radiation at room temperature. Reductions of the collected data were performed with the Bruker Suite software package. The numerical absorption corrections were carried out with the SADABS program and integrated 18399

DOI: 10.1021/jacs.7b11263 J. Am. Chem. Soc. 2017, 139, 18397−18405

Article

Journal of the American Chemical Society 540 eV, and the numerical integration of the Brillouin zone was performed using a 4 × 4 × 4 Monkhorst−Pack scheme k-point grid sampling for K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3.

3. RESULTS AND DISCUSSION 3.1. Synthesis and Thermal Behavior. During the synthesis process for the pure polycrystalline samples, we tried to obtain pure phases of K7MIIY2(B5O10)3 (MII = Ca, Sr, Ba) at the stoichiometric ratio via traditional solid-state reaction methods. Unfortunately, before melting at about 1000 °C, the observed XRD patterns confirmed the coexistence of targeted K7MIIY2(B5O10)3 phases and YBO3 (PDF no. 32-1428), which would lead to some inaccurate results during the physicochemical properties tests. Faced with this, several attempts were made to improve the purity, including changing the raw materials, adjusting the molar ratios, using other synthetic methods (molten salt and sol−gel method), changing the temperature program, etc. Finally, pure polycrystalline samples of K7MIIY2(B5O10)3 were successfully obtained by mixing the raw materials via solid-state reaction methods at a molar ratio of K2CO3:MIICO3 (MII = Ca, Sr, Ba):Y2O3:B2O3 = 3.5:0.9:0.8:7.5. As shown in Figure 1a, the observed XRD patterns are in good agreement with the corresponding theoretical ones. Thermal stabilities of selected K 7 CaY 2 (B 5 O 1 0 ) 3 , K7SrY2(B5O10)3, and K7BaY2(B5O10)3 polycrystalline samples were evaluated by TG and DSC. For each of the three compounds, only one endothermic peak is observed on the heating part of the DSC curves and no obvious weight loss is observed on their TG curves in the range from 40 °C to 1000 °C, which tentatively proves that the tested compounds can be stable at least up to 965 °C, 946 °C, and 922 °C, respectively (Figure 1b). There are no peaks on the cooling part of DSC curves, which is caused by high viscosity of the B-rich systems and can be confirmed by the glassy products of the tested samples. Till now, whether or not the K7MIIY2(B5O10)3 series belongs to congruent melting compounds is still unknown based only on the results of TG and DSC. Therefore, the polycrystalline samples (2.0 g) of K7MIIY2(B5O10)3 were placed into a platinum crucible and heated to 1000 °C for homogeneous melting and then cooled to room temperature very slowly (1.0 °C/h). The PXRD patterns of the residues show diffraction patterns different from those of the initial powders (Figures S2−S4 in Supporting Information), which indicates that those compounds melt incongruently and that a suitable flux should be introduced to decrease the temperature during crystal growth. Obvious inclusions were observed in the as-grown K7BaY2(B5O10)3 crystal, indicating that more suitable growth process parameters (flux, orientation of the seed, temperature fields, etc.) should be further explored. Therefore, the crystal growth of related compounds will also be examined with the aim to use these crystals as UV nonlinear optical materials. 3.2. Crystal Structure Description. Crystallographic analysis reveals that all 12 reported borates are isostructural and crystallize in the same noncentrosymmetric trigonal space group, R32, with only slight differences in the related unit cell parameters. The volume increases continuously from Ca-, Sr-, to Ba-based compounds with the decrease of a values and increase of c values in two series K7MIIRE2(B5O10)3 (MII = Ca, Sr, and Ba; RE = Y and Lu), respectively. The ball-and-stick topological view of the frameworks are shown in Figure 2a, which represents a three-dimensional (3D) framework composed of KO6, MIIO6 (MII = Ca, Sr, Ba), REO6 (RE = Y,

Figure 2. Crystal structure of K7MIIRE2(B5O10)3. (a) The whole crystal structure of K7MIIRE2(B5O10)3 viewed along the c axis. (b) Topological representation of K7MIIRE2(B5O10)3: the [B5O10]5− and REO6 basic units are regarded as 4-c and 6-c nodes. (c) [B5O10]5− communities. (d) Layered structures of K7MIIRE2(B5O10)3. (e) RE− B−O-based single layer in K7MIIRE2(B5O10)3.

Lu, Gd, K/RE), or BiO6 octahedra and isolated [B5O10]5− clusters. The asymmetric unit of K7MIIRE2(B5O10)3 contains three, one, one, three, and five crystallographic independent K, MII, RE, B, and O atoms, respectively. With respect to the B−O structural framing, the B(1, 2) and B(3) atoms are three and four coordinated with O atoms into the triangular [B(1,2)O3]3− and tetrahedral [B(3)O4]5− units, respectively. The centered [B(3)O4]5− group is surround by two pairs of [B(1,2)O3]3− groups to form a [B5O10]5− cluster, (Figure 2c) which are arranged in isolated configuration without connecting to any other B or B-containing units. For the RE atoms, they are sixcoordinated for the REO6 octahedra, which continuously share four of six equatorial oxygens with the nearest [B5O10]5− units to form a RE−B−O-based single layer (Figure 2d and 2e), and then those single layers are further connected to form double layers. Further, the whole 3D structures are stacked through the connection of the RE−O bonds and the [B5O10]5− clusters in an approximately vertical manner between double layers. The K and MII atoms are identically six-coordinated, forming the KO6 and MIIO6 octahedra and filling the interstices between those RE−B−O-based layers. Furthermore, the topology of the K7MIIRE2(B5O10)3 series can be simplified by considering the [B5O10]5− clusters and the REO6 octahedra as four-connected (4-c) and six-connected (6-c) nodes (Figure 2b), respectively. As a result, a hea-type topological 4,6-c 2-nodal net is formed with a Schläfli symbol of {43;63}3{46;69}2 for this series of structures. From a chemical point of view, the K7MIIRE2(B5O10)3 and K7MIIBi2(B5O10)3 series can be regarded as derivatives of the cosubstitution strategy based on β-BaB2O4 (Figures 3 and 4). Expanding the stoichiometry of β-BaB2O4 to 15 times that of original value results in Ba7Ba2Ba6(B3O6)5. Based on this, two practical routes from maternal structures to derivatives are given: 15Ba2+ = 14K++ 2M2+ + 4RE3+(or 4 Bi3+) and 15Ba2+ = 14K++ (K/RE)4+ + 4RE3+. (1) [Ba7]14+, [Ba2]4+, and [Ba6]12+ are respectively replaced by the K14, MII2 , and MIII 4 units with the 18400

DOI: 10.1021/jacs.7b11263 J. Am. Chem. Soc. 2017, 139, 18397−18405

Article

Journal of the American Chemical Society

Figure 3. Structural similarity between β-BaB2O4 (left) and K7MIIRE2(B5O10)3 (right).

isolated [B3O6]3− units can also be regarded as the polymerization of three [BO3]3− units in Ba2Mg(BO3)266 through three shared O atoms in Ba2Mg(B3O6)2,67 which can also occur between Ba3Y(BO3)368 and Ba3Y(B3O6)3.69 Furthermore the evolution among borates, the structural evolution of anionic groups can also be found between different compound systems, such as borates and phosphates (BO3 → PO4),54,70 vanadates and phosphates (VO4 → PO4),55 carbonates and borates (CO3 → BO3),71 silicates and germanates (SiO4 → GeO4),72 etc. With this, chemical cosubstitution strategy is confirmed to be a feasible strategy to design multiple compounds with diverse compositions based on the same template or maternal structures. However, the relative stability of the cation−anion interactions is also indispensable for the successful synthesis of these series of derivatives. All the K, MII, and Y atoms in the three structures are the central atoms of six-coordinated polyhedra, which means that the substitutable positions (Ca, Sr, and Ba) are in a similar atomic environment and can improve the feasibility of cosubstitution behaviors. The Lewis acidity values73 of K+, M2+, and Y3+ are defined as the formal valence divided by the coordination number (CN = 6). Those values (Table 2) are close to each other in three related compounds, which can also ensure the feasibility of mutual substitution. Furthermore, the global instability index (GII) and bond strain index (BSI) values were also calculated.74−76 Generally, a BSI or GII value larger than 0.05 vu (valence unit) indicates that a structure is strained while values larger than 0.20 vu imply that a structure is unstable.74,75 As listed in Table 2, the BSI values are larger than 0.05 vu and the GII values are smaller than 0.20 vu for K7CaY2(B5O10)3 (0.06, 0.13), K7SrY2(B5O10)3 (0.05, 0.12), and K7BaY2(B5O10)3 (0.05, 0.12), indicating that they are all strained and stable. Those values are very close to each other, which is reasonable as they possess nearly the same BVS values and atomic coordination environment (CN = 6).

Figure 4. Evolution based on chemical cosubstitution strategy from βBaB2O4 to a series of derivatives.

same chemical valence states, resulting in K7MIIMIII 2 B15O30, where MII is divalent Ca2+, Sr2+, and Ba2+, and MIII is the trivalent rare-earth Y3+, Lu3+, and Gd3+ for the derivatives K7MIIRE2(B5O10)3, while MII and MIII are divalent Pb2+, Sr2+, and trivalent Bi3+ for the derivatives K7PbBi2(B5O10)3 and K7SrBi2(B5O10)3, respectively. (2) [Ba7]14+, [Ba2]4+, and [Ba6]12+ are respectively replaced by the K14, KMIII, and MIII 4 units with the same chemical valence states, resulting in III 3+ K7.5MIII and 2.5 B15O30, where M is the trivalent rare-earth Y Gd3+ for the derivatives K7.5Y2.5(B5O10)3 and K7.5Gd2.5(B5O10)3, respectively. For the B−O anionic parts, the isolated [B5O10]5− double rings in two series of derivatives can be considered as the recombination of two [B3O6]3− single rings in β-BaB2O4 through a shared B atom. The similar structural evolution between B−O units can also be found in Ba2Mg(BO3)2 and Ba2Mg(B3O6)266,67 (Figure S5 in Supporting Information). The

Table 2. Bond Valence Sum, Lewis Acidity Value, Bond Strain Index (BSI), and Global Instability Index (GII) for K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3 BVS

Lewis acidity value (vu)

compound

K

MII

Y

K

MII

Y

BSI (vu)

GII (vu)

K7CaY2(B5O10)3 K7SrY2(B5O10)3 K7BaY2(B5O10)3

0.805−1.261 0.815−1.238 0.944−1.293

2.091 2.178 2.060

3.207 3.114 3.087

0.13−0.21 0.14−0.21 0.16−0.22

0.35 0.36 0.34

0.53 0.52 0.51

0.06 0.05 0.05

0.13 0.13 0.12

18401

DOI: 10.1021/jacs.7b11263 J. Am. Chem. Soc. 2017, 139, 18397−18405

Article

Journal of the American Chemical Society

Figure 5. Infrared spectra (a) and UV−vis−NIR spectra (b) of K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3.

3.3. Vibrational Spectroscopy and Optical Properties. To specify and compare the coordination model of the B atoms in K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3, the infrared spectra were examined. As shown in Figure 5a, the strong bands at 1191, 1256, and 1386 cm − 1 for K 7 CaY 2 (B 5 O 1 0 ) 3 , 1212, 1255, and 1375 cm − 1 for K7SrY2(B5 O10) 3, and 1180, 1256, and 1378 cm −1 for K7BaY2(B5O10)3, are mainly due to the asymmetric stretching of [BO3]3− units. The strong bands at around 941 and 1027 cm−1, 941 and 1028 cm−1, and 930 and 1039 cm−1 in the curves are attributed to the [BO4]5− asymmetric stretching vibrations for the Ca-, Sr-, Ba-based compounds, respectively. The absorption bands around 734 and 776 cm − 1 for K7CaY2(B5O10)3, 732 and 778 cm−1 for K7SrY2(B5O10)3, and 745 and 776 cm−1 for K7BaY2(B5O10)3 can be assigned to both the [BO3]3− and [BO4]5− bending modes. Based on this, the existence of the [BO3]3− triangles and the [BO4]5− tetrahedra is further confirmed, which is consistent with the results derived from the single-crystal X-ray structural analyses and other reported borates.77,78 Furthermore, the UV−vis−NIR diffuse reflectance spectra for three compounds are shown in Figure 5b. Results show that their cutoff edges are all lower than 190 nm, indicating that all three crystals may have potential NLO applications in the UV spectral region. 3.4. Second-Harmonic Generation Properties. Plots of the SHG intensity versus particle sizes for ground polycrystalline samples of K7MIIY2(B5O10)3 (MII = Ca, Sr, Ba) are shown in Figure 6. Detailed features of those curves indicate that all three crystals are type I phase-matchable based on the rule proposed by Kurtz and Perry.62 The comparison of the SHG signal produced by the K7MIIY2(B5O10)3 polycrystalline samples and the KDP samples in the same particle sizes ranging from 200 to 250 μm reveals that Ca-, Sr-, Ba-based compounds exhibit a suitable SHG response of ∼0.9, 1.1, and 1.2 × KDP, respectively. Such SHG responses are large enough for UV NLO applications. To further investigate the NLO properties of K7MIIY2(B5O10)3 (MII = Ca, Sr, Ba), we also calculated the second-order coefficients dij based on the first-principles calculation. The space group of K7MIIY2(B5O10)3 (MII = Ca, Sr, Ba) is R32, which belongs to the class 32 point group and has only two nonvanishing independent SHG tensor component (d11 and d14) under the restriction of Kleinman symmetry. Among the two tensors, d14 is very small and close to zero, therefor, only d11 sensor of three compounds were calculated and the calculated d11 is 0.65, 0.67, and 0.69 pm/V

Figure 6. Powder SHG data for K7MIIY2(B5O10)3 (MII = Ca, Sr, Ba) at 1064 nm. The curves are drawn to guide the eye and are not a fit to the data.

for Ca-, Sr-, Ba-based compounds, respectively. The values are comparable to that of KDP (d36 = 0.39 pm/V) and also in good agreement with the experimental values and tendency (Ca < Sr < Ba), which verifies the validity of the pseudopotential methods we employed. 3.5. Electronic Structure Calculations. The characteristic features of calculated band structures for K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3 are plotted in Figures S6−S8 in Supporting Information, which show a pair of bonding and antibonding orbitals separated by gaps, respectively. It is shown that all three compounds are direct band gap crystals with the valence band (VB) maximum and the conduction band (CB) minimum located at the same Γ point. The values of band gaps were calculated to be 4.53, 4.51, and 4.47 eV for Ca-, Sr-, Ba-based crystals, respectively, and the values are smaller than experimental values resulting from discontinuity of the exchange-correlation energy functional. These differences will be corrected using a so-called scissors energy shift when evaluated optical properties based on DFT band structure calculation result. Therefore, the scissors are adopted in the following optical properties calculations.79 The partial density of state (PDOS) projected on the constitutional atoms of K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3 crystals are given in Figure 7. The VBs below the Fermi level are mainly composed of three parts in the energy range from 0 to −20 eV. (1) The higher energy subband 18402

DOI: 10.1021/jacs.7b11263 J. Am. Chem. Soc. 2017, 139, 18397−18405

Article

Journal of the American Chemical Society

Figure 7. Partial density of states (PDOS) of K7CaY2(B5O10)3 (a), K7SrY2(B5O10)3 (b), and K7BaY2(B5O10)3 (c).

Figure 8. SHG-density of unoccupied and occupied states of K7CaY2(B5O10)3 (a, d), K7SrY2(B5O10)3 (b, e), and K7BaY2(B5O10)3 (c, f) in the virtualelectron (VE) process.

lying between −8.5 eV and the Fermi level is dominated by O 2p orbitals with partly contributions of B 2p and Y 4d oritals. The upper two peaks of the VB between 0 and −2 eV mainly consist of O 2p orbitals with contributions of Y 4d orbitals. Likewise, the band between −2 and −8.5 eV is produced by the hybridization of B 2p, O 2p, and Y 5s orbitals. (2) Furthermore,

one narrow band with a sharp peak is at about −11 eV due to almost entirely to a contribution from K 3p orbitals. (3) The lower parts of the VB between −16 and −20 eV are made up entirely of the B and O orbitals, especially O 2p and B 2p orbitals between −20 and −16 eV, due to strong hybridization of the B and O orbitals in the [B5O10]5− groups. The CB above 18403

DOI: 10.1021/jacs.7b11263 J. Am. Chem. Soc. 2017, 139, 18397−18405

Article

Journal of the American Chemical Society

X-ray crystallographic file in CIF format; checkcif files (ZIP)

the Fermi level is dominated by K 3p, B 2p, O 2p and Ca, Sr, Ba ns values (n = 4, 5, 6), while the bottom of CB mostly consists of K 3p and Y 4d atoms, respectively. 3.6. Origin of the SHG Effects. To further explain the donation of individual atoms in K 7 CaY 2 (B 5 O 1 0 ) 3 , K7SrY2(B5O10)3, and K7BaY2(B5O10)3 to the SHG effect in real space, we used the SHG-density method.80 The SHG process is denoted by two virtual transition processes, namely virtual electron (VE) and virtual hole (VH) processes. The VE process contributes close to 83.2%, 84.4%, and 89.5% to the largest SHG tensors (d11) of Ca-, Sr-, Ba-based crystals, respectively, indicating that SHG effects of the three compounds mainly come from the VE processes. Figure 8 gives a clear description that SHG-densities on the O atoms are dominant contributors in the three compounds. The [B5O10]5− group as the basic anionic unit is composed of four [B(1,2)O3]3− and one centered [B(3)O4]5− units. From the density of the SHG effect in the [B5O10]5− groups of K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3, it shows that three-coordinated [B(1)O3]3− and [B(2)O3]3− as well as the YO6 octahedra are the major sources of the SHG effect, while the contributions from four-coordinated [B(3)O4]5− units and the K atoms are negligibly small in the unoccupied states of the three crystals. In occupied states, the O(1), O(2), O(3), O(4), and O(5) atoms in [B5O10]5− and YO6 groups are still the dominant contributors to the SHG effect, while the K and Y atoms make a very small contribution. On the basis of this, we can draw the conclusion that the suitable SHG effect in the three crystals can be seen as synergistic effects of the [B5O10]5− and YO6 basic building blocks.

Accession Codes

CCDC 1576439−1576450 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 [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: + 44 1223 336033.



*[email protected] *[email protected] *[email protected] ORCID

Miriding Mutailipu: 0000-0002-1331-0185 Ying Wang: 0000-0001-6642-543X Shilie Pan: 0000-0003-4521-4507 Author Contributions #

These authors contributed equally.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by Xinjiang Key Laboratory Foundation (grant no. 2014KL009), the National Key Research Project (grant no. 2016YFB1102302, 2016YFB0402104), National Natural Science Foundation of China (grant nos. 21501194, 51425206, 51602341, 91622107), National Basic Research Program of China (grant no. 2014CB648400), and the Science and Technology Project of Urumqi (grant no. P161010003).

4. CONCLUSION Guided by the chemical cosubstitution strategy, taking the classic β-BaB2O4 as a maternal structure, we designed and synthesized a series of noncentrosymmetric borates by simultaneously replacing the Ba2+ atoms in β-BaB2O4 with multiple rare-, alkali-, and alkaline-earth metals. Three of them, namely K7CaY2(B5O10)3, K7SrY2(B5O10)3, and K7BaY2(B5O10)3, were subjected to experimental and theoretical characterizations. Results show that they all exhibit suitable SHG responses as large as ∼0.9, 1.1, and 1.2 × KDP and that the UV cutoff edges are all lower than 190 nm. First-principles calculations demonstrate that the suitable SHG response originates from the cooperative effects of the [B5O10]5− and distorted [YO6]9− basic building blocks. These results confirm the feasibility of the chemical cosubstitution strategy to design NLO materials and that the three selected crystals may have potential application as UV NLO materials.



AUTHOR INFORMATION

Corresponding Authors



REFERENCES

(1) Tran, T. T.; Yu, H. W.; Rondinelli, J. M.; Poeppelmeier, K. R.; Halasyamani, P. S. Chem. Mater. 2016, 28, 5238−5258. (2) Shi, G. Q.; Wang, Y.; Zhang, F. F.; Zhang, B. B.; Yang, Z. H.; Hou, X. L.; Pan, S. L.; Poeppelmeier, K. R. J. Am. Chem. Soc. 2017, 139, 10645−10648. (3) Jiang, X. X.; Luo, S. Y.; Kang, L.; Gong, P. F.; Huang, H. W.; Wang, S. C.; Lin, Z. S.; Chen, C. T. ACS Photonics 2015, 2, 1183− 1191. (4) Liang, F.; Kang, L.; Gong, P. F.; Lin, Z. S.; Wu, Y. C. Chem. Mater. 2017, 29, 7098−7102. (5) Yu, H. W.; Zhang, W. G.; Young, J. S.; Rondinelli, J. M.; Halasyamani, P. S. Adv. Mater. 2015, 27, 7380−7385. (6) Wei, Q.; Wang, J. J.; He, C.; Cheng, J. W.; Yang, G. Y. Chem. Eur. J. 2016, 22 (31), 10759−10762. (7) Huang, H. W.; Liu, L. J.; Jin, S. F.; Yao, W. J.; Zhang, Y. H.; Chen, C. T. J. Am. Chem. Soc. 2013, 135, 18319−18322. (8) Zhao, S. G.; Gong, P. F.; Luo, S. Y.; Liu, S. J.; Li, L. N.; Asghar, M. A.; Khan, T.; Hong, M. C.; Lin, Z. S.; Luo, J. H. J. Am. Chem. Soc. 2015, 137, 2207−2210. (9) Kong, F.; Huang, S. P.; Sun, Z. M.; Mao, J. G.; Cheng, W. D. J. Am. Chem. Soc. 2006, 128, 7750−7751. (10) Li, L.; Wang, Y.; Lei, B. H.; Han, S. J.; Yang, Z. H.; Poeppelmeier, K. R.; Pan, S. L. J. Am. Chem. Soc. 2016, 138, 9101− 9104. (11) Yu, P.; Wu, L.-M.; Zhou, L.-J.; Chen, L. J. Am. Chem. Soc. 2014, 136, 480−487. (12) Zhao, S. G.; Gong, P. F.; Luo, S. Y.; Bai, L.; Lin, Z. S.; Tang, Y. Y.; Zhou, Y. L.; Hong, M. C.; Luo, J. H. Angew. Chem., Int. Ed. 2015, 127, 4291−4295.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.7b11263. Atomic coordinates and equivalent isotropic displacement parameters, selected bond lengths for K7MIIRE2(B5O10)3 (MII = Ca, Sr, Ba, K/RE0.5; RE = Y, Lu, Gd) and K7 MIIBi2(B5O10)3 (MII = Pb, Sr); experimental powder-XRD patterns and calculated band structures for K 7 CaY 2 (B 5 O 10 ) 3 , K 7 SrY 2 (B 5 O 10 ) 3 , K7BaY2(B5O10)3; structural evolution from Ba2Mg(BO3)2 to Ba2Mg(B3O6)2 (PDF) 18404

DOI: 10.1021/jacs.7b11263 J. Am. Chem. Soc. 2017, 139, 18397−18405

Article

Journal of the American Chemical Society (13) Luo, M.; Ye, N.; Zou, G. H.; Lin, C. S.; Cheng, W. D. Chem. Mater. 2013, 25, 3147−3153. (14) Zou, G. H.; Ye, N.; Huang, L.; Lin, X. S. J. Am. Chem. Soc. 2011, 133, 20001−20007. (15) Tran, T. T.; He, J.; Rondinelli, J. M.; Halasyamani, P. S. J. Am. Chem. Soc. 2015, 137, 10504−10507. (16) Song, Y.; Ye, N.; Luo, M.; Lin, C. Chem. Mater. 2017, 29, 896− 903. (17) Chao, T.-L.; Chang, W.-J.; Wen, S.-H.; Lin, Y.-Q.; Chang, B.-C.; Lii, K.-H. J. Am. Chem. Soc. 2016, 138, 9061−9064. (18) Li, L. Y.; Li, G. B.; Wang, Y. X.; Liao, F. H.; Lin, J. H. Chem. Mater. 2005, 17, 4174−4180. (19) Huang, Q.; Hwu, S.-J. Inorg. Chem. 2003, 42, 655−657. (20) Mo, X. H.; Ferguson, E.; Hwu, S.-J. Inorg. Chem. 2005, 44, 3121−3126. (21) Becker, P. Adv. Mater. 1998, 10, 979−992. (22) Mutailipu, M.; Zhang, M.; Su, X.; Yang, Z. H.; Pan, S. L. Inorg. Chem. 2017, 56, 5506−5509. (23) Mutailipu, M.; Zhang, M.; Su, X.; Yang, Z. H.; Chen, Y. N.; Pan, S. L. Chem. - Eur. J. 2017, 23, 13910−13918. (24) Wu, H. Q.; Ju, P.; He, H.; Yang, B. F.; Yang, G. Y. Inorg. Chem. 2013, 52, 10566−10570. (25) Belokoneva, E. L. Cryst. Res. Technol. 2008, 43, 1173−1182. (26) Chen, C. T.; Wu, B. C.; Jiang, A. D.; You, G. M. Sci. Sin., Ser. B 1985, 28, 235−243. (27) Chen, C. T.; Wu, Y. C.; Jiang, A. D.; Wu, B. C.; You, G. M.; Li, R. K.; Lin, S. J. J. Opt. Soc. Am. B 1989, 6, 616−621. (28) Xia, Y. N.; Chen, C. T.; Tang, D. Y.; Wu, B. C. Adv. Mater. 1995, 7, 79−81. (29) Lin, Z. S.; Lin, J.; Wang, Z. Z.; Chen, C. T.; Lee, M. H. Phys. Rev. B 2000, 62, 1757−1764. (30) Hu, Z. G.; Yoshimura, M.; Mori, Y.; Sasaki, T. J. Cryst. Growth 2004, 260, 287−290. (31) Wang, S. C.; Ye, N. J. Am. Chem. Soc. 2011, 133, 11458−11461. (32) Zhao, S. G.; Gong, P. F.; Bai, L.; Xu, X.; Zhang, S. Q.; Sun, Z. H.; Lin, Z. S.; Hong, M. C.; Chen, C. T.; Luo, J. H. Nat. Commun. 2014, 5, 4019. (33) Zhang, M.; Su, X.; Pan, S. L.; Wang, Z.; Zhang, H.; Yang, Z. H.; Zhang, B. B.; Dong, L. Y.; Wang, Y.; Zhang, F. F.; Yang, Y. J. Phys. Chem. C 2014, 118, 11849−11856. (34) Wu, H. P.; Pan, S. L.; Poeppelmeier, K. R.; Li, H. Y.; Jia, D. Z.; Chen, Z. H.; Fan, X. Y.; Yang, Y.; Rondinelli, J. M.; Luo, H. S. J. Am. Chem. Soc. 2011, 133, 7786−7790. (35) He, R.; Lin, Z. S.; Lee, M. H.; Chen, C. T. J. Appl. Phys. 2011, 109, 103510. (36) Iwai, M.; Kobayashi, T.; Furuya, H.; Mori, Y.; Sasaki, T. Jpn. J. Appl. Phys., Part 2 1997, 36, L276−L279. (37) Aka, G.; Kahn-Harari, A.; Mougel, F.; Vivien, D.; Salin, F.; Coquelin, P.; Colin, P.; Pelenc, D.; Damelet, J. P. J. Opt. Soc. Am. B 1997, 14, 2238−2247. (38) Zhang, S. J.; Cheng, Z. X.; Zhang, S. J.; Zhang, J. G.; Han, J. R.; Chen, H. C. Chin. Phys. Lett. 1999, 16, 184−186. (39) Gravereau, P.; Chaminade, J. P.; Pechev, S.; Nikolov, V.; Ivanova, D.; Peshev, P. A. Solid State Sci. 2002, 4, 993−998. (40) Zhang, G. C.; Wu, Y. C.; Li, Y. G.; Chang, F.; Pan, S. L.; Fu, P. Z.; Chen, C. T. J. Cryst. Growth 2005, 275, 1997−2001. (41) Wu, Y. C.; Liu, J. G.; Fu, P. Z.; Wang, J.; Zhou, H.; Wang, G. L.; Chen, C. T. Chem. Mater. 2001, 13, 753−755. (42) Zhao, S. G.; Zhang, G. C.; Yao, J. Y.; Wu, Y. C. CrystEngComm 2012, 14, 5209−5214. (43) Ye, N.; Stone-Sundberg, J. L.; Hruschka, M. A.; Aka, G.; Kong, W.; Keszler, D. A. Chem. Mater. 2005, 17, 2687−2692. (44) Gheorghe, L.; Khaled, F.; Achim, A.; Voicu, F.; Loiseau, P.; Aka, G. Cryst. Growth Des. 2016, 16, 3473−3479. (45) Gao, J. H.; Li, S. Inorg. Chem. 2012, 51, 420−424. (46) Bartschke, J.; Knappe, R.; Boller, K. J.; Wallenstein, R. IEEE J. Quantum Electron. 1997, 33, 2295−2300. (47) Ok, K. M. Acc. Chem. Res. 2016, 49, 2774−2785.

(48) Zou, G. H.; Lin, C. S.; Jo, H.; Nam, G.; You, T.; Ok, K. M. Angew. Chem., Int. Ed. 2016, 55, 12078−12082. (49) Gautier, R.; Li, X. Y.; Xia, Z. G.; Massuyeau, F. J. Am. Chem. Soc. 2017, 139, 1436−1439. (50) Xia, Z. G.; Poeppelmeier, K. R. Acc. Chem. Res. 2017, 50, 1222− 1230. (51) Li, F.; Xia, Z. G.; Gong, Y.; Gu, L.; Liu, Q. L. J. Mater. Chem. C 2017, 5, 9281−9287. (52) Zhou, J.; Xia, Z. G.; Molokeev, M. S.; Zhang, X. W.; Peng, D. S.; Liu, Q. L. J. Mater. Chem. A 2017, 5, 15031−15037. (53) Zhang, H.; Zhang, M.; Pan, S. L.; Yang, Z. H.; Wang, Z.; Bian, Q.; Hou, X. L.; Yu, H. W.; Zhang, F. F.; Wu, K.; Yang, F.; Peng, Q. J.; Xu, X. Y.; Chang, K. B.; Poeppelmeier, R. P. Cryst. Growth Des. 2015, 15, 523−529. (54) Fletcher, J. G.; Glasser, F. P.; Howie, R. A. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1991, 47, 12−14. (55) Yu, H. W.; Young, J.; Wu, H. P.; Zhang, W. G.; Rondinelli, J. M.; Halasyamani, P. S. J. Am. Chem. Soc. 2016, 138, 4984−4989. (56) Zhang, B. B.; Shi, G. Q.; Yang, Z. H.; Zhang, F. F.; Pan, S. L. Angew. Chem., Int. Ed. 2017, 56, 3916−3919. (57) Cao, X. L.; Hu, C. L.; Xu, X.; Kong, F.; Mao, J. G. Chem. Commun. 2013, 49, 9965−9967. (58) Rashchenko, S. V.; Bekker, T. B.; Bakakin, V. V.; Seryotkin, Yu. V.; Shevchenko, V. S.; Kokh, A. E.; Stonoga, S. Y. Cryst. Growth Des. 2012, 12, 2955−2960. (59) SAINT, version 7.60A; Bruker Analytical X-ray Instruments, Inc., Madison, WI, 2008. (60) Sheldrick, G. M. SHELXTL, version 6.14; Bruker Analytical Xray Instruments, Inc., Madison, WI, 2008. (61) Spek, A. L. J. Appl. Crystallogr. 2003, 36, 7−13. (62) Kurtz, S. K.; Perry, T. T. J. Appl. Phys. 1968, 39, 3798−3813. (63) Perdew, J. P.; Chevary, J. A.; Vosko, S. H.; Jackson, K. A.; Pederson, M. R.; Singh, D. J.; Fiolhais, C. Phys. Rev. B 1992, 46, 6671− 6687. (64) Blaha, P.; Schwarz, K.; Madsen, G.; Kvaniscka, D.; Luitz, J. Wien2k; Vienna University of Technology, Vienna, Austria, 2001. (65) Rappe, A.; Rabe, K.; Kaxiras, E.; Joannopoulos, J. Phys. Rev. B 1990, 41, 1227−1230. (66) Akella, A.; Keszler, D. A. Mater. Res. Bull. 1995, 30, 105−111. (67) Li, R. K.; Ma, Y. Y. CrystEngComm 2012, 14, 5421−5424. (68) Li, X. Z.; Chen, X. L.; Wu, L.; Cao, Y. G.; Zhou, T.; Xu, Y. P. J. Alloys Compd. 2004, 370, 53−58. (69) Li, X. Z.; Wang, C.; Chen, X. L.; Li, H.; Jia, L. S.; Wu, L.; Du, Y. X.; Xu, Y. P. Inorg. Chem. 2004, 43, 8555−8560. (70) Wang, Y.; Pan, S. L.; Huang, S. S.; Dong, L. Y.; Zhang, M.; Han, S. J.; Wang, X. Dalton Trans. 2014, 43, 12886−12893. (71) Liu, L. L.; Yang, Y.; Huang, J. B.; Dong, X. Y.; Yang, Z. H.; Pan, S. L. Chem. - Eur. J. 2017, 23, 10451−10459. (72) Li, L. P.; Yang, Z. H.; Lei, B. H.; Kong, Q. R.; Lee, M. H.; Zhang, B. B.; Pan, S. L.; Zhang, J. Phys. Chem. Chem. Phys. 2016, 18, 6077−6084. (73) Hawthorne, F. C. Acta Crystallogr., Sect. B: Struct. Sci. 1994, B50, 481−510. (74) Salinas-Sanchez, A.; Garcia-Munoz, J. L.; Rodriguez-Carvajal, J.; SaezPuche, R.; Martinez, J. L. J. Solid State Chem. 1992, 100, 201−211. (75) Preiser, C.; Losel, J.; Brown, I. D.; Kunz, M.; Skowron, A. Acta Crystallogr., Sect. B: Struct. Sci. 1999, B55, 698−711. (76) Marvel, M. R.; Lesage, J.; Baek, J.; Halasyamani, P. S.; Stern, C. L.; Poeppelmeier, K. R. J. Am. Chem. Soc. 2007, 129, 13963−13969. (77) Neumair, S. C.; Vanicek, S.; Kaindl, R.; Többens, D. M.; Martineau, C.; Taulelle, F.; Senker, J.; Huppertz, H. Eur. J. Inorg. Chem. 2011, 2011 (27), 4147−4152. (78) Wang, J. H.; Wei, Q.; Cheng, J. W.; He, H.; Yang, B. F.; Yang, G. Y. Chem. Commun. 2015, 51, 5066−5068. (79) Chan, M. K. Y.; Ceder, G. Phys. Rev. Lett. 2010, 105, 196403. (80) Lo, C. H.; Lee, M. H. Masters Thesis; Tamkang University: Taipei, Taiwan, 2005.

18405

DOI: 10.1021/jacs.7b11263 J. Am. Chem. Soc. 2017, 139, 18397−18405