Facile Growth of an Ultraviolet Hydroisocyanurate Crystal with Strong

2 days ago - Chemistry, Chinese Academy of Sciences, Beijing 100190, China. §. Physics and Space Science College, China Wes. t Normal University,...
1 downloads 0 Views 714KB Size
Communication Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Facile Growth of an Ultraviolet Hydroisocyanurate Crystal with Strong Nonlinearity and a Wide Phase-Matching Region from π‑Conjugated (HC3N3O3)2− Groups Xianghe Meng,†,‡,⊥ Fei Liang,†,‡,⊥ Kaijin Kang,§,⊥ Jian Tang,§ Tixian Zeng,*,§ Zheshuai Lin,† and Mingjun Xia*,†,‡

Downloaded via NOTTINGHAM TRENT UNIV on August 16, 2019 at 02:35:05 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



Beijing Center for Crystal Research and Development, Key Laboratory of Functional Crystals and Laser Technology, Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Beijing 100190, China § Physics and Space Science College, China West Normal University, Nanchong 637002, China ‡ Center of Materials Science and Optoelectronics Engineering, University of Chinese Academy of Sciences, Beijing 100049, China S Supporting Information *

strong optical anisotropy from delocalized π bonds.22 Subsequently, more attention has been paid to inorganic conjugated systems, such as (NO2)−, (CO3)2−, (NO3)−, and (HCOO)−, to discover superior UV NLO materials.23,24 Accordingly, some new crystals were reported including NaNO2,25 KSrCO3F,23 Sr2(OH)3NO3,24 and Na(HCOO)· H2O;26 however, no practical UV coherent radiation was achieved to our best knowledge. Therefore, it is still attractive but challenging for researchers to explore new conjugated systems for UV NLO materials.27−37 In the (B3O6)3− group, the central B atom only provides an empty pπ orbital in the sp2 hybridization configuration. In comparison, the C atom can offer an involved pπ orbital as well as an additional pπ electron. Therefore, it is natural to find a Crich inorganic ring, just like benzene, as the desirable building block for the UV NLO crystal. Hence, the planar (C3N3O3)3− anion, isoelectric with the (B3O6)3− unit, was singled out as an outstanding anionic group for both UV birefringent and NLO materials. The (C 3 N 3 O 3 ) 3− group exhibits a shorter interatomic distance and an enhanced pπ−pπ conjugated interaction in comparison with the (B3O6)3− unit, thus leading to a larger linear polarization difference and a stronger SHG response, concurrently. In 2013, the Meyer group first reported C3N3O3-based cynaurates by cyclotrimerization of linear (CNO)− groups in a closed system and high-temperature conditions.38 Subsequently, they discovered NLO crystals M3(C3N3O3)2 (M = Ca, Sr) isostructural to BBO39,40 but showing stronger SHG effects. Afterward, Liang et al. elucidated the relationship between the large π-conjugated orbitals and SHG responses in theory,41,42 which laid a solid foundation for cynaurates as UV NLO crystals. More recently, hydroisocyanurates with (HxC3N3O3)x−3 (x = 0−3) groups, maintaining the excellent gene of large πconjugated orbitals, were also expected to show a large birefringence and a strong SHG response.43 Otherwise, for hydroisocyanurates, the large single crystals with good quality were easily grown in a simple aqueous method,44−48 while cyanurate crystals synthesized in a fused quartz tube were

ABSTRACT: Mixed-alkali-metal hydroisocyanurate nonlinear-optical crystal RbLi(HC3N3O3)·2H2O was grown by a facile aqueous solution method. It exhibits a short phase-matching wavelength (λPM = 239 nm) and a strong second-harmonic-generation response (deff = 2.1KDP) benefitting from the well-aligned planar π-conjugated (HC3N3O3)2− anions, as confirmed by the first-principles calculations. Moreover, solar blind ultraviolet radiation can be reached using the title crystal via a fourthharmonic-generation technique of a Nd-based laser, i.e., 1064 nm/4 = 266 nm.

S

olar blind ultraviolet (UV) light specifies the UV ray in the wavelength range of 200−280 nm that can be absorbed by the ozone layer near the Earth’s surface.1 Solar blind UV radiation has been urgently demanded for their practical applications in the military, academic, and industry fields.2 An effective manner for obtaining coherent radiation in the solar blind UV region is through a cascading frequency conversion technique using nonlinear-optical (NLO) crystals.3−5 Accordingly, UV NLO materials with excellent optical properties are imperatively needed.6−11 Generally, a practical UV NLO crystal should simultaneously satisfy some essential criteria, including an intrinsic acentric structure, a broad UV transparent window, a large second-harmonic-generation (SHG) coefficient, and moderate birefringence for phase matching (PM).12,13 Up to now, most of the commercially available UV NLO crystals satisfying these stringent requirements are mainly borates such as β-BaB2O4 (BBO),14 LiB3O5 (LBO),15 CsB3O5 (CBO),16 KBe2BO3F2 (KBBF),17 CsLiB6O10 (CLBO),18 etc. Among them, BBO, CLBO, and KBBF crystals belong to solar blind NLO materials because of their short UV PM abilities from large optical anisotropy,19 although several other crystals also show short UV absorption edges. Especially the BBO crystal is the most famous one owing to its wide applications in fourth- and fifth-harmonic-generation Nd-based lasers and tunable Ti-sapphire fourth-harmonic-generation lasers.20,21 It features the conjugated (B3O6)3− anionic group composed of three (BO3)3− units in a coplanar arrangement by sharing terminal O atoms, which hold high SHG susceptibility and © XXXX American Chemical Society

Received: July 18, 2019

A

DOI: 10.1021/acs.inorgchem.9b02152 Inorg. Chem. XXXX, XXX, XXX−XXX

Communication

Inorganic Chemistry usually smaller in size and poorer in quality, unfavorable for practical applications.49−52 Later, several hydroisocyanurates were reported, among which noncentrosymmetric KLi(HC3N3O3)·2H2O (KLHCY)53 and Cs3Na(H2C3N3O3)4· 3H2O54 were proposed as promising potential UV NLO crystals. In order to further enrich hydroisocyanurate chemistry and target promising UV NLO materials, we selected the A1A2-HCNO system (A1 and A2 represent alkali metals) as a model and grew a mixed-alkali-metal hydroisocyanurate solar blind UV NLO crystal, RbLi(HC3N3O3)·2H2O (1). It exhibits a strong NLO effect (2.1KDP) and a short PM wavelength (λPM = 239 nm) because of the well-aligned π-conjugated anions. In addition, theoretical calculations were applied to analyze the contribution of (HC3N3O3)2− to optical anisotropy and SHG response. The single crystals of 1 were easily prepared by using LiOH· H2O, RbOH, and H3C3N3O3 as starting reagents under aqueous solution conditions. The distribution of the Rb, C, N, and O elements was certified by energy-dispersive X-ray spectroscopy (Figure S1). The phase purity of the as-grown crystals was checked by powder X-ray diffraction (XRD) refinement (Figure 1a). Thermal analysis of 1 revealed two key

Figure 2. (a) Fragment of the crystal structure of 1. (b) Wavy layered structure of [Li(HC3N3O3)O2]5− (the H atoms in H2O have been removed for clarity of the LiO5 polyhedra). (c) Planar π-conjugated (HC3N3O3)2− 6-MR group. (d and e) Coordination environments of Li and Rb.

long bond (2.613 Å; Figure 2d). Rb atoms are connected with six O atoms and one N atom in (HC3N3O3)2− and one O atom from H2O (Figure 2e). As displayed in Figure S2, the fingerprints of the IR spectra of 1 are observed. The peaks at 422 and 670 cm−1 belong to the bending vibrations of N−C−O. The bands at 1090 and 1368 cm−1 are considered to be the stretching vibrations of C− O and C−N, respectively. The peaks around 1400−1700 cm−1 are attributed to the stretching vibrations of the (HC3N3O3)2− 6-MR. In addition, the bands in the range of 2700−3500 cm−1 are assigned to the N−H and hydrogen bonds. The strongest peak in the Raman spectra at 715 cm−1 is attributed to the inplane C−O bond stretching vibration mode. The peak at 976 cm−1 is assigned to the breathing vibrations of three N atoms. The peaks at the 1426 and 1646 cm−1 modes belong to the coupling breathing and bending modes of the N−H bonds (Figure S3).33,46,54 According to the UV−vis−near-IR (NIR) reflectance spectrum, the band gap of 1 is deduced as 5.18 eV, corresponding to the UV absorption cutoff edge of 239 nm (Figure 3a). 1 has a high transmittance of up to 80% at

Figure 1. (a) Refinement plot of the powder XRD data for 1. (b) TG and DTA curves for 1.

steps of mass losses by the thermal gravimetric (TG) and differential thermal analysis (DTA) measurements (Figure 1b). The first mass loss of 14.3% happened at about 141.5 °C, which is very consistent with the calculated value of 14.1% for the loss of two crystal H2O molecules per formula unit. The next weight loss occurred at approximately 427.3 °C and accounts for the calculated mass loss of 16.7% for further decomposition of the (HC3N3O3)2− group.45 1 crystallizes in the polar space group Pna21 with unit cell parameters of a = 15.6805(9) Å, b = 3.7415(2) Å, c = 12.7706(7) Å, and Z = 4, and it is isostructural to KLHCY and different from RbLi(H2C3N3O3)2·2H2O with the space group P21/m. The detailed crystallographic data are listed in Tables S1−S5. 1 features 2D ribbon-like layers consisting of πconjugated (HC3N3O3)2− six-membered-ring (6-MR) anions, Li+ cations, and H2O molecules stacked along the a axis, with Rb+ evenly residing between the layers (Figure 2a). The ribbon-like layers are built from [Li(HC3N3O3)O2]5− sheets (Figure 2b), where (HC3N3O3)2− 6-MRs are the fundamental building motifs, with C−N and C−O bond distances of 1.336(4)−1.378(5) and 1.250(4)−1.264(4) Å, respectively (Figure 2c). Li atoms are five-coordinated to form the LiO5 polyhedron with three and two O atoms from H2O and (HC3N3O3)2− anionic groups, respectively, with four normal bond lengths of (1.951, 2.022, 2.089, and 2.158 Å) and one

Figure 3. (a) UV−vis−NIR spectrum of 1. (b) Powder SHG measurements of 1 at 1064 nm.

wavelengths longer than 255 nm and a wide transparency window over the UV-to-NIR spectral range. The band gap of 1 is comparable to those of K2Mg(H2C3N3O3)4·4H2O (5.32 eV),46 RbLi(H2C3N3O3)2·2H2O (5.16 eV),33 and Cs3Na(H2C3N3O3)4·3H2O (5.46 eV).54 This band gap is the key parameter and large enough to consider 1 as the NLO crystal B

DOI: 10.1021/acs.inorgchem.9b02152 Inorg. Chem. XXXX, XXX, XXX−XXX

Communication

Inorganic Chemistry

d33 = −1.62 pm/V). Unfortunately, the maximum component d33 is invalid for effective deff in the SHG process because of its mm2 point group. Accordingly, the simulated results are in good agreement with the powder SHG effect. In conclusion, a new solar blind NLO crystal exploration opens a marvelous chapter with the successful synthesis of 1. It exhibits a wide UV transmittance, a large SHG response, and, more importantly, moderate birefringence achieving PM wavelengths up to 239 nm from the planar π-conjugated (HC3N3O3)2− groups in the structure. Therefore, the task of exploring C3N3O3-based compounds is extremely imminent.60 Moreover, the simple method of preparing crystals in an aqueous solution is easy. Further, the bulk crystal will be grown and solar blind UV radiation at 266 nm is achieved by using the title NLO crystal via the fourth harmonic generation of Nd lasers, i.e., 1064 nm/4 = 266 nm.

used in the solar blind region. Owing to its NCS point group, 1 is expected to pursue NLO property. Taking KDP of the same particle size as the reference, the powder SHG signal intensities of 1 were determined to be under 1064 nm. As seen from Figure 3b, 1 is a type I PM compound. Also, the SHG intensity of 1 is about 2.1 times that of KDP and nearly 4 times larger than that of Cs3Na(H2C3N3O3)4·3H2O.50 To further gain deep insight into the structure−property relationships of KLHCY and 1, first-principles calculations were carried out using the CASTEP package.55 As shown in Figure S4, both are direct-band-gap semiconductors with wide forbidden gaps of around 4.70 eV, which are slightly smaller than the experimental values, with a difference of around 0.5 eV. The valence band maximum and conduction band minimum locate at the Γ point. The detailed electron states show that the near-gap bands are dominantly composed of C 2p, N 2p, and O 2p orbitals, relating to occupied nonbonding states and unoccupied anti-π orbitals (Figure S5). This case is very similar to that in other conjugated systems, such as NaNO2,56 BBO,57 and KSrCO3F.58 Moreover, linear refractive indices are simulated by ab initio density functional theory. The experimental values of KLHCY are listed in ref 53. Commonly, the band gap calculated by the Perdew−Burke− Ernzerhof (PBE) functional is underestimated owing to its intrinsic shortcomings. Thus, a scissor operator correction method is widely adopted for accurate optical property simulations, in which the correcting energy is set as the difference between the PBE and experimental gaps. However, there are also some accidental exceptions in conjugated systems. For example, in NaNO2,56 the scissor operator must be 5.0 eV, but not 1.05 eV, so as to get credible simulation results. Also, for KLHCY, the scissor operator must be set as 2.4 eV, but not 0.5 eV. As depicted in Figure 4a, the calculated



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.9b02152. Experimental synthesis, computational methods, and additional tables and figures (PDF) Accession Codes

CCDC 1938661 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.



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Fei Liang: 0000-0002-4932-1329 Zheshuai Lin: 0000-0002-9829-9893 Mingjun Xia: 0000-0001-8092-6150 Author Contributions ⊥

X.M., F.L., and K.K. contributed equally.

Notes

The authors declare no competing financial interest.



Figure 4. Refractive dispersion curves of (a) KLHCY and (b) 1.

ACKNOWLEDGMENTS This work was financially supported by the Natural Science Foundation of China (Grants 51502307, 51772304, 51890864, and U1731123), the Youth Innovation Promotion Association of the Chinese Academy of Sciences (Grant 2018035), Chunhui Plan of the Ministry of Education of China (Grant Z2016122), and the Fujian Institute of Innovation, Chinese Academy of Sciences (Grant FJCXY18030101).

refractive indices of KLHCY are basically consistent with the experimental results. The calculated birefringence of 1 is 0.18 at 514.6 nm (Figure 4b), which is slightly smaller than that of KLHCY (0.22), owing to the reduced (HC3N3O3)2− spatial density, but larger than that of BBO (0.116).14 A large birefringence is very favorable for obtaining a sufficient PM capacity in the whole transparency range of 1. Therefore, 1 can be employed as a solar blind UV NLO crystal. The corrected scissor operator is also adopted in their SHG coefficients. According to the Kleiman symmetry relations,59 they have three independent components, d31, d32, and d33. The SHG coefficients of 1 are calculated to be d31 = 0.71 pm/V, d32 = 0.42 pm/V, and d33 = −1.15 pm/V, and they are comparable to those of KLHCY (d31 = 1.04 pm/V, d32 = 0.35 pm/V, and



REFERENCES

(1) Xu, Z.; Sadler, B. M. Ultraviolet Communications: Potential and State-of-the-art. IEEE Commun. Mag. 2008, 46, 67−73. (2) Razeghi, M. Short-wavelength solar-blind detectors - Status, prospects, and markets. Proc. IEEE 2002, 90, 1006−1014. (3) Savage, N. Ultraviolet lasers. Nat. Photonics 2007, 1, 83−85.

C

DOI: 10.1021/acs.inorgchem.9b02152 Inorg. Chem. XXXX, XXX, XXX−XXX

Communication

Inorganic Chemistry

Material with Large SHG Responses. J. Mater. Chem. C 2015, 3, 5268−5274. (25) Iio, K. Nonlinear Optical Property of Sodium Nitrite. I. Second Harmonic Generation at Room Temperature. J. Phys. Soc. Jpn. 1973, 34, 138−147. (26) Ito, H.; Naito, H.; Inaba, H. New Phase-matchable Nonlinear Optical Crystals of the Formate Family. IEEE J. Quantum Electron. 1974, 10, 247−252. (27) Li, M.; Li, B.; Zhou, L.; Zhang, Y.; Cao, Q.; Wang, R.; Xiao, H. Fluorescence-sensitive adsorbent based on cellulose using for mercury detection and removal from aqueous solution with selective ″on-off″ response. Int. J. Biol. Macromol. 2019, 132, 1185−1192. (28) Shi, G.; Wang, Y.; Zhang, F.; Zhang, B.; Yang, Z.; Hou, X.; Pan, S.; Poeppelmeier, K. R. Finding the Next Deep-Ultraviolet Nonlinear Optical Material: NH4B4O6F. J. Am. Chem. Soc. 2017, 139, 10645− 10648. (29) Wang, Y.; Zhang, B.; Yang, Z.; Pan, S. Cation-Tuned Synthesis of Fluorooxoborates: Towards Optimal Deep-Ultraviolet Nonlinear Optical Materials. Angew. Chem., Int. Ed. 2018, 57, 2150−2154. (30) Luo, M.; Liang, F.; Song, Y.; Zhao, D.; Xu, F.; Ye, N.; Lin, Z. M2B10O14F6 (M = Ca, Sr): Two Noncentrosymmetric Alkaline Earth Fluorooxoborates as Promising Next-Generation Deep-Ultraviolet Nonlinear Optical Materials. J. Am. Chem. Soc. 2018, 140, 3884− 3887. (31) Zhou, Z.; Qiu, Y.; Liang, F.; Palatinus, L.; Poupon, M.; Yang, T.; Cong, R.; Lin, Z.; Sun, J. CsSiB3O7: A Beryllium-Free DeepUltraviolet Nonlinear Optical Material Discovered by the Combination of Electron Diffraction and First-Principles Calculations. Chem. Mater. 2018, 30, 2203−2207. (32) Li, Y.; Liang, F.; Zhao, S.; Li, L.; Wu, Z.; Ding, Q.; Liu, S.; Lin, Z.; Hong, M.; Luo, J. Two Non-π-Conjugated Deep-UV Nonlinear Optical Sulfates. J. Am. Chem. Soc. 2019, 141, 3833−3837. (33) Lu, J.; Yue, J. N.; Xiong, L.; Zhang, W. K.; Chen, L.; Wu, L. M. Uniform Alignment of Non-π-Conjugated Species Enhances Deep Ultraviolet Optical Nonlinearity. J. Am. Chem. Soc. 2019, 141, 8093− 8097. (34) Xiong, L.; Chen, J.; Lu, J.; Pan, C. Y.; Wu, L. M. Monofluorophosphates: A New Source of Deep-Ultraviolet Nonlinear Optical Materials. Chem. Mater. 2018, 30, 7823−7830. (35) Luo, M.; Liang, F.; Song, Y.; Zhao, D.; Ye, N.; Lin, Z. Rational Design of the First Lead/Tin Fluorooxoborates MB2O3F2 (M = Pb, Sn), Containing Flexible Two-Dimensional [B6O12F6]∞ Single Layers with Widely Divergent Second Harmonic Generation Effects. J. Am. Chem. Soc. 2018, 140, 6814−6817. (36) Huang, H.; Tu, S.; Zeng, C.; Zhang, T.; Reshak, A. H.; Zhang, Y. Macroscopic Polarization Enhancement Promoting Photo- and Piezoelectric-Induced Charge Separation and Molecular Oxygen Activation. Angew. Chem., Int. Ed. 2017, 56, 11860−11864. (37) Yu, H.; Li, J.; Zhang, Y.; Yang, S.; Han, K.; Dong, F.; Ma, T.; Huang, H. Three-in-One Oxygen Vacancies: Whole Visible-Spectrum Absorption, Efficient Charge Separation, and Surface Site Activation for Robust CO2 Photoreduction. Angew. Chem., Int. Ed. 2019, 58, 3880−3884. (38) Kalmutzki, M.; Stroebele, M.; Meyer, H. J. From Cyanate to Cyanurate: Cyclotrimerization Reactions Towards the Novel Family of Metal Cyanurates. Dalton Trans 2013, 42, 12934−12939. (39) Kalmutzki, M.; Strobele, M.; Wackenhut, F.; Meixner, A. J.; Meyer, H. J. Synthesis and SHG Properties of Two New Cyanurates: Sr3(O3C3N3)2 (SCY) and Eu3(O3C3N3)2 (ECY). Inorg. Chem. 2014, 53, 12540−12545. (40) Kalmutzki, M.; Strobele, M.; Wackenhut, F.; Meixner, A. J.; Meyer, H. J. Synthesis, Structure, and Frequency-doubling Effect of Calcium Cyanurate. Angew. Chem., Int. Ed. 2014, 53, 14260−14263. (41) Liang, F.; Kang, L.; Zhang, X.; Lee, M.; Lin, Z.; Wu, Y. Molecular Construction Using (C3N3O3)3‑ Anions: Analysis and Prospect for Inorganic Metal Cyanurates Nonlinear Optical Materials. Cryst. Growth Des. 2017, 17, 4015−4020. (42) Liang, F.; Wang, N.; Liu, X.; Lin, Z.; Wu, Y. Co-crystal LiCl· (H3C3N3O3): A Promising Solar-blind Nonlinear Optical Crystal with

(4) Atuchin, V. V.; Liang, F.; Grazhdannikov, S.; Isaenko, L. I.; Krinitsin, P. G.; Molokeev, M. S.; Prosvirin, I. P.; Jiang, X.; Lin, Z. Negative thermal expansion and electronic structure variation of chalcopyrite type LiGaTe2. RSC Adv. 2018, 8, 9946−9955. (5) Atuchin, V. V.; Subanakov, A. K.; Aleksandrovsky, A. S.; Bazarov, B. G.; Bazarova, J. G.; Gavrilova, T. A.; Krylov, A. S.; Molokeev, M. S.; Oreshonkov, A. S.; Stefanovich, S. Y. Structural and Spectroscopic Properties of New Noncentrosymmetric Self-activated Borate Rb3EuB6O12 with B5O10 Units. Mater. Des. 2018, 140, 488−494. (6) Tran, T. T.; Yu, H. W.; Rondinelli, J. M.; Poeppelmeier, K. R.; Halasyamani, P. S. Deep Ultraviolet Nonlinear Optical Materials. Chem. Mater. 2016, 28, 5238−5258. (7) You, F.; Liang, F.; Huang, Q.; Hu, Z.; Wu, Y.; Lin, Z. Pb2GaF2(SeO3)2Cl: Band Engineering Strategy by Aliovalent Substitution for Enlarging Bandgap while Keeping Strong Second Harmonic Gneration Response. J. Am. Chem. Soc. 2019, 141, 748− 752. (8) Zou, G.; Jo, H.; Lim, S. J.; You, T. S.; Ok, K. M. Rb3VO(O2)2CO3: A Four-in-One Carbonatoperoxovanadate Exhibiting an Extremely Strong Second-Harmonic Generation Response. Angew. Chem., Int. Ed. 2018, 57, 8619−8622. (9) Song, Y.; Luo, M.; Liang, F.; Ye, N.; Lin, Z. The secondharmonic generation intensification derived from localization conjugated pi-orbitals in O22‑. Chem. Commun. 2018, 54, 1445−1448. (10) Halasyamani, P. S.; Rondinelli, J. M. The must-have and niceto-have experimental and computational requirements for functional frequency doubling deep-UV crystals. Nat. Commun. 2018, 9, 2972. (11) Halasyamani, P. S.; Poeppelmeier, K. R. Noncentrosymmetric oxides. Chem. Mater. 1998, 10, 2753−2769. (12) Liang, F.; Kang, L.; Gong, P.; Lin, Z.; Wu, Y. Rational Design of Deep-Ultraviolet Nonlinear Optical Materials in Fluorooxoborates: Toward Optimal Planar Configuration. Chem. Mater. 2017, 29, 7098− 7102. (13) Wang, Y.; Pan, S. Recent Development of Metal Borate Halides: Crystal Chemistry and Application in Second-order NLO Materials. Coord. Chem. Rev. 2016, 323, 15−35. (14) Chen, C. T.; Wu, B. C.; Jiang, A. D.; You, G. M. A New-Type Ultraviolet SHG Crystal - β-BaB2O4. Sci. China B 1985, 28, 235−243. (15) Chen, C. T.; Wu, Y. C.; Jiang, A. D.; Wu, B. C.; You, G. M.; Li, R. K.; Lin, S. J. New Nonlinear-Optical Crystal: LiB3O5. J. Opt. Soc. Am. B 1989, 6, 616−621. (16) Wu, Y. C.; Sasaki, T.; Nakai, S.; Yokotani, A.; Tang, H. G.; Chen, C. T. CsB3O5: A New Nonlinear Optical Crystal. Appl. Phys. Lett. 1993, 62, 2614−2615. (17) Mei, L. F.; Wang, Y. B.; Chen, C. T.; Wu, B. C. NonlinearOptical Materials Based on MBe2BO3F2 (M = Na, K). J. Appl. Phys. 1993, 74, 7014−7015. (18) Mori, Y.; Kuroda, I.; Nakajima, S.; Sasaki, T.; Nakai, S. New Nonlinear Optical Crystal: Cesium Lithium Borate. Appl. Phys. Lett. 1995, 67, 1818−1820. (19) Chen, C. T.; Wu, Y. C.; Li, R. K. The Anionic Group Theory of the Non-Linear Optical Effect and Its Applications in the Development of New High-Quality NLO Crystals in the Borate Series. Int. Rev. Phys. Chem. 1989, 8, 65−91. (20) Nikogosyan, D. N. Beta-Barium Borate (BBO) - A Reivew of Its Properties and Applications. Appl. Phys. A: Solids Surf. 1991, 52, 359−368. (21) Wilhelm, T.; Piel, J.; Riedle, E. Sub-20-fs pulses tunable across the visible from a blue-pumped single-pass noncollinear parametric converter. Opt. Lett. 1997, 22, 1494−1496. (22) Shen, Y.; Zhao, S.; Luo, J. The role of cations in second-order nonlinear optical materials based on π-conjugated [BO3]3− groups. Coord. Chem. Rev. 2018, 366, 1−28. (23) Zou, G.; Ye, N.; Huang, L.; Lin, X. Alkaline-Alkaline Earth Fluoride Carbonate Crystals ABCO3F (A = K, Rb, Cs; B = Ca, Sr, Ba) as Nonlinear Optical Materials. J. Am. Chem. Soc. 2011, 133, 20001− 20007. (24) Huang, L.; Zou, G.; Cai, H.; Wang, S.; Lin, C.; Ye, N. Sr2(OH)3NO3: the First Nitrate as a Deep UV Nonlinear Optical D

DOI: 10.1021/acs.inorgchem.9b02152 Inorg. Chem. XXXX, XXX, XXX−XXX

Communication

Inorganic Chemistry Giant Nonlinearity From Coplanar π-conjugated Groups. Chem. Commun. 2019, 55, 6257−6260. (43) Nichol, G. S.; Clegg, W.; Gutmann, M. J.; Tooke, D. M. Stoichiometry-dependent Structures: an X-ray and Neutron Singlecrystal Diffraction Study of the Effect of Reaction Stoichiometry on the Crystalline Products Formed in the Potassium-cyanurate System. Acta Crystallogr., Sect. B: Struct. Sci. 2006, 62, 798−807. (44) Gross, P.; Höppe, H. A. The Sodium (Iso)Cyanurates Nax[H3‑xC3N3O3]·yH2O (x = 1−3, y = 0, 1): A Key-Series for Understanding the Crystal Chemistry of Metal (Iso)Cyanurates. Z. Anorg. Allg. Chem. 2019, 645, 257−266. (45) Gross, P.; Höppe, H. A. An Expedition on Alkali and AlkalineEarth Isocyanurate Hydrates: Structure Elucidation, Thermogravimetry, and Spectroscopy. Z. Anorg. Allg. Chem. 2017, 643, 1692−1703. (46) Meng, X.; Liang, F.; Kang, K.; Tang, J.; Huang, Q.; Yin, W.; Lin, Z.; Xia, M. A Rich Structural Chemistry in π-conjugated Hydroisocyanurates: Layered Structures of A2B(H2C3N3O3)4·nH2O (A = K, Rb, Cs; B = Mg, Ca; n = 4, 10) with High Ultraviolet Transparency and Strong Optical Anisotropy. Dalton Trans 2019, 48, 9048−9052. (47) Meng, X.; Liang, F.; Tang, J.; Kang, K.; Zeng, T.; Yin, W.; Guo, R.; Lin, Z.; Xia, M. Parallel Alignment of π-Conjugated Anions in Hydroisocyanurates Enhancing Optical Anisotropy. Inorg. Chem. 2019, 58, 8948−8952. (48) Wang, N.; Liang, F.; Yang, Y.; Zhang, S.; Lin, Z. A New Ultraviolet Transparent Hydra-cyanurate K2(C3N3O3H) with Strong Optical Anisotropy from Delocalized π-bonds. Dalton Trans 2019, 48, 2271−2274. (49) Xia, M.; Zhou, M.; Liang, F.; Meng, X.; Yao, J.; Lin, Z.; Li, R. Noncentrosymmetric Cubic Cyanurate K6Cd3(C3N3O3)4 Containing Isolated Planar π-Conjugated (C3N3O3)3‑ Groups. Inorg. Chem. 2018, 57, 32−36. (50) Tang, J.; Liang, F.; Meng, X.; Kang, K.; Yin, W.; Zeng, T.; Xia, M.; Lin, Z.; Yao, J.; Zhang, G.; Kang, B. Ba3(C3N3O3)2: A New Phase of Barium Cyanurate Containing Parallel π-Conjugated Groups as a Birefringent Material Replacement for Calcite. Cryst. Growth Des. 2019, 19, 568−572. (51) Li, Z.; Liang, F.; Guo, Y.; Lin, Z.; Yao, J.; Zhang, G.; Yin, W.; Wu, Y.; Chen, C. Ba2M(C3N3O3)2 (M = Mg, Ca): Potential UV Birefringent Materials with Strengthened Optical Anisotropy Originating from the (C3N3O3)3‑ Group. J. Mater. Chem. C 2018, 6, 12879−12887. (52) Kang, K.; Liang, F.; Meng, X.; Tang, J.; Zeng, T.; Yin, W.; Xia, M.; Lin, Z.; Kang, B. Ba2M(C3N3O3)2 (M = Sr, Pb): Band Engineering from p-π Interaction via Homovalent Substitution in Metal Cyanurates Containing Planar pi-Conjugated Groups. Inorg. Chem. 2019, 58, 9553−9556. (53) Lin, D.; Luo, M.; Lin, C.; Xu, F.; Ye, N. KLi(HC3N3O3)·2H2O: Solvent-drop Grinding Method toward the Hydro-isocyanurate Nonlinear Optical Crystal. J. Am. Chem. Soc. 2019, 141, 3390−3394. (54) Meng, X.; Liang, F.; Tang, J.; Kang, K.; Huang, Q.; Yin, W.; Lin, Z.; Xia, M. Cs3Na(H2C3N3O3)4·3H2O: Mixed Alkali-metal Hydro-isocyanurate Nonlinear Optical Material Containing πConjugated Six-Membered Ring Units. Eur. J. Inorg. Chem. 2019, 2019, 2791−2795. (55) Clark, S. J.; Segall, M. D.; Pickard, C. J.; Hasnip, P. J.; Probert, M. J.; Refson, K.; Payne, M. C. First Principles Methods Using CASTEP. Z. Kristallogr. - Cryst. Mater. 2005, 220, 567−570. (56) Lin, Z. S.; Wang, Z. Z.; Chen, C. T.; Lee, M. H. Calculations for the linear and nonlinear optical coefficients of NaNO2 crystal. Acta. Phys. Sin. 2001, 50, 1145−1149. (57) Lin, J.; Lee, M. H.; Liu, Z. P.; Chen, C. T.; Pickard, C. J. Mechanism for Linear and Nonlinear Optical Effects in β-BaB2O4 Crystals. Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 60, 13380− 13389. (58) Kang, L.; Luo, S.; Huang, H.; Ye, N.; Lin, Z.; Qin, J.; Chen, C. Prospects for Fluoride Carbonate Nonlinear Optical Crystals in the UV and Deep-UV Regions. J. Phys. Chem. C 2013, 117, 25684− 25692.

(59) Kleinman, D. A. Nonlinear Dielectric Polarization in Optical Media. Phys. Rev. 1962, 126, 1977−1979. (60) Xia, M.; Liang, F.; Meng, X.; Wang, Y.; Lin, Z. Intrinsic Zero Thermal Expansion in Cube Cyanurate K6Cd3(C3N3O3)4. Inorg. Chem. Front. 2019, DOI: 10.1039/C9QI00709A.

E

DOI: 10.1021/acs.inorgchem.9b02152 Inorg. Chem. XXXX, XXX, XXX−XXX