DOI:
10.1039/D5QI01311A
(Research Article)
Inorg. Chem. Front., 2025, Advance Article
Y(C6H5SO3)3·9H2O: strategic integration of ligand substitution and π-conjugated group linkage in sulfates for enhanced ultraviolet nonlinear optical properties
Received
16th June 2025
, Accepted 28th July 2025
First published on 19th August 2025
Abstract
Ultraviolet (UV) nonlinear optical (NLO) materials are critical for laser frequency conversion to extend wavelength ranges. Sulfates have emerged as a promising family of UV NLO materials, yet their NLO performance is constrained by the low optical anisotropy of the constituent [SO4]2− groups. In this study, we investigate the [C6H5SO3]− group, formed by substituting a ligand of the non-π-conjugated [SO4]2− unit with a π-conjugated benzene ring, and demonstrate that this group has a strong second harmonic generation (SHG) response and large birefringence. Leveraging this optimized structural unit, we successfully synthesize the first UV NLO benzenesulfonate, Y(C6H5SO3)3·9H2O. This compound exhibits enhanced UV optical properties, including a short cutoff edge of 280 nm, a relatively large SHG efficiency (0.8 × KDP), and a large birefringence of 0.16 at 1064 nm. This work not only presents Y(C6H5SO3)3·9H2O as a competitive UV NLO material, but also highlights the [C6H5SO3]− group as a highly effective structural motif for the future development of advanced UV NLO materials.
Introduction
Nonlinear optical (NLO) crystals are indispensable materials for coherent laser frequency conversion, playing a pivotal role in the rapid development of laser technology and its diverse applications, including ultrafine spectroscopy, entangled photon pair generation, environmental monitoring, and explosive detection.1–3 To date, commercial NLO materials such as LBO (LiB3O5),4 β-BBO (β-BaB2O4),5 KTP (KTiOPO4),6 KDP (KH2PO4)7 and LN (LiNbO3)8 have demonstrated excellent performance, characterized by suitable transparency ranges, large second-harmonic generation (SHG) responses, and moderate birefringence. These properties stem from the well-arranged NLO-active units within their structures, as elucidated by, e.g., anionic group theory.9 However, as laser technology continues to evolve, the discovery of new NLO crystals with good performance has been imperative. Yet, surpassing the performance of commercial crystals using traditional structural units remains a great challenge. Thus, the development of new NLO active units is crucial for the exploration of next-generation NLO materials.10–15
Very recently, the non-π-conjugated [SO4]2− unit has garnered significant attention owing to its ability to achieve wide bandgaps, strong SHG responses and rich structural diversity. For example, Zhao et al. reported NH4NaLi2(SO4)2,16 which exhibits a bandgap of 5.5 eV and a SHG response of 1.1 × KDP. Similarly, Zou et al. demonstrated CsSbF2SO4
17 with a bandgap of 4.76 eV and a SHG response of 3 × KDP. Despite their promising properties, materials whose anionic groups are exclusively composed of non-π-conjugated [SO4]2− groups typically exhibit low optical anisotropy. This is attributed to their nearly isotropic responses to external optical fields. This limitation typically results in small birefringence (∼0.02), which would hinder the achievement of short phase matching wavelengths.
To address the challenge of low birefringence in sulfates, two primary strategies – typically employed separately – have been developed. The first strategy involves the incorporation of additional units with large structural anisotropy while maintaining the integrity of [SO4]2− groups. For instance, planar π-conjugated groups, e.g., the [BO3]3− group, and stereochemically active lone pair (SCALP) cation-centred units, e.g., the [SbO3Cl2] group, have been successfully introduced into sulfates, achieving large birefringence values. Notable examples include (NH4)2B4SO10 (0.056 at 1064 nm)18 reported by Pan et al. and RbSbSO4Cl2 (0.112 at 1064 nm)19 reported by Zou et al. The second strategy focuses on enhancing polarizability anisotropy by modifying the ligands of the [SO4]2− tetrahedron. Partial substitution of oxygen atoms in [SO4]2− with other atoms or groups can disrupt the symmetry of the tetrahedron and deform its electron cloud structure, thereby improving polarizability anisotropy. This approach has led to the development of units such as [SO3F]−, [SO3CH3]−, [SO3CF3]− and [SO3NH2]−, which exhibit relatively large birefringence, enabling short phase-matching wavelengths in NLO sulfates. For example, Ye et al. reported C(NH2)3SO3F20 with a birefringence of 0.133; Lin et al. achieved a birefringence of 0.05 in CsSO3CF3
21 and Luo et al. discovered Sr(SO3NH2)2
22 with a birefringence of 0.05.
In this work, we combine the two aforementioned strategies by substituting one oxygen atom in the [SO4]2− tetrahedron with an additional group exhibiting large structural anisotropy, aiming to develop a new NLO active unit. Specifically, given its substantial hyperpolarizability, the planar π-conjugated benzene ring is chosen as the additional group. The [C6H5SO3]− group, which fuses the π-conjugated benzene ring with the non-π-conjugated [SO4]2− tetrahedron, is then constructed and subjected to theoretical investigation. This innovative group exhibits both large hyperpolarizability and significant polarizability anisotropy. Based on this optimized structural unit, we successfully synthesize the first UV NLO benzenesulfonate, Y(C6H5SO3)3·9H2O. This compound crystallizes in a non-centrosymmetric space group and exhibits a wide UV cutoff edge of 280 nm, a relatively large powder SHG response of 0.8 × KDP, a large birefringence of 0.16 at 1064 nm and a short phase matching wavelength of 300 nm, indicating its potential as a UV NLO material. Furthermore, first principles calculations are performed to elucidate its NLO mechanism, highlighting the [C6H5SO3]− group as a promising NLO active unit for further exploration and material design.
Experimental section
Reagents
Benzenesulfonic acid (98%, Aladdin) and YCl3·6H2O (97%, Aladdin) were used directly in the synthesis process without further processing.
Synthesis
Benzenesulfonic acid (7.11 g, 4.5 mmol) was dissolved in 15 mL of deionized water. YCl3·6H2O (2.925 g, 1.5 mmol) was then added to the solution, and the mixture was stirred thoroughly until complete dissolution was achieved. The solution was permitted to evaporate gradually at a controlled temperature of 25 °C, which effectively facilitated the crystallization process. After a period of ten days, the formation of single crystals of Y(C6H5SO3)3·9H2O was distinctly observed, as they precipitated and settled at the bottom of the beaker.
Single-crystal structure analysis
Single crystals of suitable dimensions (0.01 × 0.01 × 0.01 mm3) were carefully selected for X-ray diffraction (XRD) analysis. Data collection was performed using a Bruker D8 Quest diffractometer equipped with a Mo Kα radiation source (λ = 0.71073 Å). Indexing and data integration were carried out using the APEX3 software suite, employing the difference vectors method. Absorption corrections were applied using the multi-scan method implemented in the SADABS program. The space group was determined using XPREP, which is integrated into the APEX3 package.23
The crystal structures were initially solved using SHELXT and subsequently refined with SHELXL-2014, employing full-matrix least-squares refinement on F2. Anisotropic displacement parameters were included in this refinement process. The structural frameworks of Y(C6H5SO3)3·9H2O were visualized and analyzed using the OLEX2 software package. The positions of individual atoms were refined by least-squares minimization using the SHELXL-2014 refinement package.24 Finally, the PLATON program was used to verify the structural symmetry and authenticity.
Powder X-ray diffraction
The powder XRD (PXRD) measurements of Y(C6H5SO3)3·9H2O were carried out on a Bruker D8 focus diffractometer (λ = 1.5418 Å and 2θ = 5–80°). The scanning step was set to 0.05°, and the fixed counting time was set to 0.5 s per step.
UV-vis-NIR diffuse reflectance spectra
The UV-vis diffuse reflection data were recorded at room temperature using powder samples and an Agilent Cary 7000 UV/vis/NIR spectrophotometer over 200–2000 nm. BaSO4 was chosen as the back base.
Birefringence
Birefringence was measured using a Nikon Eclipse polarizing microscope E200MV POL under a visible light filter. The birefringence was calculated using the formula Δn = ΔR/T, where ΔR, Δn, and T represent the optical path difference, the birefringence, and the thickness of the crystal, respectively. A transparent strip crystal was chosen to ensure accuracy. The thickness of the sample was measured using a Bruker Smart Apex II.
Thermal analysis
Thermogravimetric analysis (TGA) was performed on a HITACHI STA 200 instrument. The crystal sample was heated in the temperature range of 30–800 °C at a heating rate of 10 °C min−1.
Powder SHG measurement
The powder SHG of Y(C6H5SO3)3·9H2O was measured using the Kurtz–Perry approach25 on a Q-switched Nd:YAG laser (1064 nm). Polycrystalline samples of Y(C6H5SO3)3·9H2O were ground into different particle size ranges (10–60, 60–100, 100–150, and 200–250 μm). Accordingly, KH2PO4 (KDP) samples with the same particle size ranges were tested for reference.
Computational method
The electronic and optical properties of the [C6H5SO3]− group and related [SO4]2− based groups, including the LUMO–HOMO (lowest unoccupied molecular orbital–highest occupied molecular orbital) gaps, hyperpolarizability (β) and polarization anisotropy (δ) were investigated using Gaussian 16 software.26,27 The calculations were performed using the B3LYP functional with the 6-31G basis set, a computational setting widely validated and employed in quantum chemical calculations.28
The CASTEP program based on density functional theory (DFT)29 was employed to calculate the linear and nonlinear optical properties in Y(C6H5SO3)3·9H2O. The exchange–correlation (XC) functional is described by the generalized gradient approximation (GGA) with Perdew, Burke and Ernzerhof (PBE) functionals.30 The ion/electron interactions are modelled using norm-conserving pseudopotentials for all the constituent elements. In this model, C 2s22p2, H 1s1, O 2s22p4, S 3s23p4 and Y 4d15s2 electrons are treated as the valence electrons. A kinetic energy cutoff of 750 eV and Monkhorst–Pack k-point meshes spanning less than 0.07 Å−3 in the Brillouin zone were chosen to ensure sufficient accuracy for the present purposes.
The SHG coefficients were calculated using the method developed by our group. In addition, the band-resolved SHG coefficient and SHG-weighted density analyses were performed to intuitively visualize the electronic orbitals and structural motifs contributing to SHG in real space.31 Our previous study revealed that the occupied-state contribution via virtual electronic transitions dominates the SHG effect in materials.32
Results and discussion
The [C6H5SO3]− group can be viewed as a benzene ring replacing one O atom in the [SO4]2− group. We theoretically investigated its LUMO–HOMO gap (ELUMO–HOMO), hyperpolarizability (β) and polarization anisotropy (δ), and compared it with those in related [SO4]2− based groups, including [SO4]2−, [SO3NH2]−, [SO3F]−, and [SO3CH3]−. Their structures and calculated results are shown in Fig. 1. As shown, the [SO4]2−, [SO3NH2]−, [SO3F]−, and [SO3CH3]− groups exhibit a tetrahedral configuration, considering the small size of the H atom, while the [SO3CF3]− group adopts a distorted tetrahedral geometry. In comparison, the [C6H5SO3]− group features a planar ring connected to a [SO3] triangular pyramid. These structural differences lead to significant variations in their optical properties. Notably, the NLO-related properties of the [C6H5SO3]− group are studied for the first time, although its polarizability anisotropy has been previously highlighted.33
 |
| Fig. 1 Geometry and key NLO properties (ELUMO–HOMO (eV), β (a.u.) and δ (a.u.)) of [SO4]2−, [SO3NH2]−, [SO3F]−, [SO3CH3]−, [SO3CF3]− and [C6H5SO3]−. | |
The results demonstrate that the [C6H5SO3]− group exhibits superior β and δ, which are approximately two and ten times larger, respectively, than those of the related tetrahedral groups. Meanwhile, a wide ELUMO–HOMO can be maintained even after the introduction of the π-conjugated benzene ring, which is conducive to applications in the UV range. These findings indicate that the incorporation of the [C6H5] moiety significantly enhances the β and δ of the [SO4]2− group, highlighting the [C6H5SO3]− group as a promising UV NLO active group.
Based on the benzenesulfonic [C6H5SO3]− unit, the first UV NLO benzenesulfonate, Y(C6H5SO3)3·9H2O, was successfully synthesized. Single crystals of Y(C6H5SO3)3·9H2O were grown using a simple solution evaporation method and a single crystal with the dimensions of 10 × 10 × 2 mm3 was obtained, as shown in Fig. 2a, which demonstrated a good crystal growth habit. Single-crystal XRD measurements were performed to solve and refine the crystal structure of Y(C6H5SO3)3·9H2O.
 |
| Fig. 2 (a) Single crystal photograph of a crystal of Y(C6H5SO3)3·9H2O. (b) Experimental and calculated XRD patterns of Y(C6H5SO3)3·9H2O. (c) Crystal structure of Y(C6H5SO3)3·9H2O. | |
Detailed structural information including crystallographic data and refinement parameters, crystal data and structure refinement, atomic coordinates and equivalent isotropic displacement parameters, anisotropic displacement parameters, selected bond lengths, selected bond angles, and hydrogen atom coordinates with isotropic displacement parameters is provided in Table 1 and Tables S1–S6.
Table 1 Crystallographic data and refinement parameters for Y(C6H5SO3)3·9H2O
|
Y(C6H5SO3)3·9H2O |
R1 = ∑||Fo| − |Fc||/∑|Fo|, wR2 = {∑w[(Fo)2 − (Fc)2]2/∑w[(Fo)2]2}1/2. |
Formula |
C18 H33 O18 S3 Y |
fw |
722.53 |
Crystal system |
Monoclinic |
Space group |
P21 |
a/Å |
11.8061(5) |
b/Å |
7.4977(2) |
c/Å |
17.4276(7) |
α (°) |
90 |
β (°) |
108.183 |
γ (°) |
90 |
V/Å3 |
1472.34(10) |
Z |
2 |
Dc (g cm−3) |
1.630 |
μ(Mo-Kα) (mm−1) |
2.27 |
GOF on F2 |
1.032 |
Flack factor |
0.043(7) |
R1, R2 (I > 2σ(I))a |
0.0406, 0.0748 |
The PXRD pattern of the title compound is shown in Fig. 2b, which is in good agreement with the calculated results of the single crystal XRD data, indicating the high purity of the obtained samples. The crystal structure of Y(C6H5SO3)3·9H2O is displayed in Fig. 2c. This compound crystallizes in the noncentrosymmetric monoclinic space group of P21 (no. 4). The asymmetric unit comprises 18, 33, 18, 3 and 1 crystallographically independent C, H, O, S and Y atoms, respectively. Six C and five H atoms constitute a benzene ring, which is further connected with [SO3] and forms a benzenesulfonic [C6H5SO3]− unit. The Y atom is coordinated by eight O atoms, two of which originate from the [C6H5SO3]− units, while the remaining six are provided by the H2O molecule. As shown in Fig. 2c, the [C6H5SO3]−groups are connected to the Y atom in two ways: one is by sharing the oxygen atom on the [C6H5SO3]− group, and the other is by hydrogen bonding mediated by water molecules nearby. The [C6H5SO3]− groups are alternately arranged on the a–c plane, forming a porous structure. The [YO8H12] groups occupy the space between coordination chains and hydrogen bonding in stabilizing the crystal structure.
Optical properties
The UV-vis diffuse spectrum of the title compound is displayed in Fig. 3a. Clearly, Y(C6H5SO3)3·9H2O exhibits good transmittance extending into the UV region. Its UV cutoff edge is located at 280 nm, corresponding to a wide band gap of 4.35 eV,34 indicating its potential for applications in the UV range. The UV cutoff edge of this compound is comparable to those of other UV NLO materials, such as Li2(H2C4N2O3)·2H2O (260 nm),35 [C(NH2)3]2SO3S (254 nm),36 and [C(NH2)3]3PO4·2H2O (250 nm),37 while it is smaller than those of (NH4)2PO3F (177 nm),38 Cs3[(BOP)2(B3O7)3] (165 nm),39 and Rb3B11P2O23 (168 nm).40
 |
| Fig. 3 (a) Transmittance of Y(C6H5SO3)3·9H2O. (b) SHG response of Y(C6H5SO3)3·9H2O. (c) Photos of a birefringence test of Y(C6H5SO3)3·9H2O. | |
The SHG response of Y(C6H5SO3)3·9H2O was evaluated using the Kurtz–Perry method under 1064 nm laser irradiation, with KDP as the reference. As shown in Fig. 3b, the SHG signal of Y(C6H5SO3)3·9H2O increases with particle size in the range from 10 to 250 μm, confirming its phase-matching capability. Finally, the SHG response reaches approximately 0.8 × KDP, a value comparable to those of Ba5P6O20 (0.8 × KDP),41 β-(C2H5N4)(NO3) (0.8 × KDP),42 and K7CaY2(B5O10)3 (1 × KDP),43 and smaller than those of Rb3Hg2(SO4)3Cl (1.5 × KDP)44 and NH4B4O6F (3 × KDP).45 It should be emphasized that, despite the large β of the [C6H5SO3]− unit, the NLO-unfavorable arrangement of these units in Y(C6H5SO3)3·9H2O may limit further enhancement of the SHG response.
The birefringence of Y(C6H5SO3)3·9H2O was measured with a polarizing microscope (Fig. 3c). The optical path difference R was found to be 880 nm and the crystal thickness T was about 6.0 μm. Using the equation Δn = ΔR/T, the birefringence was determined to be 0.147. This value exceeds those of commercial UV NLO materials, such as BBO46 and LBO,4 and is comparable to those of recently reported UV NLO materials, e.g., Zn(SCN)2 (0.164 at 546 nm),47 Na[BF2(CN)2] (0.152 at 546.1 nm),48 and Rb2SbFP2O7 (0.15 at 546 nm).49 From a structural perspective, the large birefringence can be attributed to the planar arrangement of the π-conjugated benzene rings in Y(C6H5SO3)3·9H2O.
First-principles calculations were performed to evaluate the phase-matching behavior of Y(C6H5SO3)3·9H2O (Fig. 4b). The calculated birefringence value at 1064 nm was 0.16, in good agreement with the experimental value. Based on the calculated refractive index dispersion curves, the shortest phase-matching wavelength was predicted to be 300 nm, indicating its potential for NLO applications in the UV region. According to the Kleinman symmetry restriction, Y(C6H5SO3)3·9H2O possesses four independent SHG coefficients, i.e., d14, d16, d22 and d23. The first-principles values were 0.07 pm V−1, 0.29 pm V−1, −0.50 pm V−1 and 0.04 pm V−1 for these SHG coefficients, respectively, with the maximum value of d22 = −0.50 pm V−1, which agree well with the experimental result. To further elucidate the relationship between structure and NLO performance, the electronic band structure and partial density of states (PDOS) in Y(C6H5SO3)3·9H2O were calculated, and the band-resolved analysis of the d22 coefficient was performed. As shown in Fig. 4a, Y(C6H5SO3)3·9H2O is a direct bandgap compound. The orbitals located at both the conduction band minimum and the valence band maximum are occupied by O and C from the [C6H5SO3]− group. Meanwhile, the band-resolved plot shows that the orbitals nearby the band gap play dominant roles in determining the SHG response, d22 in this case, indicating that the [C6H5SO3]− group makes a dominant contribution to the NLO properties. To further clearly show the contributions of these orbitals, the SHG weighted densities are shown in Fig. 4c. Clearly, the orbitals are mainly from the [C6H5SO3]− group. For further exploration, [C6H5SO3]− groups may be used as good UV NLO active units and both larger SHG responses and birefringence are expected if they are arranged in an orderly manner.
 |
| Fig. 4 (a) Calculated band structure, partial density of states, and band-resolved d22. (b) Calculated refractive index and birefringence. (c) The VE-occupied SHG density. | |
Conclusions
In conclusion, the integration of a π-conjugated benzene ring into a non-π-conjugated [SO4]2− tetrahedron led to the discovery of the [C6H5SO3]− group as an exceptional UV NLO active unit, characterized by its large hyperpolarizability and significant polarizability anisotropy. Capitalizing on this discovery, we successfully synthesized the first UV NLO benzenesulfonate, Y(C6H5SO3)3·9H2O. This compound exhibits a wide bandgap, a relatively strong SHG response, a significant birefringence, and a short phase-matching wavelength, collectively highlighting its potential as a promising UV NLO material. Additionally, first-principles calculations were conducted to unravel the underlying NLO mechanism. This work not only introduces a new candidate for UV NLO materials but also establishes the [C6H5SO3]− group as a high-performance UV NLO active unit for further exploration and material design.
Conflicts of interest
There are no conflicts of interest to report.
Data availability
The data supporting this article have been included as part of the SI.
TG/DTG curves of Y(C6H5SO3)3·9H2O, crystal data and structure refinement, atomic coordinates (×104) and equivalent isotropic displacement parameters (Å2 × 103), anisotropic displacement parameters (Å2 × 103), selected bond lengths, selected bond angles (deg), hydrogen atom coordinates (Å × 104) and isotropic displacement parameters (Å2 × 103) of Y(C6H5SO3)3·9H2O. CCDC 2366861 contains the supplementary crystallographic data for this paper. See DOI: https://doi.org/10.1039/d5qi01311a.
CCDC 2366861 contains the supplementary crystallographic data for this paper.50
Acknowledgements
This work is supported by the National Natural Science Foundation of China (Grant No. 22275201 and 22133004). P. G. acknowledges support from the Youth Innovation Promotion Association CAS.
References
- A. L. Cavalieri, N. Müller, T. Uphues, V. S. Yakovlev, A. Baltuška, B. Horvath, B. Schmidt, L. Blümel, R. Holzwarth, S. Hendel, M. Drescher, U. Kleineberg, P. M. Echenique, R. Kienberger, F. Krausz and U. Heinzmann, Attosecond spectroscopy in condensed matter, Nature, 2007, 449, 1029–1032 CrossRef CAS
. - T. T. Tran, H. Yu, J. M. Rondinelli, K. R. Poeppelmeier and P. S. Halasyamani, Deep Ultraviolet Nonlinear Optical Materials, Chem. Mater., 2016, 28(15), 5238–5258 CrossRef CAS
. - X. Chen, H. Jo and K. M. Ok, Lead Mixed Oxyhalides Satisfying All Fundamental Requirements for High-Performance Mid-Infrared Nonlinear Optical Materials, Angew. Chem., Int. Ed., 2020, 59(19), 7514–7520 CrossRef CAS
. - C. Chen, Y. Wu, A. Jiang, B. Wu, G. You, R. Li and S. Lin, New nonlinear-optical crystal: LiB3O5, J. Opt. Soc. Am. B, 1989, 6(4), 616–621 CrossRef CAS
. - C. Chen, Y. Wang, Y. Xia, B. Wu, D. Tang, K. Wu, Z. Wenrong, L. Yu and L. Mei, New development of nonlinear optical crystals for the ultraviolet region with molecular engineering approach, J. Appl. Phys., 1995, 77(6), 2268–2272 CrossRef CAS
. - Y. S. Liu, L. Drafall, D. Dentz and R. Belt, Nonlinear optical properties of KTiOPO4 (KTP), [in] Proceedings of the Conference on Lasers and Electro-Optics, OSA Technical Digest., 1981, WF4 Search PubMed
. - J. J. De Yoreo, Burnham A. K. and Whitman P. K. Developing KH2PO4 and KD2PO4 crystals for the world's most power laser, Int. Mater. Rev., 2002, 47(3), 113–152 CrossRef CAS
. - K. Guithi, H. E. Sekrafi, A. B. J. Kharrat, K. Khirouni and W. Boujelben, Synthesis, structural and optical characterization of LiNbO3 material for optical applications, J. Opt., 2023, 52(3), 1494–1506 CrossRef
. - C. Chen, Y. Wu and R. Li, 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(1), 65–91 Search PubMed
. - L. Wu, C. Lin, H. Tian, Y. Zhou, H. Fan, S. Yang, N. Ye and M. Luo, Angew. Chem., Int. Ed., 2024, 63, e202315647 CrossRef CAS
. - L. Wu, C. Lin, H. Tian, T. Yan, X. B. Li, H. Fan, Y. Fan, S. Yang and M. Luo, Angew. Chem., Int. Ed., 2025, 64, e202500877 CrossRef CAS PubMed
. - D. Lin, M. Luo, C. Lin, F. Xu and N. Ye, J. Am. Chem. Soc., 2019, 141(8), 3390–3394 CrossRef CAS
. - Y. Deng, L. Huang, X. Dong, L. Wang, K. M. Ok, H. Zeng, Z. Lin and G. Zou, Angew. Chem., Int. Ed., 2020, 59, 21151 CrossRef CAS
. - Y. Tian, W. Zeng, X. Dong, L. Huang, Y. Zhou, H. Zeng, Z. Lin and G. Zou, Angew. Chem., Int. Ed., 2024, 63, e202409093 CrossRef CAS
. - X. Dong, L. Huang and G. Zou, Acc. Chem. Res., 2025, 58(1), 150–162 CrossRef CAS PubMed
. - Y. Li, F. Liang, S. Zhao, L. Li, Z. Wu, Q. Ding, S. Liu, Z. Lin, M. Hong and J. Luo, Two Non-π-Conjugated Deep-UV Nonlinear Optical Sulfates, J. Am. Chem. Soc., 2019, 141(9), 3833–3837 CrossRef CAS
. - X. Dong, L. Huang, C. Hu, H. Zeng, Z. Lin, X. Wang, K. M. Ok and G. Zou, CsSbF2SO4: An Excellent Ultraviolet Nonlinear Optical Sulfate with a KTiOPO4 (KTP)-type Structure, Angew. Chem., Int. Ed., 2019, 58(20), 6528–6834 CrossRef CAS
. - Z. Li, W. Jin, F. Zhang, Z. Chen, Z. Yang and S. Pan, Achieving Short-Wavelength Phase-Matching Second Harmonic Generation in Boron-Rich Borosulfate with Planar [BO3] Units, Angew. Chem., Int. Ed., 2022, 61(4), e202112844 CrossRef CAS
. - F. He, Y. Deng, X. Zhao, L. Huang, D. Gao, J. Bi, X. Wang and G. Zou, RbSbSO4Cl2: an excellent sulfate nonlinear optical material generated due to the synergistic effect of three asymmetric chromophores, J. Mater. Chem. C, 2019, 7(19), 5748–5754 RSC
. - M. Luo, C. Lin, D. Lin and N. Ye, Rational Design of the Metal-Free KBe2BO3F2·(KBBF) Family Member C(NH2)3SO3F with Ultraviolet Optical Nonlinearity, Angew. Chem., Int. Ed., 2020, 59(37), 15978–15981 CrossRef CAS
. - B. Xu, P. Gong, F. Liu, X. Zhang, H. Huo and Z. Lin, (SO3CF3)−: A Non-π-Conjugated Motif for Nonlinear Optical Crystals Transparent into the Deep-Ultraviolet Region, Adv. Opt. Mater., 2023, 12(5), 2301725 CrossRef
. - X. Hao, M. Luo, C. Lin, G. Peng, F. Xu and N. Ye, M(NH2SO3)2 (M=Sr, Ba): Two Deep-Ultraviolet Transparent Sulfamates Exhibiting Strong Second Harmonic Generation Responses and Moderate Birefringence, Angew. Chem., Int. Ed., 2021, 60(14), 7621–7625 CrossRef CAS PubMed
. - G. M. Sheldrick, A short history of SHELX, Acta Crystallogr., Sect. A:Found. Crystallogr., 2008, 64, 112–122 CrossRef CAS PubMed
. - G. M. Sheldrick, Crystal structure refinement with SHELXL, Acta Crystallogr., Sect. C:Struct. Chem., 2015, 71, 3–8 Search PubMed
. - S. K. Kurtz and T. T. Perry, A Powder Technique for the Evaluation of Nonlinear Optical Materials, J. Appl. Phys., 1968, 39, 3798–3813 CrossRef CAS
. - R. Kumar, S. K. Yadav, R. Seth and A. Singh, Designing of gigantic first-order hyperpolarizability molecules via joining the promising organic fragments: a DFT study, J. Mol. Model., 2023, 29, 5 CrossRef CAS
. - I. Budagovsky, M. Smayev, A. Baranov, A. Kuznetsov, A. Zolot'ko and A. Bobrovsky, Optical anisotropy induced in amorphous azobenzene-containing polymers by light beams of various types, Opt. Mater., 2024, 155, 115823 CrossRef CAS
. - X. Wang, D. Toroz, S. Kim, S. L. Clegg, G.-S. Park and D. Di Tommaso, Density functional theory based molecular dynamics study of solution composition effects on the solvation shell of metal ions, Phys. Chem. Chem. Phys., 2020, 22(28), 16301–16313 RSC
. - M. D. Segall, J. D. L. Philip, M. J. Probert, C. J. Pickard, P. J. Hasnip, S. J. Clark and M. C. Payne, First-principles simulation: ideas, illustrations and the CASTEP code, J. Phys.:Condens. Matter, 2002, 14(11), 2717–2744 CrossRef CAS
. - J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J. Singh and C. Fiolhais, Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation, Phys. Rev. B:Condens. Matter Mater. Phys., 1992, 46(11), 6671–6687 CrossRef CAS
. - Z. Lin, X. Jiang, L. Kang, P. Gong, S. Luo and M.-H. Lee, First-principles materials applications and design of nonlinear optical crystals, J. Phys. D:Appl. Phys., 2014, 47(25), 253001 CrossRef
. - R. He, Z. S. Lin, M.-H. Lee and C. T. Chen, Ab initio studies on the mechanism for linear and nonlinear optical effects in YAl3(BO3)4, Jpn. J. Appl. Phys., 2011, 109(10), 103510 CrossRef
. - Z. Bai and K. M. Ok, Designing Sulfate Crystals with Strong Optical Anisotropy through pi-Conjugated Tailoring, Angew. Chem., Int. Ed., 2024, 63(4), e202315311 CrossRef CAS
. - S. P. Tandon and J. P. Gupta, Measurement of Forbidden Energy Gap of Semiconductors by Diffuse Reflectance Technique, Phys. Status Solidi B, 1970, 38(1), 363–367 CrossRef
. - D. Lin, M. Luo, C. Lin, L. Cao and N. Ye, An Optimal Arrangement of (H2C4N2O3)2− Groups in the First Non-Centrosymmetric Alkali Barbiturate Li2(H2C4N2O3)·2H2O Inducing a Giant Second Harmonic Generation Response and a Striking Birefringence, Cryst. Growth Des., 2020, 20(8), 4904–4908 CrossRef CAS
. - Y. Liu, X. Liu, Z. Xiong, B. Liu, J. Xu, L. Li, S. Zhao, Z. Lin, M. Hong and J. Luo, 2D van der Waals Layered [C(NH2)3]2SO3S Exhibits Desirable UV Nonlinear-Optical Trade-Off, Inorg. Chem., 2021, 60(19), 14544–14549 CrossRef CAS
. - X. Wen, C. Lin, M. Luo, H. Fan, K. Chen and N. Ye, [C(NH2)3]3PO4·2H2O: A new metal-free ultraviolet nonlinear optical phosphate with large birefringence and second-harmonic generation response, Sci. China Mater., 2021, 64(8), 2008–2016 CrossRef CAS
. - B. Zhang, G. Han, Y. Wang, X. Chen, Z. Yang and S. Pan, Expanding Frontiers of Ultraviolet Nonlinear Optical Materials with Fluorophosphates, Chem. Mater., 2018, 30(15), 2008–2016 Search PubMed
. - H. Liu, H. Wu, Z. Hu, J. Wang, Y. Wu and H. Yu, Cs3[(BOP)2(B3O7)3]: A Deep-Ultraviolet Nonlinear Optical Crystal Designed by Optimizing Matching of Cation and Anion Groups, J. Am. Chem. Soc., 2023, 145(23), 12691–12700 CrossRef CAS PubMed
. - S. Wang, N. Ye, W. Li and D. Zhao, Alkaline Beryllium Borate NaBeB3O6 and ABe2B3O7 (A = K, Rb) as UV Nonlinear Optical Crystals, J. Am. Chem. Soc., 2010, 132(25), 8779–8786 CrossRef CAS
. - S. Zhao, P. Gong, S. Luo, L. Bai, Z. Lin, Y. Tang, Y. Zhou, M. Hong and J. Luo, Tailored synthesis of a nonlinear optical phosphate with a short absorption edge, Angew. Chem., Int. Ed., 2015, 54(14), 4217–4221 CrossRef CAS
. - Q. Xia, X. Jiang, C. Jiang, H. Zhang, Y. Hu, L. Qi, C. Wu, G. Wei, Z. Lin, Z. Huang, M. G. Humphrey and C. Zhang, pH-Dependent Switching Between Nonlinear-Optical-Active Nitrate-Based Supramolecular Polymorphs, Angew. Chem., Int. Ed., 2025, 64, e202503136 CrossRef CAS PubMed
. - M. Mutailipu, Z. Xie, X. Su, M. Zhang, Y. Wang, Z. Yang, M. R. S. A. Janjua and S. Pan, Chemical Cosubstitution-Oriented Design of Rare-Earth Borates as Potential Ultraviolet Nonlinear Optical Materials, J. Am. Chem. Soc., 2017, 139(50), 18397–18405 CrossRef CAS
. - Y. Song, Y. Han, T. Yan and M. Luo, New Ultraviolet Nonlinear Optical Crystal Rb3Hg2(SO4)3Cl, J. Inorg. Mater., 2023, 38(7), 778–784 CrossRef CAS
. - G. Shi, Y. Wang, F. Zhang, B. Zhang, Z. Yang, X. Hou, S. Pan and K. R. Poeppelmeier, Finding the Next Deep-Ultraviolet Nonlinear Optical Material: NH4B4O6F, J. Am. Chem. Soc., 2017, 139(31), 10645–10648 CrossRef CAS
. - R. Appel, C. D. Dyer and J. N. Lockwood, Design of a broadband UV-visible α-barium borate polarizer, Appl. Opt., 2002, 41(13), 2470–2480 CrossRef CAS
. - H. Zhang, X. Jiang, Y. Zhang, K. Duanmu, C. Wu, Z. Lin, J. Xu, J. Yang, Z. Huang, M. G. Humphrey and C. Zhang, Toward Strong UV–Vis–NIR Second-Harmonic Generation by Dimensionality Engineering of Zinc Thiocyanates, J. Am. Chem. Soc., 2024, 146(41), 28329–28338 CrossRef CAS
. - M. Y. Li, X. Liu, Y. C. Liu, R. X. Wang, J. Guo, L. M. Wu and L. Chen, Unprecedented Deep Ultraviolet (DUV) Birefringence in Fluorides Constructed from Linear π-Group Anisotropic Structure Building Units, Angew. Chem., Int. Ed., 2025, 64(8), e202423054 CrossRef CAS
. - X. Dong, H. Huang, L. Huang, Y. Zhou, B. Zhang, H. Zeng, Z. Lin and G. Zou, Unearthing Superior Inorganic UV Second-Order Nonlinear Optical Materials: A Mineral-Inspired Method Integrating First-Principles High-Throughput Screening and Crystal Engineering, Angew. Chem., Int. Ed., 2024, 63(11), e202318976 CrossRef CAS
. - X. Zhang, B. Xu, D. Xiao, X. Zhang, P. Gong and Z. Lin, CCDC 2366861: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2kfxc9.
|
This journal is © the Partner Organisations 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.