DOI:
10.1039/D4QI01044B
(Research Article)
Inorg. Chem. Front., 2024,
11, 4329-4335
Synergistic combination of different types of functional motif in Rb(NO3)(SO3NH3) for realizing excellent ultraviolet optical nonlinearity†
Received
25th April 2024
, Accepted 5th June 2024
First published on 7th June 2024
Abstract
The second harmonic generation (SHG) response, birefringence, and ultraviolet (UV) absorption edge are three crucial parameters of UV nonlinear optical (NLO) crystals. However, striking a balance among these parameters has been a challenging endeavor. The synergistic combination of mixed chromophores is regarded as an effective strategy to achieve this task, leveraging the benefits of both functional motifs. Herein, a novel UV NLO crystal, Rb(NO3)(SO3NH3), was successfully designed by integrating the π-conjugated [NO3] unit with the non-π-conjugated [SO3NH3] unit. As anticipated, the π-conjugated [NO3] units contribute to a robust SHG effect and birefringence, while the non-π-conjugated [SO3NH3] units help in widening the band gap and controlling the birefringence within an optimal range by decoupling the π-conjugated interactions and reducing the system's anisotropy. Consequently, it exhibits a strong SHG response (7 × KDP), moderate birefringence (0.07@546 nm), and a short UV absorption edge (208 nm), marking it as a promising UV NLO crystal.
Introduction
Ultraviolet (UV) nonlinear optical (NLO) crystals are vital components of solid-state lasers, holding substantial technical value in areas such as laser micro-machining and optical communication.1–4 Consequently, the pursuit of novel UV NLO crystals has long been a focal point in the realms of laser and materials science.5–7 Typically, a proficient UV NLO crystal should possess a short UV absorption edge (less than 300 nm), a potent second-harmonic generation (SHG) effect (greater than 5 times that of KDP), and a moderate birefringence (between 0.06 and 0.1).8,9 However, achieving a balance among these three critical performances often proves challenging, resulting in a dearth of practical UV NLO crystals. Hence, the design and synthesis of new UV NLO crystals with superior performance remains one of the most pressing challenges in the field.
Functional motif theory proposes that functional motifs serve as microstructural units, the composition and arrangement of which determine the properties of materials.10 Typically, three critical properties of NLO crystals – the SHG effect, birefringence, and UV absorption edge – are intimately linked with three key parameters of the microscopic functional motifs: hyperpolarizability, polarizability anisotropy, and the HOMO–LUMO gap. At present, π-conjugated units and non-π-conjugated units are frequently observed functional motifs in UV NLO crystals.11,12 Notably, π-conjugated units like [BO3], [CO3], and [NO3] usually exhibit substantial hyperpolarizabilities and polarizability anisotropies, owing to their delocalized π-bonds.11 Consequently, NLO materials with a π-conjugated system often demonstrate significant SHG effects and adequate birefringence. Nonetheless, it is crucial to note that the lower splitting energy of delocalized π-bonds means π-conjugated units tend to possess a smaller HOMO–LUMO gap than non-π-conjugated units.11 Furthermore, the large polarizability anisotropy can lead to excessive birefringence when the density of these π-conjugated units is high or coplanar, resulting in spatial walk-off effects that can significantly degrade the output beam quality.13 A prime example of this is the commercially available β-BaB2O4 (BBO) crystal. Despite its excellent SHG efficiency (>5 × KDP) and short UV absorption edge (189 nm), its excessive birefringence (0.113@1064 nm) severely limits its application in high-power output fields.14,15 In contrast, non-π-conjugated units, due to the absence of a delocalized π-bond, have a restricted electron delocalization range. This typically results in a larger HOMO–LUMO gap and smaller polarizability anisotropy and hyperpolarizability.16 While a smaller polarization anisotropy can effectively prevent spatial walk-off effects, it also makes it challenging to generate adequate birefringence, thereby limiting the phase-matching wavelength range of the material.17 Additionally, the reduced hyperpolarization of non-π conjugated systems typically leads to a smaller SHG effect compared to π conjugated systems. As such, the challenge lies in expanding the band gap and maintaining birefringence within a suitable range while ensuring a sufficient SHG effect – a difficult yet meaningful endeavor.
Based on the above ideas, we hope to design high-performance UV NLO materials by combining π conjugated and non-π conjugated units, including the following three considerations: (i) the substantial hyperpolarization and polarizability anisotropy of the π-conjugated unit can yield significant SHG effects and birefringence; (ii) the relatively lower polarizability anisotropy of the non-π-conjugated units could potentially reduce the system's anisotropy, thereby preventing the generation of excessive birefringence; (iii) as per a prior study by Chen et al., non-π-conjugated units can partially isolate the π-conjugated interactions through π-conjugation confinement, which could potentially facilitate bandgap expansion.
In the current study, we opted for the triangular π-conjugated [NO3] unit in conjunction with our previously identified polar tetrahedral non-π-conjugated [SO3NH3] unit. The [NO3] unit demonstrates a stronger hyperpolarizability and polarizability anisotropy compared to other triangular π-conjugated units such as [CO3] and [BO3] units. Conversely, the [SO3NH3] unit possesses a considerable HOMO–LUMO gap and can further amplify hyperpolarizability through hydrogen-bonding interactions with the [NO3] unit. In addition, as alkali metal ions lack d–d and f–f transitions, they are more favorable for UV light transmission. Consequently, Rb cations were selected as counterions to maintain charge equilibrium. As a result of this approach, we successfully synthesized a novel UV NLO crystal, Rb(NO3)(SO3NH3), which displays a substantial SHG effect (7 × KDP), moderate birefringence (0.07@546 nm), and a short UV absorption edge (208 nm).
Experimental section
Reagents
RbNO3 (Adamas, 99.0%), SO3NH3 (Adamas, 99.0%).
Synthesis
RbNO3 and SO3NH3 were dissolved in 5 ml H2O and evaporated in an oven at 55 °C. After 3 days, RbNO3SO3NH3 crystals were obtained.
Single-crystal X-ray diffraction
The crystal structure of RbNO3SO3NH3 was determined using SC-XRD data collected with a Rigaku Mercury CCD diffractometer and graphite-monochromated Mo-Kα radiation (λ = 0.7103 Å). The structure was solved using ShelXT and refined with ShelXL on the OLEX2 package.18 Table S1† shows the crystallographic data and structure refinements for RbNO3SO3NH3. Tables S2–S4† display the atomic coordinates, equivalent isotropic displacement parameters, selected bond lengths and angles, and anisotropic displacement parameters for RbNO3SO3NH3.
Power X-ray diffraction
The PXRD data of RbNO3SO3NH3 were obtained using a Miniflex600 instrument with Cu Kα radiation (λ = 1.540598 Å) in the 2θ range from 5 to 60°.
Energy-dispersive X-ray spectroscopy analysis
Microprobe elemental analyses were recorded on a field emission scanning electron microscope (FESEM, SU-8010) with an energy dispersive X-ray spectrometer (EDS).
Thermal analysis
Thermogravimetric analysis (TGA) was conducted using a Netzsch STA449F3 simultaneous analyzer under flowing N2 at a rate of 10 °C min−1.
UV-Vis diffuse reflectance spectroscopy
The diffuse reflectance spectrum of RbNO3SO3NH3 was measured using a PerkinElmer Lamda950 UV/vis/NIR spectrophotometer, with BaSO4 as the standard.
Birefringence measurement
The polarizing microscope was used to measure the birefringence of RbNO3SO3NH3 with 546.1 nm light. The birefringence was calculated using the following formula: | R = (|Ne − No|) × T = Δn × T | (S1) |
R denotes optical path difference, Δn represents birefringence, and T is the thickness of the crystal.
Second harmonic generation measurements
The polycrystalline SHG signals for RbNO3SO3NH3 were measured on a 1064 nm solid-state laser with KDP crystals as reference samples. The crystals were ground and sieved into six different sizes: 25–45 μm, 45–62 μm, 62–75 μm, 75–109 μm, 109–150 μm, and 150–212 μm.
Theoretical calculations
Density functional theory (DFT) was used to calculate the band structure of RbNO3SO3NH3.19 The exchange–correlation energy was treated by the generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE) functional.20 The valence electrons for the following atoms are Li 2s1; C 2s2 2p2; O 2s2 2p4; N 2s2 2p3; and H 1s1. The energy cutoff for RbNO3SO3NH3 was set at 600 eV, and the Monkhorst–Pack grid size was 2 × 2 × 2.21 The HSE06 functional was used to calculate the band gap and densities of states (DOS) for RbNO3SO3NH3.22 The NLO coefficient dij was calculated by using an expression originally proposed by Rashkeev et al. and developed by Lin et al.23
Results and discussion
RbNO3SO3NH3 crystallized in the non-centrosymmetric (NCS) space group Pmc21 (no. 26) of the orthorhombic crystal system with lattice parameters a = 5.6303(3) Å, b = 7.4552(5) Å, and c = 8.0537(5) Å. In the structure, each N atom is coordinated with three O atoms to form the triangular unit [NO3] (Fig. 1a), and each S atom is coordinated with one NH3 and three O atoms to form the polar tetrahedral unit [SO3NH3] (Fig. 1b). The [SO3NH3] unit and the [NO3] unit are connected by hydrogen bonds to form a [NO3SO3NH3] one-dimensional polar chain, in which [SO3NH3] and [NO3] units are distributed alternately (Fig. 1c). These one-dimensional polar chains are further connected by Rb ions and extended along the b and c directions, resulting in the 3D structure of RbNO3SO3NH3 (Fig. 1d). In the structure, the dipole moments of each one-dimensional chain are superimposed along the c-direction within the constraints of the RbO9 polyhedra, which may contribute to the strong SHG response (Fig. 1e).
 |
| Fig. 1 (a) [NO3] unit. (b) [SO3NH3] unit. (c) [NO3SO3NH3]− one-dimensional polar chain. 3D crystal structure of RbNO3SO3NH3 (d) in the b–c plane and (e) in the a–c plane. | |
RbNO3SO3NH3 was synthesized via the aqueous solution method. PXRD confirmed the purity of the crystals (Fig. S1†), while EDS confirmed that the crystals were composed of Rb, O, N, and S elements (Fig. S2†). Thermal analysis indicated that RbNO3SO3NH3 remained stable up to 127 °C (Fig. S3†).
Diffuse reflectance spectroscopy showed that the UV absorption edge is about 208 nm, corresponding to a bandgap of 5.96 eV (Fig. 2a). HSE06 calculations showed that RbNO3SO3NH3 has a direct band gap of 5.657 eV (Fig. 2b), matching closely with the measured value. Such a wide bandgap is significantly larger than that of RbNO3 (5.251 eV). As described above, this may be because [SO3NH3] units intercept the delocalized π–π interactions between [NO3] units, and thus enlarge the band gap. To confirm this inference, the PDOS of RbNO3 and RbNO3SO3NH3 were calculated, respectively (Fig. S4†). One can find that the conduction band minimum (CBM) of RbNO3SO3NH3 is occupied by N 2p and O 2p orbitals. Compared to RbNO3, the N–O π*–π* component of the CBM in RbNO3SO3NH3 is significantly lower, and thus pulls up the bottom of the CB.
 |
| Fig. 2 (a) UV/Vis diffuse-reflectance spectrum of RbNO3SO3NH3; (b) calculated electronic band structures for RbNO3SO3NH3 based on HSE06; (c) power SHG measurement at 1064 nm; (d) calculated SHG coefficients of RbNO3SO3NH3. | |
The SHG effect of RbNO3SO3NH3 was tested based on the Kurtz–Perry method,24 with KDP as the standard. The results showed that RbNO3SO3NH3 exhibited phase-matched behavior and SHG effects up to 7 × KDP (Fig. 2c). According to the formula deff, sample = deff, KDP(I22ω, sample/I22ω, KDP)1/2 (deff, KDP = 0.26 pm V−1),25,26 the effective SHG coefficient deff of RbNO3SO3NH3 was 1.82 pm V−1. In addition, the calculated maximum SHG coefficient d31 was −3.27 pm V−1 (Fig. 2d), which is consistent with the experimental values. According to functional motif theory, the SHG effect of RbNO3SO3NH3 should be contributed by the functional motif [NO3] and [SO3NH3] units. Based on the anion group theory,27 the SHG contribution of the [NO3] unit is directly proportional to its number density (nV−1) and structural criterion (C), while the SHG contribution of the [SO3NH3] unit is proportional to its dipole moment per unit volume. For RbNO3SO3NH3, the number density (nV−1) and the structural criterion (C) of the [NO3] unit are 0.0059 and 0.7702, respectively, and the dipole moments per unit volume are 0.022 D Å−3. These values are all comparable to those we previously reported for KNO3SO3NH3,28 and thus should be the main reason for the large SHG effects. In addition, according to our previous study, hydrogen bonding between the [SO3NH3] unit and [NO3] unit can change the electron distribution in the π-orbital, which can further enhance the SHG effect. The hyperpolarization of the [SO3NH3NO3] unit in the structure is 147, significantly higher than the sum of the hyperpolarization of the [SO3NH3] unit and the [NO3] unit. This confirms that the hydrogen-bonding interactions between the [SO3NH3] unit and [NO3] unit also contribute to the formation of large SHG effects. Interestingly, although the number density (nV−1) and structural criterion (C) of [NO3] units, as well as the dipole moments per unit volume of [SO3NH3] units in RbNO3SO3NH3, are comparable to those previously reported for KNO3SO3NH3, the hyperpolarization of [SO3NH3NO3] units in RbNO3SO3NH3 is greater than that in KNO3SO3NH3. However, RbNO3SO3NH3 exhibits a smaller SHG effect than KNO3SO3NH3. To elucidate the microscopic mechanism behind this SHG difference, the SHG weighted electron density was calculated. As shown in Fig. S5,† the maximum tensor d31 of RbNO3SO3NH3 and the maximum tensor d21 of KNO3SO3NH3 indeed are both contributed by the [NO3] units and the [SO3NH3] units. However, the contribution of the [NO3] unit to RbNO3SO3NH3 is significantly less than its contribution to KNO3SO3NH3. As a result, the maximum tensor d31 of RbNO3SO3NH3 is significantly smaller than that of KNO3SO3NH3, which may be the main reason for the SHG effect difference. In addition, since Rb ions possess a greater electron density around them than is around K ions, RbNO3SO3NH3 possesses a higher refractive index than KNO3SO3NH3 (Fig. 3c). This difference in refractive index may cause the 532 nm laser light to have a different phase-matching angle in the two crystals, which may also be one of the reasons for the SHG effect difference.
 |
| Fig. 3 (a) Single crystal of RbNO3SO3NH3 under the polarizing microscope; (b) RbNO3SO3NH3 crystal achieves complete extinction under the compensator; (c) the calculated refractive indices of RbNO3SO3NH3. | |
The birefringence of RbNO3SO3NH3 was tested using a polarized light microscope. The results indicate that the birefringence of RbNO3SO3NH3 at 546.1 nm is 0.07 (Fig. 3a and b), which is consistent with the theoretical calculations (Fig. 3c). It is noteworthy that this value is significantly higher than that of most NLO materials that only contain non-π conjugated functional motifs, such as KH2PO4 (0.035@1064 nm),29 Ba3P3O10Cl (0.03@1064 nm),30 K4Mg4(P2O7)3 (0.0108@1064 nm),31 Li2SO4·H2O (0.023@1064 nm),32 La(NH4)(SO4)2 (0.03@1064 nm),33 NaNH4PO3F·H2O (0.053@589.3 nm),34 Sr(NH2SO3)2 (0.056@589.3 nm),35 Ba(NH2SO3)2 (0.028@546.1 nm),35 and Ba(SO3CH3)2 (0.04@589.3 nm). In addition, compared with most nitrate NLO materials, such as Pb2(NO3)2(H2O)F2 (0.230@1064 nm),36 Sc(IO3)2(NO3) (0.348@546 nm),37 Pb2(BO3)(NO3) (0.174@1064 nm),38 and Cs2Pb(NO3)2Br2 (0.147@546 nm),39 RbNO3SO3NH3 has a more suitable birefringence, which is equivalent to that of KBBF (0.077@546 nm).40 In general, the optical properties of a crystal are determined by the electronic states near the forbidden band. As shown in Fig. S4,† the vicinity of the forbidden band of the title compound is occupied by the N-p and O-p states, indicating that its birefringence is entirely contributed by the [NO3] units. On the one hand, this may be due to the large polarizability anisotropy of the [NO3] unit preventing the too-small birefringence, and on the other hand, according to our previous study,41,42 this noncoplanar arrangement of π-conjugated units can effectively avoid the overlarge birefringence. These factors may be the main reasons for the moderate birefringence of the title compound. Further, the calculated refractive index dispersion curves indicate that the birefringence of RbNO3SO3NH3 is 0.086@546 nm, which is basically in agreement with the experimental results.
To evaluate the comprehensive performance of RbNO3SO3NH3, radargrams were plotted while commercial UV-NLO crystals KDP, LBO and BBO were used for comparison. As shown in Fig. 4, RbNO3SO3NH3 displays larger SHG coefficients than KDP, BBO, and LBO, while also maintaining a short UV absorption edge. More importantly, it has more suitable birefringence, which not only contributes to the phase-matching performance of the material in the UV region but also can effectively prevent the occurrence of the walk-away effect. Overall, RbNO3SO3NH3 achieves a balance between strong SHG effects, moderate birefringence, and short UV absorption edges, making it a promising UV NLO material.
 |
| Fig. 4 Radar chart (there are three directions in the radar chart, which represent the ultraviolet absorption edge, maximum SHG coefficient, and birefringence, respectively). | |
Conclusions
In summary, a novel UV NLO material, Rb(NO3)(SO3NH3), has been successfully synthesized through the integration of [NO3] and [SO3NH3] units. The synergistic interaction between these units contributes to a potent SHG response (7 × KDP). In addition, the one-dimensional chain structure composed of [NO3] and [SO3NH3] units, coupled with the reduction of system anisotropy by the [SO3NH3] units, engenders suitable birefringence (0.07@546 nm). Meanwhile, due to the spaced arrangement of [NO3] and [SO3NH3] units in the chain, the π-conjugated interactions between [NO3] units are decoupled by thr [SO3NH3] units, resulting in its short UV absorption edge of 208 nm. Consequently, Rb(NO3)(SO3NH3) strikes a balance among three key properties, marking it as a promising UV NLO crystal.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (22305038, 22222510) and the Natural Science Foundation of Fujian Province (2021J011080).
References
- N. Savage, Ultraviolet lasers, Nat. Photonics, 2007, 1, 83–85 CrossRef CAS.
- D. Cyranoski, China's crystal cache: a Chinese laboratory is the only source of a valuable crystal. David Cyranoski investigates why it won't share its supplies, Nature, 2009, 457, 953–955 CrossRef CAS PubMed.
- F. Xu, G. Zhang, M. Luo, G. Peng, Y. Chen, T. Yan and N. Ye, A powder method for the high-efficacy evaluation of electro-optic crystals, Natl. Sci. Rev., 2021, 8, nwaa104 CrossRef CAS PubMed.
- D. F. Eaton, Nonlinear Optical Materials, Science, 1991, 253, 281–287 CrossRef CAS PubMed.
- G. H. Zou and K. M. Ok, Novel ultraviolet (UV) nonlinear optical (NLO) materials discovered by chemical substitutionoriented design, Chem. Sci., 2020, 11, 5404–5409 RSC.
- X. H. Dong, H. B. Huang, L. Huang, Y. Q. Zhou, B. B. Zhang, H. M. Zeng, Z. E. Lin and G. H. 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, e202318976 CrossRef CAS PubMed.
- Q. Liu, C. Hu, X. Su, M. Abudoureheman, S. L. Pan and Z. H. Yang, LiGeBO4: a nonlinear optical material with a balance between deep-ultraviolet cut-off edge and large SHG response induced by hand-in-hand tetrahedra, Inorg. Chem. Front., 2019, 6, 914–919 RSC.
- H. X. Fan, N. Ye and M. Luo, New Functional Groups Design toward High Performance Ultraviolet Nonlinear Optical Materials, Acc. Chem. Res., 2023, 56, 3099–3109 CrossRef CAS PubMed.
- L. L. Wu, C. S. Lin, H. T. Tian, Y. Q. Zhou, H. X. Fan, S. D. Yang, N. Ye and M. Luo, Mg(C3O4H2)(H2O)2: A New Ultraviolet Nonlinear Optical Material Derived from KBe2BO3F2 with High Performance and Excellent Water-Resistance, Angew. Chem., Int. Ed., 2024, 63, e202315647 CrossRef CAS PubMed.
- X. M. Jiang, S. Deng, M. H. Whangbo and G. C. Guo, Material Research from the Viewpoint of Functional Motifs, Natl. Sci. Rev., 2022, 9, nwac017 CrossRef CAS PubMed.
- L. Xiong, L. M. Wu and L. Chen, A General Principle for DUV NLO Materials: p-Conjugated Confinement Enlarges Band Gap, Angew. Chem., Int. Ed., 2021, 60, 25063–25067 CrossRef CAS PubMed.
- H. T. Qiu, F. M. Li, C. C. Jin, Z. H. Yang, J. J. Li, S. L. Pan and M. R. D. Mutailipu, Fluorination Strategy Towards Symmetry Breaking of Boron-centered Tetrahedron for Poly-fluorinated Optical Crystals, Angew. Chem., Int. Ed., 2024, 63, e20231619 Search PubMed.
- X. H. Meng, X. Y. Zhang, Q. X. Liu, Z. Y. Zhou, X. X. Jiang, Y. G. Wang, Z. S. Lin and M. J. Xia, Nonlinear Optical Materials Perfectly Encoding π-Conjugated Anions in the RE5(C3N3O3) (OH)12 (RE = Y, Yb, Lu) Family with Strong Second Harmonic Generation Response and Balanced Birefringence, Angew. Chem., Int. Ed., 2022, e202214848 Search PubMed.
- C. T. Chen, Y. B. Wang, B. C. Wu, K. C. Wu, W. L. Zeng and L. H. Yu, Design and synthesis of an ultraviolet-transparent nonlinear optical crystal Sr2Be2B2O7, Nature, 1995, 373, 322–324 CrossRef CAS.
- L. L. Cao, H. T. Tian, D. H. Lin, C. S. Lin, F. Xu, Y. L. Han, T. Yan, J. D. Chen, B. X. Li, N. Ye and M. Luo, A flexible functional module to regulate ultraviolet optical nonlinearity for achieving a balance between a second-harmonic generation response and birefrin-gence, Chem. Sci., 2022, 13, 6990 RSC.
- H. T. Tian, C. S. Lin, X. Zhao, F. Xu, C. Wang, N. Ye and M. Luo, Ba(SO3CH3)2: A Deep-Ultraviolet Transparent Crystal with Excel-lent Optical Nonlinearity Based on a New Polar Non-π-Conjugated NLO Building Unit SO3CH3−, CCS Chem., 2023, 5, 2497–2505 CrossRef CAS.
- H. T. Tian, N. Ye and M. Luo, Sulfamide: A Promising Deep-Ultraviolet Nonlinear Optical Crystal Assembled from Polar Covalent [SO2(NH2)2] Tetrahedra, Angew. Chem., Int. Ed., 2022, 61, e202200395 CrossRef CAS PubMed.
- G. M. Sheldrick, A short history of SHELX, Acta Crystallogr., 2008, 64, 112–122 CrossRef CAS PubMed.
- W. Kohn, Nobel Lecture: Electronic structure of matter-wave functions and density functionals, Rev. Mod. Phys., 1999, 71, 1253–1266 CrossRef CAS.
- J. P. Perdew, K. Burke and M. Ernzerhof, Generalized Gradient Approximation Made Simple, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
- H. J. Monkhorst and J. D. Pack, Special points for Brillouin-zone integrations, Phys. Rev. B: Solid State, 1976, 13, 5188–5192 CrossRef.
- L. Kang, S. Y. Luo, H. W. Huang, N. Ye, Z. S. Lin, J. G. Qin and C. T. Chen, Prospects for Fluoride Carbonate Nonlinear Optical Crystals in the UV and Deep-UV Regions, J. Phys. Chem. C, 2013, 117, 25684–25692 CrossRef CAS.
- J. Lin, M. H. Lee, Z. P. Liu, C. T. Chen and C. J. Pickard, Mechanism for linear and nonlinear optical effects in beta-BaB2O4 crystals, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 60, 13380 CrossRef CAS.
- S. K. Kurtz and T. T. Perry, A power technique for evaluation of nonlinear optical materials, J. Appl. Phys., 1968, 39, 3798–3813 CrossRef CAS.
- W. F. Chen, B. W. Liu, S. M. Pei, X. M. Jiang and G. C. Guo, [K2PbX][Ga7S12] (X = Cl, Br, I): The First Lead-Containing Cationic Moieties with Ultrahigh Second-Harmonic Generation and Band Gaps Exceeding the Criterion of 2.33 eV, Adv. Sci., 2023, 10, 2207630 CrossRef CAS PubMed.
- D. Eimerl, Electro-optic, linear, and nonlinear optical properties of KDP and its isomorphs, Ferroelectrics, 1987, 72, 95–139 CrossRef CAS.
- N. Ye, Q. Chen, B. Wu and C. Chen, Searching for new nonlinear optical materials on the basis of the anionic group theory, J. Appl. Phys., 1998, 84, 555–558 CrossRef CAS.
- H. T. Tian, C. S. Lin, X. Zhao, S. H. Fang, H. Li, C. Wang, N. Ye and M. Luo, Design of a new ultraviolet nonlinear optical material KNO3SO3NH3 exhibiting an unexpected strong second harmonic generation response, Mater. Today Phys., 2022, 28, 100849 CrossRef CAS.
- Z. S. Lin, Z. Z. Wang, C. T. Chen and M. H. Lee, Mechanism of linear and nonlinear optical effects of KDP and urea crystals, J. Chem. Phys., 2003, 118, 2349 CrossRef CAS.
- P. Yu, L. M. Wu, L. J. Zhou and L. Chen, Deep-Ultraviolet Nonlinear Optical Crystals: Ba3P3O10X (X = Cl, Br), J. Am. Chem. Soc., 2014, 136, 480–487 CrossRef CAS PubMed.
- H. Yu, J. Young, H. Wu, W. Zhang, J. M. Rondinelli and P. S. Halasyamani, M4Mg4(P2O7)3 (M = K, Rb): Structural Engineering of Pyrophosphates for Nonlinear Optical Applications, Chem. Mater., 2017, 29, 1845–1855 CrossRef CAS.
- P. Becker, S. Ahrweiler, P. Held, H. Schneeberger and L. Bohatý, Thermal expansion, pyroelectricity and linear optical properties of Li2SeO4·H2O and Li2SO4·H2O, Cryst. Res. Technol., 2003, 38, 881–889 CrossRef CAS.
- C. Wu, X. X. Jiang, Y. L. Hu, C. B. Jiang, T. H. Wu, Z. S. Lin, Z. P. Huang, M. G. Humphrey and C. Zhang, A Lanthanum Ammonium Sulfate Double Salt with a Strong SHG Response and Wide Deep-UV Transparency, Angew. Chem., Int. Ed., 2022, 61, e202115855 CrossRef CAS PubMed.
- J. Lu, J. N. Yue, L. Xiong, W. K. Zhang, L. Chen and L. M. Wu, Uniform alignment of non-π-conjugated species enhances
deep ultraviolet optical nonlinearity, J. Am. Chem. Soc., 2019, 141, 8093–8097 CrossRef CAS PubMed.
- X. Hao, M. Luo, C. S. 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., 2020, 60, 7621–7625 CrossRef PubMed.
- G. Peng, Y. Yang, Y. H. Tang, M. Luo, T. Yan, Y. Q. Zhou, C. S. Lin, Z. S. Lin and N. Ye, Collaborative enhancement from Pb2+ and F− in Pb2(NO3)2(H2O)F2 generates the largest second harmonic genera-tion effect among nitrates, Chem. Commun., 2017, 53, 9398–9401 RSC.
- C. Wu, X. X. Jiang, X. J. Wang, Z. S. Lin, Z. P. Huang, X. F. Long, M. G. Humphrey and C. Zhang, Giant optical anisotropy in the UV-transparent 2D nonlinear optical material Sc(IO3)2(NO3), Angew. Chem., Int. Ed., 2021, 60, 3464–3468 CrossRef CAS PubMed.
- J. L. Song, C. L. Hu, X. Xu, F. Kong and J. G. Mao, A Facile Synthetic Route to a New SHG Material with Two Types of Parallel p-Conjugated Planar Triangular Units, Angew. Chem., Int. Ed., 2015, 54, 3679–3682 CrossRef CAS PubMed.
- Y. Long, X. H. Dong, H. M. Zeng, Z. E. Lin and G. H. Zou, Layered perovskite-like nitrate Cs2Pb(NO3)2Br2 as a multifunctional optical material, Inorg. Chem., 2022, 61, 4184–4192 CrossRef CAS PubMed.
- C. T. Chen, G. L. Wang, X. Y. Wang and Z. Y. Xu, Deep-UV nonlinear optical crystal KBe2BO3F2-discovery, growth, optical properties and applications, Appl. Phys. B, 2009, 97, 9–25 CrossRef CAS.
- H. T. Tan, C. S. Lin, Y. Q. Zhou, X. Zhao, H. X. Fan, T. Yan, N. Ye and M. Luo, Design of the Ionic Organic Nonlinear Optical Material NH4[LiC3H(CH3)O4] with Ultrawide Band Gap and Moderate Birefringence, Angew. Chem., Int. Ed., 2023, e202304858 Search PubMed.
- H. T. Tian, C. S. Lin, B. X. Li, X. Zhao, T. Yan, N. Ye and M. Luo, Designing Strong Polarity and High Configurational Entropy Flexible Units toward Excellent Ultraviolet Nonlinear Optical Materials, Adv. Funct. Mater., 2024, 2402295 CrossRef.
Footnote |
† Electronic supplementary information (ESI) available: Crystallographic and structure data (Tables S1–S4), and figures containing the measurement results and detail of RbNO3SO3NH3 (Fig. S1–S5). CCDC 2351069 for RbNO3SO3NH3. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4qi01044b |
|
This journal is © the Partner Organisations 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.