DOI:
10.1039/D4QI01515K
(Research Article)
Inorg. Chem. Front., 2024,
11, 5587-5597
Halide-driven polarity tuning and optimized SHG-bandgap balance in (C4H11N2)ZnX3 (X = Cl, Br, I)†
Received
15th June 2024
, Accepted 16th July 2024
First published on 16th July 2024
Abstract
In the quest for organic–inorganic metal halides (OIMHs) that harmoniously combine large second harmonic generation (SHG) efficiency with broad bandgaps, our study introduces a series of noncentrosymmetric and polar piperazine zinc halides, (C4H11N2)ZnX3 (X = Cl, Br, I). The distinct influence of halide anion size on the configuration of ZnNX3 tetrahedra is uncovered, revealing a design principle for enhancing SHG activity and manipulating bandgap characteristics. Specifically, (C4H11N2)ZnCl3 is the first example of deep-ultraviolet (deep-UV) nonlinear optical (NLO) OIMHs, with UV transparency lower than 190 nm and moderate SHG effect, 0.8 times KH2PO4 (KDP). Meanwhile, compared with (C4H11N2)ZnI3 (2.1 times KDP, 4.52 eV), (C4H11N2)ZnBr3 boasts a widened bandgap of 5.53 eV while maintaining a striking SHG response (2.5 times KDP), representing a pinnacle in the SHG-bandgap balance among OIMHs with Eg exceeding 5.0 eV. Computational analyses underscore the critical roles of halide orbitals and ZnNX3 distortions in dictating SHG efficiency, with increasing halide polarizability correlating with heightened SHG contributions. This study paves the way for a novel approach to designing short-UV NLO crystals based on OIMHs.
Introduction
The interplay of light and matter at the nexus of laser physics and optical materials science unveils captivating phenomena exemplified by second harmonic generation (SHG).1–7 This effect empowers nonlinear optical (NLO) crystals to generate a vast laser spectrum, spanning from far-infrared to the elusive deep-ultraviolet (deep-UV, λ < 200 nm) realm. Lasers penetrating the short or even deep-UV regions hold immense potential for ultrafast spectroscopy, high-density information storage, and precision instrument fabrication.8–17 However, despite their alluring capabilities, commercialization of UV NLO crystals remains sluggish, dominated by established materials like KH2PO4 (KDP), β-BaB2O4 (BBO), and LiB3O5 (LBO). This lag stems from stringent performance requirements for efficient UV laser output. Ideal NLO crystals for this domain must exhibit NCS space groups, robust SHG coefficients (dij > KDP), wide bandgaps for transparency in the short or deep-UV region (Eg > 5.0 eV), moderate birefringence for phase matching (Δn > 0.04), and favorable crystal growth habits.18–25
Anionic groups play a pivotal role in designing NLO crystals. Planar π-conjugated (BO3)3− and (B3O6)3− anions, recognized by anionic group theory, exhibit high anisotropy, large first-order hyperpolarizability, and wide HOMO–LUMO bandgaps – key factors for outstanding NLO properties.26–31 Non-π-conjugated tetrahedra like (PO4)3− and (SO4)2−, with their wider HOMO–LUMO gaps, also contribute as NLO building blocks, exemplified by commercially available KH2PO4.32–35 Recent discoveries have yielded novel sulfate- and phosphate-based NLO crystals, such as NH4NaLi2(SO4)2 (1.1 × KDP),36 Ba3P3O10Cl (0.5 × KDP),37 RbNaMgP2O7 (1.5 × KDP),38 Li9Na3Rb2(SO4)7 (1.3 × KDP),39 and ZrF2(SO4) (3.2 × KDP).40 However, tetrahedral groups often suffer from low anisotropy and polarizability due to their high symmetry, leading to challenges within NCS space groups and minimal birefringence (Δn < 0.01).32–35 To address this limitation, researchers have explored disrupting the high symmetry of tetrahedra by partially substituting O atoms with other elements.3,9,10,41 This strategy, exemplified by the enhanced birefringence of LiSO3F (0.057) compared to Li2SO4 (0.004), demonstrates the effectiveness of introducing distortion.34 Additionally, S–O–Org units, where Org represents an organic group, have emerged as promising SHG functional units, leading to NLO crystals,42–46 like Ba(SO3CH3)2 (1.5 × KDP),45 Cs2Mg(NH2SO3)4·4H2O (2.3 × KDP),42 and SO2(NH2)2 (4 × KDP).44
On the other hand, it has been observed that the creation of M–X–Org (M: metal; X: halide; Org: organic group) asymmetric units can yield NLO OIMHs with a strong SHG effect, such as (C4H11N2)CdI3 (6 × KDP).47–49 Furthermore, in comparison to SO3F or S–O–Org units, M–X–Org units are more readily obtained at the experimental level, often through the controlled utilization of HX based on previously obtained MX4-containing OIMHs.47 However, the bandgap of (C4H11N2)CdI3, 4.10 eV, is insufficient for short UV NLO crystal applications.47 Besides, in the case of [C6H5(CH2)4NH3]4BiX7·H2O, the wider bandgap of the Br compound (3.52 eV) compared to the isomorphous I compound (2.29 eV) does not maintain a stronger SHG effect.50 In fact, the SHG effect of the Br compound (0.4 × KDP) is significantly smaller than that of the I compound (1.3 × AGS), in which KDP: d33 = 0.39 pm V−1 and AGS: d36 = 12.6 pm V−1. This reveals that the electronegativity-driven halogen substitution strategy does not ensure both a wide bandgap (Eg > 5.0 eV) and a strong SHG response (dij > KDP), thereby presenting a challenge in the exploration of high performance NLO in short UV band.47–58
To address this challenge, our efforts have been concentrated on the piperazine–Zn–Cl/Br/I system for several reasons. Firstly, the piperazine cation exhibits a wider HOMO–LUMO gap compared to the pyrazine cation (Fig. S1†), despite having a similar N-hybridized structure but lacking a π-conjugated configuration. Furthermore, d10 transition metals such as Zn2+, Cd2+, and Hg2+ only display weak d–d electron transitions, indicating good transparency in the UV regions. And, in contrast to toxic Cd2+ and Hg2+, Zn2+ is an environmentally friendly cation. Secondly, the Zn2+ cation is a typical Soft Lewis Acid, while the N-hybridized piperazine is a Soft Lewis Base. Therefore, they readily form a Zn–N coordination bond, constructing polarizable Zn–X–PIP groups, which can serve as SHG functional units. Finally, it is noteworthy that replacing iodide with the smaller radius bromide can induce changes in both the polar axis and dipole moment.59 Consequently, we have grounds to believe that halogen substitution may not only impact optical performance due to electronegativity differences but also potentially enhance the arrangement of SHG motifs through size effects, thereby influencing SHG performance. As a result, the size effect of halides could counterbalance the significant decrease in SHG effect resulting from variances in bandgap during halide substitution.
As a result, we have successfully synthesized three novel piperazine zinc halides, namely (C4H11N2)ZnX3 (X = Cl, Br, I). Notably, the UV cutoff edge of (C4H11N2)ZnCl3 falls below 190 nm, and it exhibits a moderate SHG effect at 0.8 times that of KDP, marking it as the first reported deep-UV NLO OIMHs. Additionally, when compared to (C4H11N2)ZnI3 (4.52 eV, 2.1 × KDP), (C4H11N2)ZnBr3 features an unprecedented balance between bandgap and SHG response, 5.53 eV and 2.5 × KDP, the largest value among all reported OIMHs with bandgap wider than 5.0 eV. This unusual relationship is primarily attributed to the ideally arranged ZnNBr3, whereas ZnNI3 not.
Experiments
Materials and synthesis
ZnO (>99%), La2O3 (>99%), Y2O3 (>99%), piperazine (>99%), HCl (36–38%, AR), HBr (48%), and HI (55–58 wt%) were used as purchased from Adamas-beta. For the synthesis of (C4H11N2)ZnCl3, the following starting materials were used: ZnO (162.78 mg, 2 mmol), piperazine (258.42 mg, 3 mmol), La2O3 (325.81 mg, 1 mmol), HCl (2 mL), and H2O (1 mL). The mixture was placed in Teflon pouches (23 mL), sealed in an autoclave, heated at 110 °C for 72 hours, and then cooled to 30 °C at a rate of 1.67 °C h−1. This procedure yielded colorless block-shaped crystals of (C4H11N2)ZnCl3 with approximately 95% yield based on Zn. For the synthesis of (C4H11N2)ZnBr3, the starting materials included ZnO (81.39 mg, 1 mmol), piperazine (129.21 mg, 1.5 mmol), HBr (0.3 mL), and H2O (2 mL). The synthesis followed the same procedure, resulting in colorless block-shaped crystals of (C4H11N2)ZnBr3 with an approximate yield of 96% based on Zn. For the synthesis of (C4H11N2)ZnI3, the starting materials comprised ZnO (40.69 mg, 0.5 mmol), piperazine (172.28 mg, 2 mmol), Y2O3 (112.50 mg, 0.5 mmol), HI (1 mL), and H2O (2 mL). The synthesis procedure was consistent with the aforementioned methods, leading to colorless block-shaped crystals of (C4H11N2)ZnI3 with an estimated yield of approximately 94% based on Zn.
Single crystal structure determination
Single-crystal X-ray diffraction data for the title compounds were collected on a Rigaku XtaLAB Synergy-DW dual-wavelength CCD diffractometer with Cu Kα radiation (λ = 1.54184 Å) at 109 K and 298 K. Data reduction was performed using CrysAlisPro, with absorption correction applied via the multi-scan method.60 The structures of (C4H11N2)ZnCl3, (C4H11N2)ZnBr3, and (C4H11N2)ZnI3 were determined by direct methods and refined by full-matrix least-squares fitting on F2 using SHELXL-2014.61 All non-hydrogen atoms were refined with anisotropic thermal parameters. The structures were checked for missing symmetry elements using PLATON, and none were found.62 The Flack parameters for (C4H11N2)ZnCl3, (C4H11N2)ZnBr3, and (C4H11N2)ZnI3 are 0.00(4), −0.06(9), and 0.053(14), respectively, confirming the correctness of their absolute structures. Additionally, twinning was observed in (C4H11N2)ZnBr3 and (C4H11N2)ZnI3, with twin laws of (−1.0, 0.0, 0.0, 0.0, −1.0, 0.0, 0.0, 0.0, −1.0). Crystallographic data, structural refinements, and crystal information of the title compounds are listed in Tables S1–S5.†
Powder X-ray diffraction
Powder X-ray diffraction (PXRD) patterns were recorded on a Rigaku Ultima IV diffractometer with graphite-monochromated Cu Kα radiation in the 2θ range of 10–70°, with a step size of 0.02°.
Thermal analysis
Thermogravimetric analysis (TGA) was conducted using a Rigaku TG-DTA 8121 unit under an Ar atmosphere, with a heating rate of 10 °C min−1 in the range of 30 to 650 °C.
Optical measurements
Ultraviolet–visible (UV-vis) spectra in the range of 200–800 nm were recorded on a PerkinElmer Lambda 750 UV-vis-NIR spectrophotometer. The reflectance spectrum was converted into an absorption spectrum using the Kubelka–Munk function.63
The infrared (IR) spectrum was recorded on a Thermo Fisher Nicolet 5700 FT-IR spectrometer in the form of KBr pellets in the range from 4000 to 400 cm−1.
Second harmonic generation measurements
Powder SHG measurements were performed using a Q-switched Nd:YAG laser generating radiation at 1064 nm, following the method of Kurtz and Perry.64 Crystalline (C4H11N2)ZnCl3, (C4H11N2)ZnBr3, (C4H11N2)ZnI3 samples were sieved into distinct particle-size ranges (45–53, 53–75, 75–105, 105–150, 150–210, and 210–300 μm). Sieved KH2PO4 (KDP) samples in the same particle-size ranges were used as references.
Elemental analysis
The elemental content was measured using a Vario EL Cube elemental analyzer from Elementar Analysensysteme GmbH, Germany. The combustion temperature was 800 °C.
Energy-dispersive X-ray spectroscope
Microprobe elemental analyses and elemental distribution maps were obtained using a field-emission scanning electron microscope (Phenom LE) equipped with an energy-dispersive X-ray spectroscope (EDS, Phenom LE).
Computational method
The electronic structures were calculated using the plane-wave pseudopotential method within density functional theory (DFT) implemented in the CASTEP code.65,66 For the exchange–correlation functional, we selected the Perdew–Burke–Ernzerhof (PBE) formulation within the generalized gradient approximation (GGA).67 The interactions between the ionic cores and the electrons were described by norm-conserving pseudopotentials.68 The following orbital electrons were treated as valence electrons: Cl-3s23p5, Br-4s24p5, I-5s25p5, Zn-3p54s23d10, H-1s1, C-2s22p2, and N-2s22p3. The basis set was determined using a cutoff energy of 750 eV. Monkhorst–Pack k-point sampling grids of 4 × 5 × 4, 4 × 5 × 4, and 1 × 1 × 1 were used for the numerical integration of the Brillouin zone.
The calculations of second-order nonlinear optical (NLO) susceptibilities were based on the length-gauge formalism within the independent particle approximation.69,70 The second-order NLO susceptibility can be expressed as:
where the subscript
L denotes the length gauge, and
χmodabc,
χinterabc, and
χintraabc represent the contributions to
χabc from interband processes, intraband processes, and the modulation of interband terms by intraband terms, respectively.
Results and discussion
Single crystals of (H11C4N2)ZnX3 (X = Cl, Br, I) were prepared through simple hydrothermal reactions. In designing the H2O to HX ratio, we referred to the preparation conditions of (C4H12N2)ZnX4 (X = Cl, Br), which utilized 0 mL H2O and 3 mL HX, as previously reported by our research group.71 We found that only when the H2O to HX ratio exceeds 2 does it facilitate the formation of coordination bonds between Zn2+ cations and halide anions. Powder XRD results indicate that the obtained samples are phase-pure (Fig. S2†). EDS analysis reveals that the Zn
:
X ratios in the Cl, Br, and I compounds are 1
:
2.86, 1
:
2.99, and 1
:
3.03, respectively (Fig. S3†). Elemental analysis of C, N, and H atoms was performed. For (C4H11N2)ZnCl3, the weight percentages were C 6.77%, H 1.63%, and N 4.02%. For (C4H11N2)ZnBr3, the weight percentages were C 8.30%, H 1.85%, and N 4.53%. For (C4H11N2)ZnI3, the weight percentages were C 8.30%, H 1.85%, and N 4.53%. IR spectra are provided in Fig. S4,† and the detailed assignments of the absorption peaks (Table S6†) are comparable with previously reported compounds.47–49 These values are in good agreement with the crystal structure solutions.
Crystal structures
Three title compounds crystallize in different space groups (Table S1†), but they all exhibit the same topology (Fig. 1 and 2). They shared structure consists of a 2D [(C4H11N2)ZnX3]∞ neutral layer, which is comprised of 0D (C4H11N2)ZnX3 units connected by N–H⋯X hydrogen bonds. Within this structure, the 0D (C4H11N2)ZnX3 units are formed by the ZnNX3 tetrahedra and piperazine cations via the bridging of Zn–N bonds.
 |
| Fig. 1 The structures of 0D (H11C4N2)ZnX3 units, in which X = Cl (a), Br (b) and I (c–d). | |
 |
| Fig. 2 Views of the structures of 2D [(C4H11N2)ZnCl3]∞ neutral layer along a direction (a), 3D (C4H11N2)ZnCl3 along c direction (b), 2D [(C4H11N2)ZnI3]∞ neutral layer along a direction, and 3D (C4H11N2)ZnI3 along b direction (d). The spheres of pink, violet-blue, blue, black, green, and dark purple respectively represent the atoms of H, Zn, N, C, Cl, and I. | |
(C4H11N2)ZnCl3 and (C4H11N2)ZnBr3 are isomorphic compounds of each other, crystallizing in P1 space group, and their unit cells have and only one asymmetric unit, containing one Zn, three Cl or Br atoms and one (C4H11N2)+ organic cation (Fig. 1). Whereas (C4H11N2)ZnI3 crystallizes in Cc space group, and its unit cell has four asymmetric unit, which is composed of a pair of (C4H11N2)ZnI3 molecule with different orientation (Fig. 1). Within their crystal structures, the six-membered (C4H11N2)+ ring shows a chair-shaped configuration with each C and N atom exhibiting sp3 hybridization, suggesting that (C4H11N2)+ is non-π-conjugated. Each Zn2+ cation bonds with three halide anions and one N atom from organic cation, forming a tetrahedral ZnNX3 groups (Fig. 1). The lengths of Zn–N are comparable with each other, with the values of 2.055(6) Å, 2.076(9) Å, and 2.074(11)–2.083(11) Å within ZnNCl3, ZnNBr3, and ZnNI3 units. The size effect of halide anions induces the difference of Zn–X bonds, with the magnitude of Zn–I (2.568(2)–2.6067(19) Å) > Zn–Br (2.360(2)–2.4068(18) Å) > Zn–Cl (2.2291(19)–2.2735(19) Å). Obviously, Zn–N distances are smaller than those of Zn–X. Each ZnNX3 tetrahedron is linked with (C4H11N2)+ organic cation via the bridging of Zn–N coordination bond, forming a 0D (C4H11N2)ZnX3 neutral unit (Fig. 1).
The assembly mode of the 0D building block is consistent across the crystal structures of the three compounds (Fig. 2). In each compound, the 0D (C4H11N2)ZnX3 units are connected to each other through hydrogen bonds (Fig. 2a and b and Table S5†). Specifically, two halogen anions from one unit are linked to another pair of 0D (C4H11N2)ZnX3 units, while two H atoms from different N–H bonds bridge another pair of units. The distances of d(D–A) in the hydrogen bonds are 3.224(7)–3.282(7) Å, 3.398(12)–3.464(10) Å, and 3.645(15)–3.740(11) Å for N–H⋯Cl, N–H⋯Br, and N–H⋯I, respectively. Therefore, each 0D (C4H11N2)ZnX3 unit forms connections with four surrounding 0D units through four differently oriented hydrogen bonds (Fig. 2a and b). These interactions lead to the formation of two-dimensional [(C4H11N2)ZnX3]∞ infinite layers (Fig. 2a and b). Finally, the stacking of these 2D layers results in the overall 3D framework of (C4H11N2)ZnX3, in which the stacking form are –A–A–A– for Cl and Br compounds whereas –A–A′–A–A′– for (C4H11N2)ZnI3 (Fig. 2c and d).
Structure comparation
We delve into the structural disparities and their underlying causes between Cl, Br, and I compounds (Fig. 3a–c). Initially, the computations reveal that the distortion index of Δdtetr for ZnNCl3, ZnNBr3, and ZnNI3 are 0.034, 0.050, and 0.078, respectively, based on the formular of
. The distortion degree of ZnNI3 significantly exceeds that of ZnNCl3 and ZnNBr3, aligning with the relative disparity in Zn–X and Zn–N bond lengths in their respective units. As the ionic radius gradually increases from Cl− to I−, it induces a progressive flexibility of the halide anion orbital. Consequently, with minimal alteration in the Zn–N radius, the elongation of Zn–X bond lengths gradually amplifies the tetrahedral distortion degree.
 |
| Fig. 3 The comparisons of radius of halide anions (a), as well as volume (b) and distortion index (c) of ZnNX3 tetrahedra. The arrangement of ZnNX3 tetrahedra in (C4H11N2)ZnCl(Br)3 (d) and (C4H11N2)ZnI3 (e). The green and blue arrow means the orientation of polarization. The symmetry of the space group of P1 (f) and Cc (g). | |
Subsequently, the ZnNCl3 and ZnNBr3 tetrahedra exhibit ideal alignment, with each tetrahedron's distortion direction being consistent, whereas neighboring ZnNI3 not (Fig. 3). This results in the polarizability of ZnNCl3 and ZnNBr3 tetrahedra being able to align along a single polarization direction, while the polarizability of ZnNI3 is partially counteracted. This phenomenon stems from the distinct space groups to which they belong (Fig. 3f and g). Cl and Br compounds crystallize in the P1 space group with the lowest symmetry and the fewest symmetry operations, with each unit cell containing only one ZnNX3 tetrahedron, ensuring consistent orientations of adjacent ZnNX3 groups. Conversely, (C4H11N2)ZnI3 crystallizes in the monoclinic Cc space group, leading to the orientations of adjacent (C4H11N2)ZnI3 molecules crossing each other, with the consistent components aligning along the polar axis. This observation is likely due to the size effect of the halide anions. The significantly larger ionic radius of I− causes the volume of the ZnNI3 tetrahedron to increase, rendering it unfeasible for the (x, y, z) symmetry operations in P1 space group to accommodate the (H11C4N2)ZnI3 molecules in space. Consequently, the latter can only be arranged in the Cc space group with more symmetry operations and energy-favorable.
Properties
TG-DTA analysis indicates that the initial decomposition temperatures of the Cl, Br, and I compounds are 300 °C, 334 °C, and 338 °C, respectively (Fig. S5†). These temperatures suggest substantial thermal stability, surpassing that of reported semi-organic NLO crystals or NLO organic–inorganic metal halides (OIMHs), such as C(NH2)3SO3F (162 °C),72 [C(NH2)3]6(PO4)2·3H2O (100 °C), (C5H6ON)(H2PO4) (166 °C),73 and (H7C3N6)(H6C3N6)HgCl3 (225 °C).48 Additionally, these three crystals exhibit air stability, showing no signs of weathering or moisture absorption even after being stored in air at room temperature for several weeks (Fig. S2†).
The ultraviolet absorption spectrum shows that the transmittance of (C4H11N2)ZnCl3 at 190 nm is greater than 80% (Fig. 4a). And the wavelength-depended F(R) (absorption coefficient/scattering coefficient) values are nearly close to zero (Fig. 4a). These are consistent with most reported deep-UV crystals, such as (NH4)2Na3Li9(SO4)7,36 HfF2(SO4)40 and Ba(NH2SO3)2,45 indicating that (C4H11N2)ZnCl3 is able to transmit in the deep-UV region and possesses a wide bandgap larger than 6.20 eV. The absorption cutoff edges for (C4H11N2)ZnBr3 and (C4H11N2)ZnI3 are 207 nm and 222 nm, respectively, corresponding to bandgaps of 5.53 eV and 4.52 eV (Fig. 4b and c). Notably, the halogen substitution from I to Cl increases the compound's bandgap and enhances its UV transmittance due to the difference in electronegativity of the halide anions. Importantly, the bandgaps of (C4H11N2)ZnCl3 and (C4H11N2)ZnBr3 are wider than all reported NLO-active OIMHs, including (C3N6H7)(C3N6H6)HgCl3 (4.40 eV),48 (C4H10NO)2Cd2Cl6 (5.45 eV),74 (L/D-C10H20N2O4)Cd5Cl12 (5.42 eV),75 (C20H20P)CuBr2 (3.62 eV)52 and L/D-C6H10N3O2ZnBr3 (5.02 eV).76 This highlights the exceptional deep-UV/UV transmittance and optical properties of (C4H11N2)ZnCl3 and (C4H11N2)ZnBr3. Their wide bandgaps also ensure good resistance to laser stability, with powder laser damage thresholds being 205.4 MW cm−2, 150.2 MW cm−2, and 87.0 MW cm−2 for Cl, Br, and I compounds, respectively, which are 51, 38 and 22 times that of AgGaS2 (AGS: 4.0 MW cm−2).
 |
| Fig. 4 UV transmittance spectra (a, b, c for Cl, Br, and I compounds, respectively). The inset shows the bandgaps and as-grown crystals. F(R) is the absorption coefficient/scattering coefficient. Oscilloscope traces of the SHG signals (150–210 μm) with 1064 nm laser radiation (d) and Phase-matching curves (e). KDP was used as references for the SHG measurements. | |
The SHG responses of (C4H11N2)ZnX3 were measured under 1064 nm laser radiation, and the results are presented in Fig. 4d. All compounds exhibit SHG activity, with efficiencies of 0.8, 2.5, and 2.1 times that of KH2PO4 (KDP) for (C4H11N2)ZnCl3, (C4H11N2)ZnBr3, and (C4H11N2)ZnI3, respectively, within the particle size range of 150–210 μm. Besides, Fig. 4e demonstrates that all compounds are phase-matchable. Combining the mentioned UV spectra, as shown in Fig. 6, we can draw some exciting conclusions. Firstly, (C4H11N2)ZnCl3 is the first reported deep-UV OIMHs with SHG activity (Fig. 5c), and its SHG response is comparable with those of known deep-UV compounds, including (NH4)2Na3Li9(SO4)7 (0.5 × KDP),36 Ba3P3O10Cl (0.6 × KDP)37 and HfBa3M2F14Cl (0.9 × KDP).78 Secondly, the SHG effect of (C4H11N2)ZnBr3 surpasses that of currently reported OIMHs with a bandgap greater than 5.0 eV (Fig. 6a), including L/D-C12H20N6O4Cd2Cl5 (0.2 × KDP),76 (C4H10NO)2Cd2Cl6 (0.73 × KDP),74 and (L/D-C10H20N2O4)Cd5Cl12 (0.25 and 0.3 × KDP).76 Additionally, the SHG effect of (C4H11N2)ZnBr3 is comparable to some recently reported semi-organic short UV NLO crystals, such as (C5H6ON)(H2PO4) (3 × KDP),77 [C(NH2)3]3PO4·2H2O (1.5 × KDP)73 and Ba(SO3NH2)2 (2.7 × KDP).46 Finally and most importantly, the substitution of I with Br in (H11C4N2)ZnI3 (2.1 × KDP, 4.52 eV) not only significantly increases the bandgap but also enhances its SHG effect. The anomalous relationship between bandgap and SHG effect due to halogen substitution is unprecedented, distinguishing it from previously reported OIMHs, such as (C6H5(CH2)4NH3)4BiBr7·H2O (0.4 × KDP, 3.52 eV) with (C6H5(CH2)4NH3)4BiI7·H2O (1.3 × AGS, 2.29 eV),50 and Cs3Pb2(CH3COO)2Br5 (4 × KDP, 3.26 eV) with Cs3Pb2(CH3COO)2I5 (8 × KDP, 2.55 eV).55
 |
| Fig. 5 SHG response and bandgap of selected NLO-active OIMHs with d10 cation. | |
 |
| Fig. 6 The comparisons of dipole moment (a and b) and SHG contributions (c) of ZnNX3 tetrahedra as well as bandgaps (d) and SHG effects (e) in (C4H11N2)ZnX3 (X = Cl, Br, I). | |
Structure–property relationship
We further analyzed the structure–property relationship of the three title compounds, especially the intrinsic reason why the Br compound can achieve SHG-bandgap balance optimization. Firstly, the size difference of halogens results in the volumes of ZnNBr3 and ZnNCl3 being much smaller than that of ZnNI3, leading to different arrangement modes of this distorted tetrahedron in the structure (Fig. 3). Secondly, although the difference in halogen polarizability leads to a much larger distortion degree of ZnNI3 than those of ZnNBr3 and ZnNCl3 (Fig. 3c), the gap between ZnNI3 and ZnNBr3 narrows in terms of dipole moment (Fig. 6a). The dipole moments of ZnNCl3, ZnNBr3 and ZnNI3 are 2.065, 5.113, and 6.692–7.306 D, respectively (Table S7†). Therefore, after further considering the arrangement of structural units and the unit cell volume, we found that the order of dipole moment per unit volume is (C4H11N2)ZnBr3 (0.020 DÅ−3) > (C4H11N2)ZnI3 (0.018 D Å−3) ≫ (C4H11N2)ZnCl3 (0.009 D Å−3) (Fig. 6b). This order is consistent with the experimentally measured SHG effects, (C4H11N2)ZnCl3 (0.8 × KDP) ≪ (C4H11N2)ZnI3 (2.1 × KDP) < (C4H11N2)ZnBr3 (2.5 × KDP) (Fig. 6e). Therefore, considering that the SHG effect and bandgap are contradictory conditions for NLO crystals, we arrive at the following conclusions: ① chloride anion has the strongest electronegativity and the weakest polarizability. And, due to the perfectly aligned arrangement of ZnNCl3 along the polarization direction in the structure, (C4H11N2)ZnCl3 can guarantee a moderate SHG effect (0.8 × KDP) while maintaining deep-UV transparency with wide bandgap (>6.2 eV); ② iodide anion has the weakest electronegativity and the strongest polarizability. Therefore, although the polarity of ZnNI3 is partially cancelled out due to its non-ideal arrangement, its smaller bandgap ensures a strong SHG effect (2.1 × KDP; 4.52 eV); ③ the slightly larger radius of bromide anion compared to chloride allows ZnNBr3 to adopt the same arrangement mode as ZnNCl3, with perfectly superimposed polarity. Therefore, compared with (C4H11N2)ZnCl3, the SHG effect of (C4H11N2)ZnBr3 is dramatically enhanced at the expense of a partial bandgap sacrifice (Fig. 6d and e). More importantly, compared with (C4H11N2)ZnI3, (C4H11N2)ZnBr3 exhibits an enhanced SHG effect while significantly widening the band gap, thus achieving optimization of the SHG-bandgap balance (Fig. 6d and e).
Systematic theoretical calculations were conducted to analyze the band structures and densities of states (DOS) of the title compounds. The calculated band structures for these compounds are illustrated in Fig. S6.† The theoretical bandgaps (4.815 eV, 4.404 eV, and 4.304 eV for Cl, Br, and I compounds, respectively) roughly align with experimental values. For (C4H11N2)ZnCl3 and (C4H11N2)ZnBr3, the theoretical bandgaps deviate significantly from experimental values due to the discontinuity of the exchange–correlation function. As shown in Fig. 7a–c, The Cl-3p, Br-4p, and I-5p states predominantly contribute to the top of the valence band (VB), while the Zn-4s state is the primary contributor to the bottom of the conduction band (CB). Additionally, with the substitution of Cl by Br and I, the contribution of X-ns2np6 orbitals to the bottom of the CB becomes significant and shows an increasing trend.
 |
| Fig. 7 The DOS (a–c) as well as the SHG density distribution in VB (d–f) and CB (g–i) for (C4H11N2)ZnX3 (X = Cl, Br, I). | |
We further calculated the optical properties (birefringence and second-order nonlinear optical coefficients) of the title compounds. A principal axis transformation was first conducted; for the triclinic systems of (C4H11N2)ZnCl3 and (C4H11N2)ZnBr3, the rotation angles between the crystallographic axes and the principal dielectric axes are shown in Table S8.† For the monoclinic system of (C4H11N2)ZnI3, the rotation angle between the crystallographic axes and the principal dielectric axes is 25.173°. As depicted in Fig. S7,† the birefringence values for the Cl, Br, and I compounds are 0.04@546 nm, 0.06@546 nm, and 0.08@546 nm, relating to the shortest type-I phase-matchable wavelength of 371 nm, 335 nm, and 451 nm, respectively. It is worth noting that from (C4H11N2)ZnI3 to (C4H11N2)ZnBr3, halide substitution has achieved a triple increase in SHG effect, bandgap, and phase-matching capability. We also calculated the effective second-order nonlinear optical coefficients (deff), which are 0.522 pm V−1, 0.847 pm V−1, and 1.344 pm V−1 for Cl, Br, and I compounds, respectively. Notably, the deff of the (C4H11N2)ZnBr3 is 2.2 times that of KDP (0.39 pm V−1), aligning closely with the experimental value (2.5 × KDP).
The origins of the SHG effects for the three title compounds were elucidated through computational analysis of the SHG density distribution (Fig. 6d–i). In the VB, the Cl-3p, Br-4p, and I-5p orbitals predominantly contribute to the SHG effect in their respective compounds. For the Cl and Br analogs, the Zn-3d and N-2p orbitals associated with the Zn-N bonds also play a significant role in the SHG effect, but a contribution that is negligible in the I analog. Furthermore, within the CB, the Zn-3s orbital is identified as the primary contributor to the SHG, with additional contributions from the Cl-3p/Br-4p/I-5p and N-2p orbitals on the Zn–N bonds. The quantitative contributions from the piperazine cation and the distorted tetrahedral ZnNX3 anion to the SHG were assessed, with the ZnNCl3, ZnNBr3, and ZnNI3 units exhibiting SHG contribution percentages of 87.8%, 95.4%, and 96.5%, respectively (Fig. 6e). This trend indicates a progressive increase in SHG contribution as the halogen substitution proceeds from chlorine to bromine to iodine, correlating with a decrease in electronegativity and an increase in polarizability of the halogen anions. Concurrently, the impact of the Zn–X bonds on the SHG effect of the compounds becomes more pronounced.
Conclusions
This study culminates in the discovery of three novel noncentrosymmetric piperazine zinc halide compounds, (C4H11N2)ZnX3 (X = Cl, Br, I), achieved through strategic halide substitution. Notably, (C4H11N2)ZnCl3 stands as the pioneering deep-UV nonlinear optical OIMH with a SHG effect of 0.8 × KDP. Our findings underscore the exceptional case of (C4H11N2)ZnBr3, which achieves a remarkable bandgap expansion to 5.53 eV without compromising on a potent SHG response (2.5 × KDP), setting a new benchmark among OIMHs with bandgaps greater than 5.0 eV. The progression from chlorine to bromine and iodine in halide substitution is shown to progressively enhance SHG contributions, correlating with reduced electronegativity and heightened polarizability of the halide ions. These findings not only broaden the scope of OIMHs with tailored optical properties but also highlight (C4H11N2)ZnBr3 as a prime candidate for short-wavelength UV NLO crystal applications. The work underscores the importance of halide choice in engineering OIMHs with optimized SHG-bandgap balance and sets the stage for further explorations into high-performance NLO materials.
Author contributions
Chen Jin and Wu Huai-Yu: conceptualization, methodology, writing – original draft, data curation, visualization; Xu Miao-Bin, Wang Ming-Chang, Chen Qian-Qian and Bing-Xuan Li: data curation; Hu Chun-Li: formal analysis; Hu Chun-Li, Du Ke-Zhao and Chen Jin: writing – review & editing, supervision.
Data availability
Crystallographic data for (H11C4N2)ZnX3 (X = Cl, Br, I) have been deposited with the CCDC under deposition numbers 2313121–2313123† and can be obtained from https://www.ccdc.cam.ac.uk/. Other data are available from the authors upon request.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
Our work has been supported by the National Natural Science Foundation of China (No. 22205037 and 22373014) and Natural Science Foundation of Fujian Province (2023J01498).
References
- C. Chen, Y. Wang, B. Wu, K. Wu, W. Zeng and L. Yu, Design and synthesis of an ultraviolet-transparent nonlinear optical crystal Sr2Be2B2O7, Nature, 1995, 373, 322–324 CrossRef CAS.
- M. Mutailipu, J. Han, Z. Li, F. Li, J. Li, F. Zhang, X. Long, Z. Yang and S. Pan, Achieving the full-wavelength phase-matching for efficient nonlinear optical frequency conversion in C(NH2)3BF4, Nat. Photonics, 2023, 17, 694–701 CrossRef CAS.
- X. Liu, Y.-C. Yang, M.-Y. Li, L. Chen and L.-M. Wu, Anisotropic structure building unit involving diverse chemical bonds: a new opportunity for high-performance second-order NLO materials, Chem. Soc. Rev., 2023, 52, 8699–8720 RSC.
- M. Mutailipu, K. R. Poeppelmeier and S.-L. Pan, Borates: A Rich Source for Optical Materials, Chem. Rev., 2021, 121, 1130–1202 CrossRef CAS PubMed.
- J. Chen, C.-L. Hu, F. Kong and J.-G. Mao, High-Performance Second-Harmonic-Generation (SHG) Materials: New Developments and New Strategies, Acc. Chem. Res., 2021, 54, 2775–2783 CrossRef CAS PubMed.
- J. Xu, X. Li, J. Xiong, C. Yuan, S. Semin, T. Rasing and X.-H. Bu, Halide Perovskites for Nonlinear Optics, Adv. Mater., 2020, 32, 1806736 CrossRef CAS PubMed.
- Y. Kang, C. Yang, J. Gou, Y. Zhu, Q. Zhu, W. Xu and Q. Wu, From Cd(SCN)2(CH4N2S)2 to Cd(SCN)2(C4H6N2)2: Controlling Sulfur Content in Thiocyanate Systems Significantly Improves the Overall Performance of UV Nonlinear Optical Materials, Angew. Chem., Int. Ed., 2024, 63, e202402086 CrossRef CAS PubMed.
- X. Zhang, L. Kang, P. Gong, Z. Lin and Y. Wu, Nonlinear Optical Oxythiophosphate Approaching the Good Balance with Wide Ultraviolet Transparency, Strong Second Harmonic Effect, and Large Birefringence, Angew. Chem., Int. Ed., 2021, 60, 6386–6390 CrossRef CAS.
- H. 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.
- Y. Li, J. Luo and S. Zhao, Local Polarity-Induced Assembly of Second-Order Nonlinear Optical Materials, Acc. Chem. Res., 2022, 55, 3460–3469 CrossRef CAS.
- M. Mutailipu, Z. Yang and S. Pan, Toward the Enhancement of Critical Performance for Deep-Ultraviolet Frequency-Doubling Crystals Utilizing Covalent Tetrahedra, Acc. Mater. Res., 2021, 2, 282–291 CrossRef CAS.
- L. Kang, F. Liang, X. Jiang, Z. Lin and C. Chen, First-Principles Design and Simulations Promote the Development of Nonlinear Optical Crystals, Acc. Chem. Res., 2020, 53, 209–217 CrossRef CAS.
- M. Mutailipu, M. Zhang, Z. Yang and S. Pan, Targeting the Next Generation of Deep-Ultraviolet Nonlinear Optical Materials: Expanding from Borates to Borate Fluorides to Fluorooxoborates, Acc. Chem. Res., 2019, 52, 791–801 CrossRef CAS.
- Z. Xia and K. R. Poeppelmeier, Chemistry-Inspired Adaptable Framework Structures, Acc. Chem. Res., 2017, 50, 1222–1230 CrossRef CAS.
- K. M. Ok, Toward the Rational Design of Novel Noncentrosymmetric Materials: Factors Influencing the Framework Structures, Acc. Chem. Res., 2016, 49, 2774–2785 CrossRef CAS.
- B. Zhang and Z. Chen, Recent advances of inorganic phosphates with UV/DUV cutoff edge and large second harmonic response, Chin. J. Struct. Chem., 2023, 42, 100033 CrossRef CAS.
- Z. Bai, J. Lee, C.-L. Hu, G. Zou and K. M. Ok, Hydrogen bonding bolstered head-to-tail ligation of functional chromophores in a 0D SbF3·glycine adduct for a short-wave ultraviolet nonlinear optical material, Chem. Sci., 2024, 15, 6572–6576 RSC.
- Y.-N. Zhang, Q.-F. Li, B.-B. Chen, Y.-Z. Lan, J.-W. Cheng and G.-Y. Yang, Na3B6O10(HCOO): an ultraviolet nonlinear optical sodium borate-formate, Inorg. Chem. Front., 2022, 9, 5032–5038 RSC.
- W.-F. Chen, J.-Y. Lu, J.-J. Li, Y.-Z. Lan, J.-W. Cheng and G.-Y. Yang, Sr2[B5O8(OH)]2·[B(OH)3]·H2O: A Strontium Borate That Shows Deep-Ultraviolet-Transparent Nonlinear Optical Properties, Chem. – Eur. J., 2024, 30, e202400739 CrossRef CAS PubMed.
- H. Su, Z. Yan, X. Hou and M. Zhang, Fluorooxoborates: A precious treasure of deep-ultraviolet nonlinear optical materials, Chin. J. Struct. Chem., 2023, 42, 100027 CrossRef CAS.
- D. Gao, H. Wu, Z. Hu, J. Wang, Y. Wu and H. Yu, Recent advances in F-containing iodate nonlinear optical materials, Chin. J. Struct. Chem., 2023, 42, 100014 CrossRef.
- Q. Shi, L. Dong and Y. Wang, Evaluating refractive index and birefringence of nonlinear optical crystals: Classical methods and new developments, Chin. J. Struct. Chem., 2023, 42, 100017 CrossRef CAS.
- J.-H. Wu, B. Zhang, T.-K. Jiang, F. Kong and J.-G. Mao, From Cs8Sb4Nb5O5F35 to Cs6Sb4Mo3O5F26: The first noncentrosymmetric fluoroantimonite with d10 transition metal, Chin. J. Struct. Chem., 2023, 42, 100016 CrossRef CAS.
- J.-H. Wu, C.-L. Hu, T.-K. Jiang, J.-G. Mao and F. Kong, Highly Birefringent d0 Transition Metal Fluoroantimonite in the Mid Infrared Band: Order–Disorder Regulation by Cationic Size, J. Am. Chem. Soc., 2023, 145, 24416–24424 CrossRef CAS PubMed.
- J. Chen, C.-L. Hu, F.-F. Mao, J.-H. Feng and J.-G. Mao, A Facile Route to Nonlinear Optical Materials: Three-Site Aliovalent Substitution Involving One Cation and Two Anions, Angew. Chem., Int. Ed., 2019, 58, 2098–2102 CrossRef CAS.
- M. Mutailipu, F. Li, C. Jin, Z. Yang, K. R. Poeppelmeier and S. Pan, Strong Nonlinearity Induced by Coaxial Alignment of Polar Chain and Dense [BO3] Units in CaZn2(BO3)2, Angew. Chem., Int. Ed., 2022, 61, e202202096 CrossRef CAS.
- H. Qiu, F. Li, Z. Li, Z. Yang, S. Pan and M. Mutailipu, Breaking the Inherent Interarrangement of [B3O6] Clusters for Nonlinear Optics with Orbital Hybridization Enhancement, J. Am. Chem. Soc., 2023, 145, 24401–24407 CrossRef CAS.
- H. Qiu, F. Li, C. Jin, Z. Yang, J. Li, S. Pan and M. Mutailipu, Fluorination Strategy Towards Symmetry Breaking of Boron-centered Tetrahedron for Poly-fluorinated Optical Crystals, Angew. Chem., Int. Ed., 2023, 63, e202316194 CrossRef PubMed.
- F. Ding, K. J. Griffith, W. Zhang, S. Cui, C. Zhang, Y. Wang, K. Kamp, H. Yu, P. S. Halasyamani, Z. Yang, S. Pan and K. R. Poeppelmeier, NaRb6(B4O5(OH)4)3(BO2) Featuring Noncentrosymmetry, Chirality, and the Linear Anionic Group BO2−, J. Am. Chem. Soc., 2023, 145, 4928–4933 CrossRef CAS.
- H. Pei, X. Wang, J. Zhang, F. Zhang, Z. Yang and S. Pan, Ba2B9O13F4·BF4: first fluorooxoborate with unprecedented infinite [B18O26F8] tubes and deep-ultraviolet cutoff edge, Sci. China: Chem., 2023, 66, 1073–1077 CrossRef CAS.
- 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, 12691–12700 CrossRef CAS PubMed.
- Y. Shang, J. Xu, H. Sha, Z. Wang, C. He, R. Su, X. Yang and X. Long, Nonlinear optical inorganic sulfates: The improvement of the phase matching ability driven by the structural modulation, Coord. Chem. Rev., 2023, 494, 215345 CrossRef CAS.
- X. Chen, Y. Li, J. Luo and S. Zhao, Recent advances in non-π-conjugated nonlinear optical sulfates with deep-UV absorption edge, Chin. J. Struct. Chem., 2023, 42, 100044 CrossRef CAS.
- M. Mutailipu and S. Pan, Emergent Deep-Ultraviolet Nonlinear Optical Candidates, Angew. Chem., Int. Ed., 2019, 59, 20302–20317 CrossRef.
- J. Dang, D. Mei, Y. Wu and Z. Lin, A comprehensive survey on nonlinear optical phosphates: Role of multicoordinate groups, Coord. Chem. Rev., 2021, 431, 213692 CAS.
- 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, 3833–3837 CrossRef CAS PubMed.
- 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.
- S. Zhao, X. Yang, Y. Yang, X. Kuang, F. Lu, P. Shan, Z. Sun, Z. Lin, M. Hong and J. Luo, Non-Centrosymmetric RbNaMgP2O7 with Unprecedented Thermo-Induced Enhancement of Second Harmonic Generation, J. Am. Chem. Soc., 2018, 140, 1592–1595 CrossRef CAS PubMed.
- Y. Li, C. Yin, X. Yang, X. Kuang, J. Chen, L. He, Q. Ding, S. Zhao, M. Hong and J. Luo, A Nonlinear Optical Switchable Sulfate of Ultrawide Bandgap, CCS Chem., 2020, 3, 2298–2306 CrossRef.
- C. Wu, C. Jiang, G. Wei, X. Jiang, Z. Wang, Z. Lin, Z. Huang, M. G. Humphrey and C. Zhang, Toward Large Second-Harmonic Generation and Deep-UV Transparency in Strongly Electropositive Transition Metal Sulfates, J. Am. Chem. Soc., 2023, 145, 3040–3046 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. Wang, X. Leng, Y. Kuk, J. Lee, Q. Jing and K. M. Ok, Deep-Ultraviolet Transparent Mixed Metal Sulfamates with Enhanced Nonlinear Optical Properties and Birefringence, Angew. Chem., Int. Ed., 2023, 63, e202315434 CrossRef.
- Z. Bai and K. M. Ok, Designing Sulfate Crystals with Strong Optical Anisotropy through π-Conjugated Tailoring, Angew. Chem., Int. Ed., 2023, 63, e202315311 CrossRef.
- H. 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.
- H. Tian, C. Lin, X. Zhao, F. Xu, C. Wang, N. Ye and M. Luo, Ba(SO3CH3)2: a Deep-Ultraviolet Transparent Crystal with Excellent Optical Nonlinearity Based on a New Polar Non-π-conjugated NLO Building Unit SO3CH3−, CCS Chem., 2022, 5, 2497–2505 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, 7621–7625 CrossRef CAS PubMed.
- H.-Y. Wu, C.-L. Hu, M.-B. Xu, Q.-Q. Chen, N. Ma, X.-Y. Huang, K.-Z. Du and J. Chen, From H12C4N2CdI4 to H11C4N2CdI3: a highly polarizable CdNI3 tetrahedron induced a sharp enhancement of second harmonic generation response and birefringence, Chem. Sci., 2023, 14, 9533–9542 RSC.
- Z. Bai, J. Lee, H. Kim, C.-L. Hu and K. M. Ok, Unveiling the Superior Optical Properties of Novel Melamine-Based Nonlinear Optical Material with Strong Second-Harmonic Generation and Giant Optical Anisotropy, Small, 2023, 19, 2301756 CrossRef CAS.
- L. Liu, Z. Bai, L. Hu, D. Wei, Z. Lin and L. Zhang, A melamine-based organic–inorganic hybrid material revealing excellent optical performance and moderate thermal stability, J. Mater. Chem. C, 2021, 9, 7452–7457 RSC.
- D. Chen, S. Hao, L. Fan, Y. Guo, J. Yao, C. Wolverton, M. G. Kanatzidis, J. Zhao and Q. Liu, Broad Photoluminescence and Second-Harmonic Generation in the Noncentrosymmetric Organic–Inorganic Hybrid Halide (C6H5(CH2)4NH3)4MX7·H2O (M = Bi, In, X = Br or I), Chem. Mater., 2021, 33, 8106–8111 CrossRef CAS.
- Y. Kang and Q. Wu, A review of the relationship between the structure and nonlinear optical properties of organic-inorganic hybrid materials, Coord. Chem. Rev., 2024, 498, 215458 CrossRef CAS.
- J. Wu, Y. Guo, J.-L. Qi, W.-D. Yao, S.-X. Yu, W. Liu and S.-P. Guo, Multi-Stimuli Responsive Luminescence and Domino Phase Transition of Hybrid Copper Halides with Nonlinear Optical Switching Behavior, Angew. Chem., Int. Ed., 2023, 62, e202301937 CrossRef CAS.
- X.-Y. Li, Q. Wei, C.-L. Hu, J. Pan, B.-X. Li, Z.-Z. Xue, X.-Y. Li, J.-H. Li, J.-G. Mao and G.-M. Wang, Achieving Large Second Harmonic Generation Effects via Optimal Planar Alignment of Triangular Units, Adv. Funct. Mater., 2023, 33, 2210718 CrossRef CAS.
- Y. Liu, Y.-P. Gong, S. Geng, M.-L. Feng, D. Manidaki, Z. Deng, C. C. Stoumpos, P. Canepa, Z. Xiao, W.-X. Zhang and L. Mao, Hybrid Germanium Bromide Perovskites with Tunable Second Harmonic Generation, Angew. Chem., Int. Ed., 2022, 61, e202208875 CrossRef CAS.
- Q.-R. Shui, H.-X. Tang, R.-B. Fu, Y.-B. Fang, Z.-J. Ma and X.-T. Wu, Cs3Pb2(CH3COO)2X5 (X=I, Br): Halides with Strong Second-Harmonic Generation Response Induced by Acetate Groups, Angew. Chem., Int. Ed., 2021, 60, 2116–2119 CrossRef CAS.
- P. Qian, Y. Li, J. Cheng, J. Li, H. Zeng, L. Huang, G. Zou and Z. Lin, Multiple Functions of l-Thioproline in the Synthesis of Chiral Metal Bromides Showing Second-Harmonic-Generation Responses, Inorg. Chem., 2024, 63, 8013–8017 CrossRef CAS PubMed.
- S.-F. Yan, Y. Guo, W. Liu, S.-P. Guo and J. Wu, Tellurium(IV) Halide Achieving Effective Nonlinear-Optical Activity: The Role of Chiral Ligands and Lattice Distortion, Inorg. Chem., 2024, 63, 73–77 CrossRef CAS PubMed.
- J.-L. Qi, J. Wu, Y. Guo, Z.-P. Xu, W. Liu and S.-P. Guo, Quasi-linear CuX2 (X = Cl, Br) motif-built hybrid copper halides realizing encouraging nonlinear optical activities, Inorg. Chem. Front., 2023, 10, 3319–3325 RSC.
- W. He, Y. Yang, C. Li, W. P. D. Wong, F. Cimpoesu, A. M. Toader, Z. Wu, X. Wu, Z. Lin, Q.-h. Xu, K. Leng, A. Stroppa and K. P. Loh, Near-90° Switch in the Polar Axis of Dion–Jacobson Perovskites by Halide Substitution, J. Am. Chem. Soc., 2023, 145, 14044–14051 CrossRef CAS.
- R. H. Blessing, An empirical correction for absorption anisotropy, Acta Crystallogr., Sect. A: Found. Crystallogr., 1995, 51, 33–38 CrossRef PubMed.
- G. M. Sheldrick, Crystal structure refinement with it SHELXL, Acta Crystallogr., 2015, 71, 3–8 CrossRef.
- A. L. Spek, Single-crystal structure validation with the program it PLATON, J. Appl. Crystallogr., 2003, 36, 7–13 CrossRef CAS.
- P. K. a. F. Munk, An Article on Optics of Paint Layers, Z. Tech. Phys., 1931, 259–274 Search PubMed.
- S. K. Kurtz and T. T. Perry, A Powder Technique for the Evaluation of Nonlinear Optical Materials, J. Appl. Phys., 2003, 39, 3798–3813 CrossRef.
- 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, J. Phys.: Condens. Matter, 2002, 14, 2717 CrossRef CAS.
- V. Milman, B. Winkler, J. A. White, C. J. Pickard, M. C. Payne, E. V. Akhmatskaya and R. H. Nobes, Electronic structure, properties, and phase stability of inorganic crystals: A pseudopotential plane-wave study, J. Quantum Chem., 2000, 77, 895–910 CrossRef CAS.
- J. P. Perdew, K. Burke and M. Ernzerhof, Generalized Gradient Approximation Made Simple, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS.
- J. S. Lin, A. Qteish, M. C. Payne and V. Heine, Optimized and transferable nonlocal separable ab initio pseudopotentials, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 4174 CrossRef PubMed.
- J. Lin, M.-H. Lee, Z.-P. Liu, C. Chen and C. J. Pickard, Mechanism for linear and nonlinear optical effects in β−BaB2O4 crystals, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 60, 13380 CrossRef CAS.
- S. N. Rashkeev, W. R. Lambrecht and B. Segall, Efficient ab initio method for the calculation of frequency-dependent second-order optical response in semiconductors, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 57, 3905 CrossRef CAS.
- Y.-P. Lin, S. Hu, J. Xu, Z. Zhang, X. Qi, X. Lu, J. Jin, X.-Y. Huang, Q. Xu, Z. Deng, Z. Xiao and K.-Z. Du, Bright green emitter of Mn-doped C4H12N2ZnX4 (X = Cl, Br) for X-ray radiography and WLEDs, Chem. Eng. J., 2023, 468, 143818 CrossRef CAS.
- 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, 15978–15981 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, 2008–2016 CrossRef CAS.
- D. Sun, D. Wang, Y. Dang, S. Zhang, H. Chen, R. Hou, K. Wu and C. Shen, Organic–Inorganic Hybrid Noncentrosymmetric (Morpholinium)2Cd2Cl6 Single Crystals: Synthesis, Nonlinear Optical Properties, and Stability, Inorg. Chem., 2022, 61, 8076–8082 CrossRef CAS.
- J. Cheng, Y. Deng, X. Dong, J. Li, L. Huang, H. Zeng, G. Zou and Z. Lin, Homochiral Hybrid Organic–Inorganic Cadmium Chlorides Directed by Enantiopure Amino Acids, Inorg. Chem., 2022, 61, 11032–11035 CrossRef CAS.
- W. Seo and K. M. Ok, Novel noncentrosymmetric polar coordination compounds derived from chiral histidine ligands, Inorg. Chem. Front., 2021, 8, 4536–4543 RSC.
- J. Lu, X. Liu, M. Zhao, X.-B. Deng, K.-X. Shi, Q.-R. Wu, L. Chen and L.-M. Wu, Discovery of NLO Semiorganic (C5H6ON)+(H2PO4)−: Dipole Moment Modulation and
Superior Synergy in Solar-Blind UV Region, J. Am. Chem. Soc., 2021, 143, 3647–3654 CrossRef CAS.
- Y. Mei, C.-L. Hu, R.-L. Tang, W.-D. Yao, W. Liu and S.-P. Guo, KBa3M2F14Cl (M = Zr, Hf): novel short-wavelength mixed metal halides with the largest second-harmonic generation responses contributed by mixed functional moieties, Chem. Sci., 2024, 15, 8500–8505 RSC.
|
This journal is © the Partner Organisations 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.