The exploration of new infrared nonlinear optical crystals based on the polymorphism of BaGa4S7

Zhen Qian , Haonan Liu , Yujie Zhang , Hongping Wu , Zhanggui Hu , Jiyang Wang , Yicheng Wu and Hongwei Yu *
Tianjin Key Laboratory of Functional Crystal Materials, Institute of Functional Crystal, Tianjin University of Technology, Tianjin 300384, China. E-mail: hwyu15@gmail.com

Received 14th June 2022 , Accepted 26th July 2022

First published on 26th July 2022


Abstract

Balancing the key performance metrics, such as a large second harmonic generation (SHG) response and a wide band gap, is an extremely important but intractable challenge for the development of infrared (IR) nonlinear optical (NLO) materials. Herein, we report two new polymorphs of BaGa4S7, β- and γ-BaGa4S7, both of which crystallize in the noncentrosymmetric and polar space groups, Pmc21 (No. 26) for β-BaGa4S7 and Cm (No. 8) for γ-BaGa4S7. They show an interesting structural evolution from AA′ [Ga6S16] chains to AA′A′A [Ga6S16] chains to AAA′AAA′ [Ga6S16] chains as their structures show a transition from the high-temperature phase to the low-temperature phase. More importantly, they both exhibit excellent NLO properties. In particular, β-BaGa4S7 exhibits the best balance along with a large phase-matching SHG response (1.2 × AgGaS2) and a wide band gap (3.94 eV). These indicate that β-BaGa4S7 will be a promising IR NLO crystal for high-power laser application. The research also provides an insight into exploring new IR NLO crystals based on structural rearrangement.


Introduction

Nonlinear optical (NLO) crystals, which can expand the spectral regions of solid-state lasers for generating coherent radiation at a variety of difficult to access wavelengths through a series of frequency conversion technologies, such as second harmonic generation (SHG), optical parametric oscillation (OPO), optical parametric amplification (OPA), etc., play crucial roles in the development of solid-state lasers.1–5 In particular, ultraviolet (UV) and infrared (IR) coherent light is difficult or impossible to generate through direct lasing. In contrast, relying on the frequency conversion of NLO crystals is the best method for these laser outputs.6,7 Currently, although a series of borate and phosphate crystals such as β-BaB2O4 (BBO),8 LiB3O5 (LBO),9 CsLiB6O10 (CLBO),10 KH2PO4 (KDP)11 and KTiOPO4 (KTP)12 have been able to achieve a high-power laser output in the UV and visible regions, crystals that can be used in the IR region are still very limited. They are mainly chalcopyrite-type AgGaS2 (AGS),13 AgGaSe2 (AGSe)14 and ZnGeP2 (ZGP),15 and some inherent performance defects, e.g. a low laser damage threshold (LDT) and two-photon absorption (TPA), also severely prohibit their performance in high-power laser output applications.16 Therefore, exploring new IR NLO crystals is still of current research interest.

Generally, an excellent IR NLO crystal needs to satisfy the following conditions, including (i) wide IR transmission windows, covering the atmospheric windows of 3–5 μm and 8–12 μm, (ii) large SHG coefficients (>0.5 × AGS), (iii) wide band-gaps and high LDTs (Eg >3.0 eV or >10 × AGS).17 Due to the merits of chalcogenides in IR optical transparency and SHG responses, they have been widely considered as the most promising materials class for the exploration of IR NLO crystals.18 In order to increase the band-gaps of chalcogenides, various strategies have been suggested, among which the most effective one is to introduce the electron-positive alkali or alkali-earth cations into structures.19 On the one hand, alkali or alkaline-earth metals have no d–d or f–f electronic transitions, so they can effectively widen the band gaps and reduce the TPAs, and thus enhance the LDTs of materials. On the other hand, the flexible coordination of alkali or alkali-earth cations is also helpful for adjusting the orientation of microscopic dipole moments of the chalcogenide MQ4 (M = Zn, Cd, Ga, In, Si, and Ge; Q = S, Se) tetrahedra to make them have the additive manner and enhance the SHG responses.20,21 Under the guidance of these ideas, lots of high-performance IR chalcogenide NLO crystals have been synthesized, including BaGa4Q7 (Q = S, Se),22,23 BaGa2GeQ6 (Q = S, Se),24 LiGaQ2 (Q = S, Se),25,26 Na2BaMQ4 (M = Ge, Sn; Q = S, Se),27 Li2Ga2GeQ6 (Q = S, Se),28etc.

Beyond these, polymorphism is also an effective strategy for exploring new IR NLO crystals because the rearrangement of the basic building units in the structure can further optimize the functional properties of materials.29 The typical examples consist of BaB2O4 (α: R-3c, β: R3),8 BaTeMo2O9 (α: Pca21, β: P21, γ: P21),30 Li2ZnGeS4 (α: Pna21, β: Pn),31,32 K2Hg3Ge2S8 (α: Aba2; β: C2),33 and Ba2[VO2F2(IO3)2]IO3 (α: Pna21, β: P21).34 In our recent research, the highly symmetric NCS β-BaGa4Se7 and β-BaGa2Se4 have been synthesized and they exhibit more excellent NLO properties than the previously reported α-phases.35,36 These indicate that structural rearrangement is indeed an ideal strategy for rationally synthesizing high-performance NLO crystals. Inspired by these, our current attention has been focused on another important IR NLO crystal, BaGa4S7, which can exhibit a larger band-gap (3.54 eV, from the crystal sample) and LDT (1.2 J cm−2 at 1.064 μm and a 15 ns pulse width)23 than β-BaGa4Se7 and β-BaGa2Se4. Our research has resulted in the discovery of two new polymorphs of BaGa4S7, which crystallize in orthorhombic (Pmc21) and monoclinic (Cm) space groups, respectively. Based on the orders they were discovered, these two new polymorphs of BaGa4S7 are named β-BaGa4S7 (Pmc21) and γ-BaGa4S7 (Cm), respectively, and the previously reported phase is named α-BaGa4S7 (Pmn21).23 As far as we know, α-, β- and γ-BaGa4S7 represent the first ternary chalcogenides with three polar and non-centrosymmetric (NCS) polymorphs. The further structural analyses show that the phase-transitions among α-, β- and γ-BaGa4S7 are caused by the rearrangement of [Ga6S16] chains in their structures, which also represents the new phase-transition mechanism. More importantly, β-BaGa4S7 exhibits excellent NLO properties, including large phase-matching SHG responses (∼1.2 × AGS) and wide band gaps (Eg = 3.94 eV), indicating that they are promising IR NLO crystals. Herein, we will report their syntheses, crystal structures, phase transitions and NLO properties.

Experimental section

Materials

Barium (Ba, 99%, Aladdin), barium sulfide (BaS, 99.9%, Beijing Hawk Science and Technology Co. Ltd), gallium (Ga, 99.99%, Aladdin), gallium sulfide (99.9%, Beijing Hawk Science and Technology Co. Ltd), sulfide (S, 99.99%, Aladdin), and germanium (Ge, 99.99%, Beijing Hawk Science and Technology Co. Ltd) all were commercially purchased without further refinement.

Syntheses

α-, β- and γ-BaGa4S7 are polymorphous. Their pure polycrystalline samples were synthesized through a series of solid-state reactions in sealed silica tubes. The stoichiometric Ba (0.48 mmol, 0.0663 g), BaS (0.32 mmol, 0.0385 g), Ga (3.22 mmol 0.2246 g) and S (6.07 mmol, 0.1704 g) were firstly loaded into graphite crucibles. Then the graphite crucibles were put into silica tubes and flame-sealed under 10−3 Pa. These tubes were heated to 720 °C in 30 h and kept at this temperature for about 100 h. Then they were cooled to room temperature and the mixture in the tube was thoroughly ground and sealed into new silica tubes again. The new silica tubes were further heated to a higher temperature, 720 °C for 30 h and the temperature was maintained for 100 h. By repeating the above process with a 20 °C higher calcination temperature than the last reaction, the phase-transition temperature from β- to α-BaGa4S7 has been determined. And the pure polycrystalline samples of α- and β-BaGa4S7 were also obtained. Furthermore, their millimeter-sized crystals were also obtained by melting the pure polycrystalline sample (for α-BaGa4S7) or with the BaS and Ge as the flux for β- and γ-BaGa4S7 at 1100, 1000 and 950 °C, respectively.

Structural refinement and crystal data

The single-crystal structure data of α-BaGa4S7 were reported by Eisenmann et al. in 1983.37 For new β- and γ-BaGa4S7, their single-crystal data were collected at 293(2) K by using a Bruker D8 VENTURE CCD diffractometer with a Mo Kα source (λ = 0.71073 Å).38 All data were integrated with SAINT and a numerical absorption correction using SADABS was applied.39 The structures were solved by direct methods using SHELXT and refined by full-matrix least-squares methods against F2 by SHELXL-2019/1.40 Crystallographic data for the structures were deposited with the Cambridge Crystallographic Data Centre. CCDC 2132789 and 2132787 contain the supplementary crystallographic data. Crystal data, data collection and structure refinement are summarized in Table 1. The final refined atomic coordinates, equivalent isotropic displacement parameters and the calculated bond valence sums (BVSs) are listed in Table S1,41 and the selected bond distances are given in Table S2.
Table 1 Crystal data and structure refinement for α-, β- and γ-BaGa4S7
a R 1 = S||Fo| − |Fc||/S|Fo| and wR2 = [Sw(Fo2Fc2)2/Sw Fo4]1/2 for Fo2 > 2s(Fo2).
Empirical formula α-BaGa4S7[thin space (1/6-em)]23 β-BaGa4S7 γ-BaGa4S7
Formula weight 640.64
Space group (number) Pmn21 (31) Pmc21 (26) Cm (8)
a [Å] 14.755(5) 14.6922(9) 18.347(3)
b [Å] 6.228(2) 6.2291(4) 14.719(3)
c [Å] 5.929(2) 11.9041(7) 6.2295(9)
β [°] 103.921(5)
Volume [Å3] 544.8(3) 1089.45(12) 1632.9(4)
Z 2 4 6
ρ calc [g cm−3] 3.905 3.906 3.909
μ [mm−1] 14.601 14.603 14.615
F (000) 584 1168 1752
Crystal colour Yellow Colourless Yellow
Completeness 99.9% 99.9% 99.9%
Data/Parameters 1159/59 3211/117 2705/174
Goodness-of-fit on F2 0.947 1.033 1.007
Final R indexes [Fo2 > 2s(Fo2)]a R 1 = 0.0201, wR2 = 0.039 R 1 = 0.0392, wR2 = 0.0579 R 1 = 0.0435, wR2 = 0.0785
Flack X parameter 0.032(13) 0.06(2) 0.07(3)


Powder X-ray diffraction

The powder X-ray diffraction (PXRD) analysis was performed on a SmartLab 9KW X-ray diffractometer at room temperature (Cu-Kα radiation). The collected data are within the 2θ range of 10–70°, with a step size of 0.01° and a step time of 2 s. As seen in Fig. S1, the polycrystalline XRD patterns are in good agreement with the calculated ones.

Energy-dispersive spectroscopy

Microprobe elemental analyses and the elemental distribution maps were measured on a field-emission scanning electron microscope (Quanta FEG 250) made by FEI. The energy-dispersive spectroscopy measurement corroborated the existence of Ba/Ga/S with an average atomic ratio of 21.39%[thin space (1/6-em)]:[thin space (1/6-em)]43.56%[thin space (1/6-em)]:[thin space (1/6-em)]35.05% and 21.39%[thin space (1/6-em)]:[thin space (1/6-em)]43.46%[thin space (1/6-em)]:[thin space (1/6-em)]35.15% for β- and γ-BaGa4S7, respectively, which are approximately equal to the theoretical one, 21.43%, 43.53%, and 35.03% (Fig. S2).

UV-Vis-near-IR (NIR) diffuse-reflectance and IR spectra

The UV/Vis/NIR diffuse reflectance spectrum was recorded with a Shimadzu SolidSpec-3700DUV UV/Vis/NIR Spectrophotometer at room temperature. The measurement range was from 300 to 2000 nm and barium sulfate was used as the diffuse reflection standard. The IR spectra were recorded on a Nicolet iS50 Fourier transform infrared spectrometer in the range 500–4000 cm−1.

Raman spectroscopy

Raman spectra of the crushed crystals of the title compounds were obtained using a LABRAM HR Evolution spectrometer equipped with a CCD detector using 532 nm radiation from a diode laser. For each sample, crystals were placed on a glass slide and a 50× objective lens was used to choose the area of the crystal specimens to be measured. The grating was set to be 600 g mm−1. The maximum power of 60 mW and a beam diameter of 35 μm were used. The spectrum was collected using an integration time of 5 s.

Power SHG measurement

The SHG responses were measured according to the method proposed by Kurtz and Perry with laser radiation at an optical wavelength of 2090 nm.42 Powder samples of polymorphs of BaGa4S7were graded into several distinct particle size ranges of 25–53, 53–75, 75–106, 106–120, 120–150, 150–180, and 180–212 μm. At the same time, a powder AGS sample with the same particle size range was used as a reference.

The details of computational methods

The electronic band structures, the partial density of states and optical properties for α-, β- and γ-BaGa4S7 were carried out using the CASTEP package based on density functional theory (DFT).43 Generalized gradient approximation (GGA) parametrized by the Perdew–Burke–Ernzerhof (PBE) functional was chosen for the exchange–correlation energy, and the pseudopotential was set as norm-conserving pseudopotential (NCP).44 The valence electrons were set as: Ba 4d105p66s2, Ga 4s24p1, and S 3s23p4. The plane-wave energy cutoff value was set at 860.0 eV. The corresponding Monkhorst–Pack κ-point meshes were adopted 2 × 5 × 6, 2 × 5 × 3 and 2 × 2 × 6 for α-, β- and γ-BaGa4S7, respectively.45

The SHG coefficients were calculated from the band wave functions using the so-called length-gauge formalism derived by Aversa and Sipe at a zero frequency limit. The static second-order nonlinear susceptibilities χ(2)αβγ can be reduced as:46–48

χ(2)αβγ = χ(2)αβγ (VE) + χ(2)αβγ (VH)

Virtual-Hole (VH), Virtual-Electron (VE) and Two-Band (TB) processes play an important role in the total SHG coefficient χ(2). The TB process can be neglected owing to the little contribution for SHG. The formulas for calculating χ(2)αβγ (VE) and χ(2)αβγ (VH) are as follows:

image file: d2qi01263d-t1.tif

image file: d2qi01263d-t2.tif

Here, α, β, γ are Cartesian components, v and v′ denote valence bands, c and c′ refer to conduction bands, and P(αβγ) denotes the full permutation. The band energy difference and momentum matrix elements are denoted as ℏωij and Pαij, respectively. As we know, the virtual electron (VE) progress of the occupied and unoccupied states is the main contribution to the overall SHG effect.49

On account of the electron transition from the valence band (VB) to the conduction band (CB), we inferred the imaginary part of the dielectric function. Then, by adopting the Kramers–Kronig transform, we calculated the real part of the dielectric function. The refractive indices n (and the birefringence Δn) were achieved from the real part of the dielectric function.

Results and discussion

Crystal structures and phase transitions

The structural data of α-, β- and γ-BaGa4S7 are listed in Table 1. It is clear that the unit cells gradually increase from α-, β- to γ-BaGa4S7. And γ-BaGa4S7 crystallizes in the lowest space group (Cm) and has the largest density (3.910 g cm−3). Therefore, γ-BaGa4S7 should be the low-temperature phase according to the principle of the thermal expansion and contraction, while α-BaGa4S7 is the high-temperature phase.
α-BaGa4S7. The structure of α-BaGa4S7 has been reported previously.23 But in order to better compare it with other polymorphs, its structure is still shown in Fig. 1. Its asymmetric unit contains one unique Ba, two unique Ga, and four unique S atom(s). In the structure, the Ga atoms are coordinated by the four S atoms to form [GaS4] tetrahedra with the Ga–S bond distances ranging from 2.242(6) to 2.335(6) Å. These [GaS4] tetrahedra further connect with each other by corner-sharing to form one-dimensional (1D) [Ga6S16] chains (Fig. 1a). The adjacent [Ga6S16] chains adopt the opposite and staggered stacking along the c-axis, and they are connected to form the 3D [Ga4S7] framework. So, if the 1D [Ga6S16] chains with the different orientation are marked as A and A′ (Fig. 1b), respectively, the structure of α-BaGa4S7 could be written as AA′AA′AA′… Thus, AA′ can be as the minimum stacking length of [Ga6S16] chains in α-BaGa4S7. And owing to the different orientation of A- and A′-[Ga6S16] chains, the microscopic polarization in the ab plane is offset. Only the polarization along the stacking direction (c-axis) will be retained. For Ba2+ cations, they are all filled in the stacking space of [Ga6S16] chains and have the sole coordinated environment, [BaS12] polyhedral with the Ba–S distances ranging from 3.417(9) to 3.672(10) Å (Fig. S3a). They are consistent with other reported ones.19 The bond valence calculations show the bond valence sums (BVSs) for Ba2+, Ga3+ and S2− are 1.81, 3.04–3.06 and 1.90–2.20, respectively.41 They are all in agreement with their ideal oxidation states.
image file: d2qi01263d-f1.tif
Fig. 1 Crystal structure of α-BaGa4S7. (a) The total crystal structure framework along the b direction, it is formed by the adjacent [Ga6S16] chains adopting the opposite and arrangement in the AA′AA′…… manner. (b) The 1D [Ga6S16] chains with different orientations as the A and A′, respectively.
β-BaGa4S7. The structure of β-BaGa4S7 is shown in Fig. 2. Its lattice parameter along the c-axis is around twice of α-BaGa4S7. In its asymmetric unit, there are two unique Ba, four unique Ga, and eight unique S atom(s). All the Ga atoms in β-BaGa4S7 are also coordinated by four S atoms to form the GaS4 tetrahedra, and these GaS4 tetrahedra are also connected to form the 1D [Ga6S16] chains (Fig. 2a). Different from α-BaGa4S7, however, the adjacent [Ga6S16] chains are arranged into the framework of β-BaGa4S7 in a manner of AA′A′AAA′A′A…(Fig. 2b). So, the minimum stacking length of [Ga6S16] chains in β-BaGa4S7 is AA′A′A, which is twice that in α-BaGa4S7. Accordingly, the Ba2+ cations in β-BaGa4S7 have two different coordination environments, [BaS12] and [BaS11] polyhedra with the Ba–S distances ranging from 3.456(3) to 3.642(3) Å (Fig. S3b). The BVSs for Ba2+, Ga3+ and S2− can be calculated as 1.78–1.80, 3.03–3.08 and 1.91–2.18, respectively. They are consistent with their respective ideal oxidation states.
image file: d2qi01263d-f2.tif
Fig. 2 Crystal structure of β-BaGa4S7. (a) The total crystal structure framework along the b direction, it is formed by the adjacent [Ga6S16] chains that adopt the opposite and arrangement in the AA′A′A… manner. (b) The 1D [Ga6S16] chains with different orientations as the A and A′, respectively.
γ-BaGa4S7. The structure of γ-BaGa4S7 is shown in Fig. 3. γ-BaGa4S7 has the largest unit cell among three polymorphs. It is around three times that of α-BaGa4S7. Accordingly, in the asymmetric unit of γ-BaGa4S7, it contains three unique Ba, six unique Ga, and twelve unique S atom(s). In its structure (Fig. 3a), all the Ga atoms are also coordinated to form the GaS4 tetrahedra, and these GaS4 tetrahedra are also connected to form the 1D [Ga6S16] chains (Fig. 3b). Notably, when these [Ga6S16] chains are connected into the 3D framework, they adopt a more complex stacking manner, i.e. AAA′AAA′… (Fig. 3a). So, a larger unit cell is necessary for γ-BaGa4S7. And owing to the complex stacking of [Ga6S16] chains, three different polyhedra are also built for three different coordinated Ba2+ cations in γ-BaGa4S7. Ba(1) and Ba(3) are coordinated into [BaS12] polyhedra with Ba–S distances ranging from 3.451(8) to 3.641(6) Å, and Ba(2) are coordinated into [BaS11] polyhedra with the Ba–S distance in the range of 3.423(6) to 3.649(6) Å (Fig. S3c). Clearly, these bond distances for all Ga–S bonds and Ba–S bonds are consistent with other reported ones. The BVSs for Ba2+, Ga3+ and S2− are 1.74–1.80, 2.99–3.04 and 1.90–2.09, respectively. They are also consistent with their respective ideal oxidation states.
image file: d2qi01263d-f3.tif
Fig. 3 Crystal structure of γ-BaGa4S7. (a) The total crystal structure frameworks along the b direction, it is formed by the adjacent [Ga6S16] chains that adopt the opposite and arrangement in the AAA′AAA′…… manner. (b) The 1D [Ga6S16] chains with different orientations as the A and A′, respectively.
The phase transitions. It is clear that the different crystal structures of α-, β- and γ-BaGa4S7 mainly originate from the different orientations and stacking manners of [Ga6S16] chains in their structures. In order to understand the relationship between the phase transitions and temperatures, a series of solid-state reactions with different calcined temperatures were performed in sealed silica tubes. As shown in Fig. 4a, when the calcination temperature is in the range of 720 to 800 °C, the main reaction product is β-BaGa4S7. When the temperature is higher than 800 °C, β-BaGa4S7 starts to transfer into α-BaGa4S7 (Fig. 4a). From 800 °C to the melting temperature (∼1100 °C), the stable phase is always α-BaGa4S7. That also indicates α-BaGa4S7 should be the high-temperature phase. The phase transition temperature from β- to α-BaGa4S7 is around 800 °C. However remarkably, calcining α-BaGa4S7 at 750 °C for 72 h (Fig. 4b), no phase-transition from α- to β-BaGa4S7 was observed, indicating that the phase-transition from β- to α-BaGa4S7 is irreversible. In addition, during the whole heating process, the main phase for γ-BaGa4S7 was not observed. That indicates γ-BaGa4S7 may be a substable phase, which can be attributed to the complex stacking of [Ga6S16] chains in the structure. In order to confirm the stabilities of α-, β- and γ-BaGa4S7. Their Bond Strain Index (BSI) and Global Instability Index (GII) were also calculated (Table S3a–c). Generally, BSI and GII values imply the strain caused by electronic-induced and lattice-induced strains, respectively.50,51 A value greater than 0.05 v.u. indicates the structure is strained, and a value greater than 0.2 v.u. indicates the structure is unable. Clearly, γ-BaGa4S7 has the largest BSI and GII values, which are 0.161 valence unit (v.u.) and 0.133 v.u. (Table S3c), respectively. These indicate that γ-BaGa4S7 is strongly strained. When γ-BaGa4S7 transforms into α- and β-BaGa4S7, the electronic-induced and lattice-induced strains can be released with the decrease in the BSI and GII values (Tables S3a and 3b).
image file: d2qi01263d-f4.tif
Fig. 4 The powder XRD patterns of BaGa4S7. (a) The variable-temperature powder XRD patterns of the β-BaGa4S7. (b) The powder XRD patterns of the α-BaGa4S7 after 750 °C for 72 h.

The linear and NLO properties of α- and β-BaGa4S7

As described above, α-, β- and γ-BaGa4S7 have the same compositions and stoichiometric ratios. The different arrangements of their basic building units (BBUs) result in their different structures. Clearly, the different structures will make them exhibit different functional properties. In order to better show the difference of their properties, their linear and NLO properties have been measured and compared in the same conditions. Remarkably, as the pure phase for γ-BaGa4S7 was not obtained, only the properties of α- and β-BaGa4S7 will be characterized.
UV-Vis-NIR spectra. The UV-Vis-NIR spectra of α- and β-BaGa4S7 in the range from 300–2500 nm are shown in Fig. 5a. Their UV cut-off edges are 330 and 315 nm, respectively. Correspondingly, their band-gaps are 3.75 and 3.94 eV (from powder sample). Therefore, it is clear that the band gap of β-BaGa4S7 is a little larger than that of α-BaGa4S7. The larger band gap of β-BaGa4S7 can also be confirmed by the crystal colors (Fig. 5a). Clearly, β-BaGa4S7 is colorless, while the color of α-BaGa4S7 is pale yellow (Fig. 5a).
image file: d2qi01263d-f5.tif
Fig. 5 Optical properties of BaGa4S7. (a) The UV/Vis-NIR diffuse reflectance spectrum of β-BaGa4S7. Inset: α- and β-BaGa4S7 crystal. (b) The FTIR spectrum. Inset: Raman spectrum of α- and β-BaGa4S7. (c) Particle size dependence of SHG intensities of β-BaGa4S7 and AGS. (d) The SHG signals for α- and β-BaGa4S7 in the same particle sizes.
IR and Raman spectra. The IR spectra of α- and β-BaGa4S7 are shown in Fig. 5b. It shows α- and β-BaGa4S7 have similar IR spectra. Both of them have no obvious absorption in a wide range from 4000 to 500 cm−1 (i.e. 2.5–20 μm), implying their wide IR transmission region. Furthermore, the Raman spectra of α- and β-BaGa4S7 were also recorded (Fig. 5b). The strong absorptions at 241, 322, 339 and 387 cm−1 for α-BaGa4S7 and 249, 345, 391 and 406 cm−1 for β-BaGa4S7 are due to asymmetric and symmetric stretching vibrations of Ga–S–Ga modes, in GaS4 tetrahedral units. For the two compounds, the other Raman peaks below 200 cm−1 are due to the Ba–S vibrations. These are consistent with those of other related chalcogenides, such as BaGa2S4 and PbGa2S4.52,53
Second harmonic generation properties. The powder SHG response of β-BaGa4S7 was studied through the Kurtz and Perry method with 2.09 μm radiation.43 The data reveal that β-BaGa4S7 is type I phase-matchable (Fig. 5c), and the SHG response is around 1.2 × AGS and 1.3 × α-BaGa4S7 under the same conditions (Fig. 5d). Obviously, the SHG response can satisfy the requirement of IR NLO crystals (0.5 × AGS).

The structure–property relationships

The first principles calculations. To understand the origin of the NLO properties, we firstly carried out first-principles calculations for α- and β-BaGa4S7 with the same parameters, and the results show that α- and β-BaGa4S7 have similar electronic structures (Fig. 6). α- and β-BaGa4S7 have the following direct band gaps, 2.38 eV for α-BaGa4S7 and 2.48 eV for β-BaGa4S7 (Fig. 6). Clearly, their band gaps are both smaller than the experimental ones owing to the discontinuity of exchange–correlation energy.54 The partial density of states analysis of transitions from occupied states to unoccupied states reveals the microcosmic origin of the optical properties from an electronic level insight. The 4s, 4p states of Ga, and 3s states of S construct the top region of valence states from −8 to 0 eV. Their wide hybridization in this area indicates that Ga–S bonds make the main contribution of the upper side of the valence band. The bottom region of conduction states mainly is composed of 4s, 4p states of Ga, 3s state of S, and 5d state of Ba. Thus, [GaS4] tetrahedra have a significant effect on its optical properties, and Ga–S polyhedra determine the SHG effects of α- and β-BaGa4S7. In addition, the calculations show that the three independent nonzero SHG coefficients d31, d32 and d33 in α-BaGa4S7 are −9.43, −8.05 and 26.83 pm V−1; the three independent nonzero SHG coefficients d31, d32 and d33 in β-BaGa4S7 are −11.24, −9.59 and −27.11 pm V−1. Both of them are also consistent with the experimental results.
image file: d2qi01263d-f6.tif
Fig. 6 The theoretical calculation for BaGa4S7. Band structures and density of states of α- and β-BaGa4S7 calculated using the PBE functional under the same conditions, the (a) and (b) for α- and β-BaGa4S7, respectively.
The dipole moment calculations based on the bond-valence modes. From the experimental and calculated results, it can be seen that β-BaGa4S7 exhibits more excellent NLO properties than the α-phase. In order to understand the structural origin resulting in their different functional properties, a more detailed structure comparison of the two polymorphs has been carried out based on the dipole moment calculations (Table 2).41 Clearly, α- and β-BaGa4S7 crystallize in the orthorhombic space group and possess a non-zero dipole moment along the c axis (Fig. 7).
image file: d2qi01263d-f7.tif
Fig. 7 The microscopic dipole moments. The orientation of the microscopic dipole moments of the GaS4 groups in α- and β-BaGa4S7 of (a) and (b). The total dipole moments along the −c axis (pink arrows).
Table 2 The dipole moments and flexibility indices (F) of [GaS4]5− groups in α- and β-BaGa4S7
Compound Polyhedron Dipole moment F
X-axis Y-axis Z-axis Magnitude Total
α-BaGa4S7 [Ga(1)S4]5− 0 0 −11.04 11.04 0.059 (esu cm Å−3) 0.216
[Ga(2)S4]5− 0 0 −5.16 5.16
Unit cell 0 0 −16.20 16.20
β-BaGa4S7 [Ga(1)S4]5− 0 0 −9.45 9.45 0.043 (esu cm Å−3) 0.218
[Ga(2)S4]5− 0 0 1.42 1.42
[Ga(3)S4]5− 0 0 −5.42 5.42
[Ga(4)S4]5− 0 0 −10.14 10.14
Unit cell 0 0 −23.59 23.59


For α-BaGa4S7, it contains two crystallographically different GaS4 tetrahedra, i.e. Ga(1)S4 and Ga(2)S4. The sum of their dipole moments in a unit cell is 16.20 D along the c-axis (Fig. 7a), and the total dipole moment is 0.059 esu cm Å−3. For β-BaGa4S7, four different GaS4 tetrahedra can be observed in a unit cell, and their sum for the dipole moment is 23.59 D along the c-axis (Fig. 7b), the total dipole moment is 0.043 esu cm Å−3. It is clear that α-BaGa4S7 can exhibit stronger polarity in the unit cell than that of β-BaGa4S7. Generally, for the polar materials, a larger net dipole moment in the unit cell often results in a larger SHG response.55 Therefore, the SHG response of α-BaGa4S7 should larger than β-BaGa4S7, but clearly this is contrary to the results of experimental and theoretical calculations. The cause of the heightened SHG response in β-BaGa4S7 is difficult to explain just based on the anionic groups, as shown by an examination of the dipole moments along the c direction (Table 2). Accordingly, different mechanisms must be investigated in order to further investigate the causes behind the β-BaGa4S7 enhanced SHG response. It is well known that the SHG responses of materials are related to not only the intrinsic dipole moments but also the induced dipole moments, which have been considered to have a more significant effect. The flexibility index was proposed to characterize the induced polarizability by valence electrons.56 In 2014, Jiang et al. proposed a simple “flexible dipole model” to ‘quantify’ the induced dipole moments by calculating an empirical “flexibility index” F.57 Clearly, the flexibility index calculations (Table 2) reveal that GaSe4 tetrahedra in β-BaGa4S7 (F = 0.218) are more “flexible” than the α-BaGa4S7 (F = 0.216), indicating that the former groups are easier to generate the larger induced polarization compared with the latter, combined with the higher number of the GaSe4 tetrahedra in the unit cell (α-BaGa4S7: 8 vs. β-BaGa4S7: 16), that makes the SHG response enhanced from α- to β-BaGa4S7.

The symmetry-adapted mode decomposition (SAMD) calculations. From the calculations above, β-BaGa4S7 exhibits a larger F-value than the α-phase. This indicates that the intrinsic dipole moments play important roles in the SHG enhancement of β-BaGa4S7. Beyond these, in order to further elucidate the contribution of atomic displancements from inversion center on NLO properties, we also performed a symmetry-adapted mode decomposition using the symmetry mode analysis tool of the Bilbao Crystallographic Server.58–60

First, a hypothetical high symmetry structure was identified for β-BaGa4S7 in centrosymmetric (CS) Pnma (No. 62) (Fig. S4). Next, the pseudosymmetric structures subjected to structural relaxation by performing DFT calculation with lattice constants were fixed at the values in the corresponding experimental structures and the atomic forces were restricted to be <8 meV Å−1. We then examined the relationship between the CS (Pnma) and NCS (Pmc21) phases in terms of orthonormal symmetry-adapted modes (Tables S4 and S5). In these two structures, the Ba(1), Ga(1), Ga(2), S(1), S(2), S(3) and S(4) in the nonpolar Pnma structure are split into Ba(1)/Ba(2), Ga(1)/Ga(4), Ga(2)/Ga(3), S(1)/S(7), S(2)/S(5), S(3)/S(8) and S(4)/S(6), respectively, in the polar Pmc21 structure, with a single polar mode with a maximum atomic displacement of 3.05 Å (Table S6). Table 3 lists the atoms that make up each Wyckoff position (WP) as well as the distortions of the atoms; Fig. S5 shows the distortions. The atomic displacements at WPs 8d contribute the most to the overall distortions of 12.34 Å, indicating that Ga and S displacements contribute the most to the local dipole. As a result, the GaS4 groups play a vital role in inversion symmetry breaking and contribute significantly to the SHG intensity in β-BaGa4S7. Furthermore, we calculated the net mode-distortion vectors in the polar axis, Ap (Table 3). Values of Ap/A of 4c and 8d less than 1 indicate that there are no “fully polarized” WPs, but through the decomposition of linearity and orthogonality along the polar axis, the atomic displacements in those orbitals contribute to a local dipole and the asymmetry in the structure.

Table 3 The α- and β-BaGa4S7 atomic decompositions, total and polar-axis distortion model amplitudes (in Å)
Compound WP Atom(s)a A A p A/Ap Total SAMDs
a The atom shown in bold for each WP makes the greatest displacement.
β-BaGa4S7 4c Ba1, S1 2.83 1.94 0.69 12.34 11.3 × 1021 (Å cm−3)
8d Ga1, Ga2, S2, S3, S4 12.01 7.21 0.60
α-BaGa4S7 2a Ba1 0.20 0.20 1 1.54 2.84 × 1021 (Å cm−3)
2b S4 0.44 0.34 0.77
4f Ga1, Ga2, S1, S2, S3 1.47 1.47 1


For comparison, we applied the same analysis to α-BaGa4S7 (see Tables S4–S6 and Fig. S3, S4) and found the total distortions to be just 1.54 Å for α-BaGa4S7 (Table 3). It is clear that the atomic displacements in both compounds are contributed by the GaS4 groups, and β-BaGa4S7 is much larger than α-BaGa4S7. To obtain the specific centric-mode displacements (SAMDs) and provide an unbiased measure of the amounts of polar distortion per unit volume in the two crystals, the values of SAMDs are also calculated, i.e., 2.84 × 1021 and 11.3 × 1021 Å cm−3 for α- and β-BaGa4S7 (Table 1), respectively. The values of SAMDs of β-BaGa4S7 are larger than α-BaGa4S7, which leads to a bigger SHG response in general and also correlates better with the experimental multiplicity intensity. These results suggest that the larger SHG responses of β-BaGa4S7 can be attributed to its larger atomic polar displacements and the easier polariability when it is radiated.

Conclusions

In summary, two new polymorphs of BaGa4S7, i.e., β- and γ-BaGa4S7, with excellent performances have been discovered and synthesized by high-temperature solution methods in sealed quartz tubes. Their phase transformations from the low-temperature phase to the high-temperature phase are systematically analyzed based on a series of solid-state reactions and the calculations of structural parameters. Furthermore, their functional properties were also measured and calculated under the same conditions. These show that all of materials are promising IR NLO materials. In particular, β-BaGa4S7 exhibits the best NLO properties among them, including the widest experimental band gap (3.94 eV) and the largest SHG response (1.2 × AGS) with phase-matching behavior. The study of the structure–property relationship reveals that the better NLO properties of β-BaGa4S7 can be attributed to the different arrangement of the GaS4 tetrahedral structure's basic building units, which suggests that structural rearrangement is an effective strategy for optimizing and exploring new IR NLO crystals. Further research on the crystal growth of β-BaGa4S7 is ongoing.

Author contributions

Writing – original draft: Zhen Qian; Writing – review and editing: Haonan Liu, Yujie Zhang and Hongping Wu; Resources: Zhanggui Hu; Supervision: Jiyang Wang and Yicheng Wu; Funding acquisition and conceptualization – ideas: Hongwei Yu.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work is supported by the National Natural Science Foundation of China (Grant No. 52172006, 22071179, 51972230, 51890864 and 51890865) and the Natural Science Foundation of Tianjin (Grant No. 20JCJQJC00060 and 21JCJQJC00090).

References

  1. S. Pei, B. Liu, X. Jiang, Y. Zou, W. Chen, Q. Yan and G. Guo, Superior Infrared Nonlinear Optical Performance Achieved by Synergetic Functional Motif and Vacancy Site Modulations, Chem. Mater., 2021, 33, 8831–8837 CrossRef CAS.
  2. J. Chen, H. Chen, F. Xu, L. Cao, X. Jiang, S. Yang, Y. Sun, X. Zhao, C. Lin and N. Ye, Mg2In3Si2P7: A Quaternary Diamond-like Phosphide Infrared Nonlinear Optical Material Derived from ZnGeP2, J. Am. Chem. Soc., 2021, 143, 10309–10316 CrossRef CAS PubMed.
  3. C. Trovatello, A. Marini, X. Xu, C. Lee and G. Cerullo, Optical parametric amplification by monolayer transition metal dichalcogenides, Nat. Photonics, 2021, 15, 1–5 CrossRef.
  4. X. Dong, Q. Jing, Y. Shi, Z. Yang, S. Pan, K. Poeppelmeier, J. Young and J. Rondinelli, Pb2Ba3(BO3)3Cl: A Material with Large SHG Enhancement Activated by Pb-Chelated BO3 Groups, J. Am. Chem. Soc., 2015, 137, 9417–9422 CrossRef CAS PubMed.
  5. Y. Wang, B. Zhang, Z. Yang and S. Pan, Cation-Tuned Synthesis of Fluorooxoborates: Towards Optimal Deep-Ultraviolet Nonlinear Optical Materials, Angew. Chem., Int. Ed., 2018, 57, 2150–2154 CrossRef CAS PubMed.
  6. G. Shi, Y. Wang, F. Zhang, B. Zhang, Z. Yang, X. Hou, S. Pan and K. Poeppelmeier, Finding the Next Deep-Ultraviolet Nonlinear Optical Material: NH4B4O6F, J. Am. Chem. Soc., 2017, 139, 10645–10648 CrossRef CAS PubMed.
  7. X. Wang, Y. Wang, B. Zhang, F. Zhang, Z. Yang and S. Pan, CsB4O6F: A Congruent-Melting Deep-Ultraviolet Nonlinear Optical Material by Combining Superior Functional Units, Angew. Chem., Int. Ed., 2017, 56, 14119–14123 CrossRef CAS PubMed.
  8. C. Chen, B. Wu, A. Jiang and G. You, A new-type ultraviolet SHG crystal β-BaB2O4, Sci. China, Ser. B, 1985, 28, 235–243 Search PubMed.
  9. 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, 616–621 CrossRef CAS.
  10. G. Ryu, C. Yoon, T. Han and H. Gallagher, Growth and characterisation of CsLiB6O10 (CLBO) crystals, J. Cryst. Growth, 1998, 191, 492–500 CrossRef CAS.
  11. H. Siegfried, Elastische und thermoelastische Eigenschaften von KH2PO4, KH2AsO4, NH4H2PO4, NH4H2AsO4 und RbH2PO4, Z. Kristallogr. – Cryst. Mater., 1964, 120, 401–414 CrossRef.
  12. J. Bierlein and H. Vanherzeele, Potassium titanyl phosphate: properties and new applications, J. Opt. Soc. Am. B, 1989, 6, 622–633 CrossRef CAS.
  13. A. Harasaki and K. Kato, New Data on the Nonlinear Optical Constant, Phase-Matching, and Optical Damage of AgGaS2, Jpn. J. Appl. Phys., 1997, 36, 700–703 CrossRef CAS.
  14. G. Catella, L. Shiozawa, J. Hietanen, R. Eckardt, R. Route, R. Feigelson, D. Cooper and C. Marquardt, Mid-IR absorption in AgGaSe2 optical parametric oscillator crystals, Appl. Opt., 1993, 32, 3948–3951 CrossRef CAS PubMed.
  15. P. Budni, L. Pomeranz, M. Lemons, C. Miller, J. Mosto and E. Chicklis, Efficient mid-infrared laser using 1.9 μm-pumped Ho:YAG and ZnGeP2 optical parametric oscillators, J. Opt. Soc. Am. B, 2000, 17, 723–728 CrossRef CAS.
  16. M. Chen, S. Zhou, W. Wei, X. Wu, H. Lin and Q. Zhu, Phase Matchability Transformation in the Infrared Nonlinear Optical Materials with Diamond-Like Frameworks, Adv. Opt. Mater., 2022, 10, 2102123 CrossRef CAS.
  17. H. Lin, W. Wei, H. Chen, X. Wu and Q. Zhu, Rational design of infrared nonlinear optical chalcogenides by chemical substitution, Coord. Chem. Rev., 2020, 406, 213150 CrossRef CAS.
  18. Q. Yue, W. Wei, H. Chen, X. Wu, H. Lin and Q. Zhu, Salt-inclusion chalcogenides: an emerging class of IR nonlinear optical materials, Dalton Trans., 2020, 49, 14338–14343 RSC.
  19. B. Liu, X. Jiang, H. Zeng and G. Guo, [ABa2Cl][Ga4S8] (A = Rb, Cs): Wide-Spectrum Nonlinear Optical Materials Obtained by Polycation-Substitution-Induced Nonlinear Optical (NLO)-Functional Motif Ordering, J. Am. Chem. Soc., 2020, 142, 10641–10645 CrossRef CAS PubMed.
  20. W. Wang, D. Mei, F. Liang, J. Zhao, Y. Wu and Z. Lin, Inherent laws between tetrahedral arrangement pattern and optical performance in tetrahedron-based mid-infrared nonlinear optical materials, Coord. Chem. Rev., 2020, 421, 213444 CrossRef CAS.
  21. M. Chen, S. Zhou, W. Wei, M. Ran, B. Li, X. Wu, H. Lin and Q. Zhu, RbBiP2S6: A Promising IR Nonlinear Optical Material with a Giant Second-Harmonic Generation Response Designed by Aliovalent Substitution, ACS Mater. Lett., 2022, 4, 1264–1269 CrossRef CAS.
  22. J. Yao, D. Mei, L. Bai, Z. Lin, W. Yin, P. Fu and Y. Wu, BaGa4Se7: A New Congruent-Melting IR Nonlinear Optical Material, Inorg. Chem., 2010, 49, 9212–9216 CrossRef CAS PubMed.
  23. X. Lin, G. Zhang and N. Ye, Growth and Characterization of BaGa4S7: A New Crystal for Mid-IR Nonlinear Optics, Cryst. Growth Des., 2009, 9, 1186–1189 CrossRef CAS.
  24. S. Grechin, P. Nikolaev, A. Ionin, I. Kinyaevskii and Y. Andreev, BaGa2GeS6 and BaGa2GeSe6 crystals for nonlinear optical frequency conversion, Quantum Electron., 2020, 50, 782–787 CrossRef CAS.
  25. V. Petrov, A. Yelisseyev, L. Isaenko, S. Lobanov and J. Zondy, Second harmonic generation and optical parametric amplification in the mid-IR with orthorhombic biaxial crystals LiGaS2 and LiGaSe2, Appl. Phys. B, 2004, 78, 543–546 CrossRef CAS.
  26. L. Isaenko, A. Yelisseyev, S. Lobanov, A. Titov, V. Petrov, J. Zondy, P. Krinitsin, A. Merkulov, V. Vedenyapin and J. Smirnova, Growth and properties of LiGaX2 (X = S, Se, Te) single crystals for nonlinear optical applications in the mid-IR, Cryst. Res. Technol., 2003, 38, 379–387 CrossRef CAS.
  27. K. Wu, Z. Yang and S. Pan, Na2BaMQ4 (M=Ge, Sn; Q=S, Se): Infrared Nonlinear Optical Materials with Excellent Performances and that Undergo Structural Transformations, Angew. Chem., Int. Ed., 2016, 55, 6713–6715 CrossRef CAS PubMed.
  28. Y. Kim, I. Seo, S. Martin, J. Baek, P. Halasyamani, N. Arumugam and H. Steinfink, Characterization of New Infrared Nonlinear Optical Material with High Laser Damage Threshold, Li2Ga2GeS6, Chem. Mater., 2008, 20, 6048–6052 CrossRef CAS.
  29. K. Wu, Y. Yang and L. Gao, A review on phase transition and structure-performance relationship of second-order nonlinear optical polymorphs, Coord. Chem. Rev., 2020, 418, 213380 CrossRef CAS.
  30. J. Zhang, Z. Zhang, W. Zhang, Q. Zheng, Y. Sun, C. Zhang and X. Tao, Polymorphism of BaTeMo2O9: A New Polar Polymorph and the Phase Transformation, Chem. Mater., 2011, 23, 3752–3761 CrossRef CAS.
  31. J. Zhang, D. Clark, J. Brant, K. Rosmus, P. Grima, J. Lekse, J. Jang and J. Aitken, α-Li2ZnGeS4: A Wide-Bandgap Diamond-like Semiconductor with Excellent Balance between Laser-Induced Damage Threshold and Second Harmonic Generation Response, Chem. Mater., 2020, 32, 8947–8955 CrossRef CAS.
  32. Y. Huang, K. Wu, J. Cheng, Y. Chu, Z. Yang and S. Pan, Li2ZnGeS4: a promising diamond-like infrared nonlinear optical material with high laser damage threshold and outstanding second-harmonic generation response, Dalton Trans., 2019, 48, 4484–4488 RSC.
  33. A. Reshak and S. Azam, Linear and nonlinear optical properties of α,-K2Hg3Ge2S8 and α-K2Hg3Sn2S8 compounds, Opt. Mater., 2014, 37, 97–103 CrossRef CAS.
  34. H. Yu, M. Nisbet and K. Poeppelmeier, Assisting the Effective Design of Polar Iodates with Early Transition-Metal Oxide Fluoride Anions, J. Am. Chem. Soc., 2018, 140, 8868–8876 CrossRef CAS PubMed.
  35. Z. Qian, Q. Bian, H. Wu, H. Yu, Z. Lin, Z. Hu, J. Wang and Y. Wu, β-BaGa4Se7: A Promising IR Nonlinear Optical Crystal Designed by Predictable Structural Rearrangement, J. Mater. Chem. C, 2022, 10, 96–101 RSC.
  36. Y. Zhang, Q. Bian, H. Wu, H. Yu, Z. Hu, J. Wang and Y. Wu, Designing A New Infrared Nonlinear Optical Material, β-BaGa2Se4 Inspired by the Phase Transition of the BaB2O4 (BBO) Crystal, Angew. Chem., 2022, 61, e202115374 CAS.
  37. B. Eisenmann, M. Jakowski and H. Schäfer, Data on BaAl4S7 and BaGa4S7, Rev. Chim. Miner., 1983, 20, 329–333 CAS.
  38. Bruker APEX2 bruker analytical X-ray instruments, Inc., Madison, Wisconsin, USA, 2005 Search PubMed.
  39. F. Liebau and X. Wang, Acta Crystallogr., Sect. A: Found. Crystallogr., 2008, 64, 299 CAS.
  40. O. Dolomanov, A. Blake, N. Champness and M. Schroder, OLEX: new software for visualization and analysis of extended crystal structures, J. Appl. Crystallogr., 2003, 36, 1283–1284 CrossRef CAS.
  41. I. Brown and D. Altermatt, Bond-valence parameters obtained from a systematic analysis of the Inorganic Crystal Structure Database, Acta Crystallogr., Sect. B, 2010, 41, 244–247 CrossRef.
  42. S. Kurtz and T. Perry, A Powder Technique for the Evaluation of Nonlinear Optical Materials, J. Appl. Phys., 1968, 39, 3798–3813 CrossRef CAS.
  43. S. Clark, M. Segall, C. Pickard, P. Hasnip, M. Probert, K. Refson and M. Payne, First principles methods using CASTEP, Z. Kristallogr., 2005, 220, 567–570 CAS.
  44. J. Perdew, K. Burke and M. Ernzerhof, Generalized Gradient Approximation Made Simple, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
  45. J. Lin, A. Qteish, M. Payne and V. Heine, Optimized and transferable nonlocal separable ab initio pseudopotentials, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 4174–4180 CrossRef PubMed.
  46. J. Lin, M. Lee, Z. Liu, C. Chen and C. Pickard, Mechanism for linear and nonlinear optical effects in β-BaB2O4 crystals, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 60, 13380–13389 CrossRef CAS.
  47. C. Aversa and J. Sipe, Nonlinear optical susceptibilities of semiconductors: Results with a length-gauge analysis, Phys. Rev. B: Condens. Matter Mater. Phys., 1995, 52, 14636–14645 CrossRef CAS PubMed.
  48. H. Monkhorst and J. Pack, Special points for Brillouin-zone integrations, Phys. Rev. B: Solid State, 1976, 13, 5188–5192 CrossRef.
  49. R. He, Z. Lin, M. Lee and C. Chen, Ab initio studies on the mechanism for linear and nonlinear optical effects in YAl3(BO3)4, J. Appl. Phys., 2011, 109, 103510 CrossRef.
  50. I. Brown, The chemical bond in inorganic chemistry: the bond valence model, 2002 Search PubMed.
  51. M. Marvel, J. Lesage, J. Baek, P. Halasyamani, C. Stern and K. Poeppelmeier, Cation−Anion Interactions and Polar Structures in the Solid State, J. Am. Chem. Soc., 2007, 129, 13963–13969 CrossRef CAS PubMed.
  52. K. Wu, S. Pan, H. Wu and Z. Yang, Synthesis, structures, optical properties and electronic structures of PbGa2Q4 (Q=S, Se) crystals, J. Mol. Struct., 2015, 1082, 174–179 CrossRef CAS.
  53. M. Nazarov, D. Noh and H. Kim, Structural and luminescent properties of calcium, strontium and barium thiogallates, Mater. Chem. Phys., 2008, 107, 456–464 CrossRef CAS.
  54. R. Godby, M. Schlüter and L. Sham, Accurate Exchange-Correlation Potential for Silicon and Its Discontinuity on Addition of an Electron, Phys. Rev. Lett., 1986, 56, 2415–2418 CrossRef CAS PubMed.
  55. H. Yu, H. Wu, S. Pan, Z. Yang, X. Hou, X. Su, Q. Jing, K. Poeppelmeier and J. Rondinelli, Cs3Zn6B9O21: A Chemically Benign Member of the KBBF Family Exhibiting the Largest Second Harmonic Generation Response, J. Am. Chem. Soc., 2014, 136, 1264–1267 CrossRef CAS PubMed.
  56. B. Levine, Erratum: Electrodynamical bond-charge calculation of nonlinear optical susceptibilities, Phys. Rev. Lett., 1969, 22, 787 CrossRef CAS.
  57. X. Jiang, S. Zhao, Z. Lin, J. Luo, P. Bristowe, X. Guan and C. Chen, The role of dipole moment in determining the nonlinear optical behavior of materials: ab initio studies on quaternary molybdenum tellurite crystals, J. Mater. Chem. C, 2014, 2, 530–537 RSC.
  58. C. Capillas, E. Tasci, G. de la Flor, D. Orobengoa, J. Perez-Mato and M. Aroyo, A new computer tool at the Bilbao Crystallographic Server to detect and characterize pseudosymmetry, Z. Kristallogr., 2011, 226, 186–196 CrossRef CAS.
  59. D. Orobengoa, C. Capillas, M. Aroyo and J. Perez-Mato, Amplimodes: symmetry-mode analysis on the Bilbao Crystallographic Server, J. Appl. Crystallogr., 2009, 42, 820–833 CrossRef CAS.
  60. J. Perez-Mato, D. Orobengoa and M. Aroyo, Mode crystallography of distorted structures, Acta Crystallogr., Sect. A: Found. Crystallogr., 2010, 66, 558–590 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available: Atomic coordinates, equivalent isotropic displacement parameters, and the bond valence sums of each atom; selected bond distances and angles; experimental and calculated XRD patterns; elemental analysis; IR transmission; dipole moment calculations. CCDC 2132789 and 2132787. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d2qi01263d

This journal is © the Partner Organisations 2022