Constructing ultraviolet nonlinear optical crystals with large second harmonic generation and short absorption edges by using polar tetrahedral S2O3 groups

Shixian Ke ac, Huixin Fan *a, Chensheng Lin a, Ning Ye d and Min Luo *ab
aKey Laboratory of Optoelectronic Materials Chemistry and Physics, Fujian Institute of Research on the Structure of Matter, Fuzhou, Fujian 350002, China. E-mail: lm8901@fjirsm.ac.cn; huixinfan@fjirsm.ac.cn
bFujian Science & Technology Innovation Laboratory for Optoelectronic Information of China, Fuzhou, Fujian 350002, China
cShanghaiTech University, Shanghai 200120, China
dTianjin Key Laboratory of Functional Crystal Materials, Institute of Functional Crystal, Tianjin University of Technology, Tianjin 350108, China

Received 27th January 2023 , Accepted 31st March 2023

First published on 1st April 2023


Abstract

It is generally considered difficult for traditional sulfates to exhibit strong second harmonic generation (SHG) responses and large birefringence values because SO4 groups are nonpolar tetrahedral structures. However, by theoretical calculations, we found that polar tetrahedral S2O3 shows great improvements in anisotropy and second order polarizability compared to SO4 while the wide band gap can be preserved. On this basis, a nonlinear optical (NLO) crystal (NH4)2S2O3 with excellent performance was synthesized. As expected, it exhibited a strong SHG response (3.3× that of KDP), suitable birefringence (0.077@546 nm) and a short absorption edge (238 nm). According to the first principles calculations, the polar tetrahedral S2O3 group is the main source for the SHG response of (NH4)2S2O3. These findings indicate that the polar S2O3 group is a kind of NLO functional group with superior comprehensive properties, and has great potential to develop more NLO crystals with superior comprehensive properties.


Introduction

As the core materials of all-solid-state lasers, nonlinear optical (NLO) crystals have been widely applied to laser medical treatment, modern laser micromachining, laser communication, etc.1–7 In the past several decades, the exploration of UV (ultraviolet) NLO crystals with excellent comprehensive properties has always been a research hotspot in this field. Generally, a high-performance UV NLO crystal should meet the following conditions: a large second-harmonic generation (SHG) coefficient, appropriate birefringence and short UV cut-off edge. According to the anionic group theory,8 π-conjugated anionic groups, such as NO3, CO3, and BO3, are ideal functional units for UV NLO crystals.45 Through lots of research studies into the π-conjugated systems, many excellent UV NLO crystals were developed including β-Rb2Al2B2O7,9 KBe2BO3F2,10,11 MB5O7F3 (M = Ca, Sr),12,13 NH4B4O6F,14 ABCO3F (A = K, Rb; B = Mg, Ca, Sr),15–17 and M2(NO3)(OH)3 (M = Sr, Ba).18,19 In recent years, non-π-conjugated tetrahedral groups, such as SO4 and PO4 groups, which are beneficial for enlarging the band gaps, have also attracted researchers’ attention. However, since the tetrahedral groups SO4 and PO4 have approximately nonpolar Td symmetry, their contributions to SHG coefficients and birefringence are limited. In fact, most reported NLO sulfates and phosphates either show small birefringence or weak SHG responses. Therefore, one of the most attractive research directions is enhancing the polarizability anisotropy or hyperpolarizability of sulfates and phosphates.

Recently, it was considered an effective strategy to design polar tetrahedral anionic groups with enlarged polarizability anisotropy and enhanced hyperpolarizability by the elemental substitution method. In 2019, Lu et al. first reported the optimization of birefringence with a polar tetrahedral anionic group.46 Accordingly, SO3F,20,21 PO3F[thin space (1/6-em)]22,23 and PO2F2[thin space (1/6-em)]24 groups have been successively uncovered as excellent NLO functional groups. Compared to nonpolar tetrahedral groups, these polar tetrahedral groups not only could retain the wide band gap, but also could improve the SHG effect and birefringence. Taking steps along this path, the polar S2O3 tetrahedral group which is obtained by replacing an oxygen atom with a sulfur atom in SO4 could be expected to be a new NLO tetrahedral unit. The NLO-related properties of the S2O3 group and SO4 group were calculated and are listed for comparison (Table 1). It is noteworthy that the hyperpolarizability (βijk) of the S2O3 group was more than 12 times that of SO4 and the polarizability anisotropy of the S2O3 group was much superior to that of SO4, indicating that the S2O3 group indeed has more potential for developing strong SHG and large birefringence.

Table 1 Calculated NLO-related properties of isolated ideal S2O3 and SO4 tetrahedra

image file: d3qi00172e-u1.tif

  P x , Py, Pz δ [βmax] E g (eV)
SO4 0.00, 0.00, 0.00 45.1 47.8 6.98
S2O3 62.7, 62.7, 95.6 73.6 578 5.79


In addition to strong SHG response and large birefringence, excellent UV NLO materials also require a short UV absorption edge. Although the S2O3 group exhibits a wide band gap, the introduction of NLO-active cations with d–d or f–f transitions would cause a significant redshift of the UV absorption edge of the material, which should be avoided. Based on these considerations, ammonium thiosulfate was discovered by screening the ICSD database47 and is expected to be an excellent UV NLO material. First, (NH4)2S2O3 crystallized in the polar non-centrosymmetric (NCS) space group, C2, which is a basic requirement for producing SHG responses. Second, in (NH4)2S2O3, the polar S2O3 group aligned along a specific orientation is conducive to a large SHG response and birefringence. Third, the ammonium ion group without a d orbital is an ideal cation for maintaining the short absorption edge of the UV material.

Guided by the above ideas, we successfully grew (NH4)2S2O3 crystals by a simple solution method. The NLO properties and the relationship between the structure and the NLO properties of the crystal were also studied. It exhibited super comprehensive properties, including a strong SHG effect of 3.3× that of KDP, a suitable birefringence of 0.077@546 nm, and a short UV absorption edge of 238 nm. These results showed that (NH4)2S2O3 is an excellent UV NLO crystal and the polar tetrahedron S2O3 group is a promising group for constructing UV NLO materials.

Experimental

Regents and synthesis

Chemicals including Na2S2O3 (0.791 g, 5 mmol, 99%, Sinopharm) and NH4Cl (0.535 g, 10 mmol, 99%, Sinopharm) were of analytical grade and used without further purification and were purchased from commercial sources. Single crystals of (NH4)2S2O3 were synthesized via a facile water solution method. A mixture of NH4Cl and Na2S2O3 was dissolved in 10 mL of deionized water. This solution was stirred until it became clear. The solution was evaporated at 298 K and after several days, bulk colorless single crystals of (NH4)2S2O3 were obtained.

Powder X-ray diffraction

Powder X-ray diffraction measurements of (NH4)2S2O3 were carried out at room temperature with a Miniflex600 powder X-ray diffractometer set with Cu Kα radiation (λ = 1.5418 Å) in the range of 2θ = 5–75° and a 0.02° scan step. The experimental X-ray powder patterns of the pure sample of (NH4)2S2O3 agree well with the simulated X-ray powder patterns of the (NH4)2S2O3 single crystal models (Fig. S1).

Single crystal X-ray diffraction

A colorless single crystal of (NH4)2S2O3 was selected using an optical microscope to determine its crystal structure. The diffraction data were collected by using graphite-monochromatic Mo Kα radiation (λ = 0.71073 Å) on a Rigaku Mercury CCD diffractometer at room temperature. Then the data were integrated using the CrystalClear program, and based on the multi-scan method the data were corrected for absorption and refinement.25 The crystal structure of (NH4)2S2O3 was solved using the SHELXTL program in OLEX2[thin space (1/6-em)]26 with the direct methods. Besides, the PLATON27 program was used to check for higher symmetry elements, and none was found. The crystallographic data and structural refinement information of (NH4)2S2O3 are summarized in Table 2. Atomic coordinates and equivalent isotropic displacement parameters, bond lengths and angles, and anisotropic displacement parameters (Å2 × 103) for (NH4)2S2O3 are listed in Tables S1–S3.
Table 2 Crystallographic data and structural refinement for (NH4)2S2O3
a R 1(F) = ∑||Fo| − |Fc||/∑|Fo|wR2(Fo2) = [∑w(Fo2Fc2)2/∑w(Fo2)2]1/2.
Formula (NH4)2S2O3
Formula mass (amu) 70.07
Temperature (K) 293(2)
λ (Å) 1.34139
Crystal system Monoclinic
Space group C2
a (Å) 10.1484(7)
b (Å) 6.4862(5)
c (Å) 8.7175(7)
α (°) 90
β (°) 93.570(7)
γ (°) 90
V3) 572.71(8)
Z 8
ρ (calcd) (g cm−3) 1.625
μ (mm−1) 5.089
F(000) 280
θ (°) 4.421–60.581
Index range −13 ≤ h ≤ 13
−8 ≤ k ≤ 7
−10 ≤ l ≤ 11
Reflections collected/unique 3357/1084
R int 0.0493
Completeness to θ = 27.42° (%) 100.0
GOF on F2 1.123
R 1/wR2 [Fo2 > 2σ(Fo2)]a 0.0432/0.1186
R 1/wR2 (all data) 0.0442/0.1194
Absolute structure parameter 0.08(3)


Energy-dispersive X-ray spectroscopy analysis

Energy dispersive X-ray spectroscopy (EDS) analyses were performed on a scanning electron microscope (FESEM, SU-8010) equipped with an energy dispersive X-ray spectroscope. The (NH4)2S2O3 crystals were fixed on an aluminum sample table by using carbon conductive adhesive and tested with a focused beam with 12 μA emission current and 20 kV accelerating voltage (Fig. S2).

Thermal analysis

Thermogravimetric analysis (TGA) was performed using a NETZSCH STA449F3 simultaneous analyzer. Clean crystals of (NH4)2S2O3 were ground into powder, and then the powder was placed in an Al2O3 crucible. With an empty Al2O3 crucible as a reference, the sample was heated from 30 °C to 800 °C at a rate of 10 °C min−1 in an atmosphere of flowing N2 (Fig. S3).

UV-vis diffuse reflectance spectroscopy

By using BaSO4 powder as the standard (100% reflectance), UV-vis-NIR diffuse reflection data were measured and recorded using a PerkinElmer Lamda-950 UV/vis/NIR spectrophotometer at room temperature in the scan range of 200–2000 nm. In accordance with the Kubelka–Munk function, the reflection value was converted to an absorption value (Fig. S4).28,29

Birefringence

The birefringence of (NH4)2S2O3 was measured by using a Nikon ECLIPSE LV100 POL polarizing microscope equipped with a Berek Compensator and a 546 nm light source (Fig. S5). The birefringence was calculated using the following formula:
ΔR = Δn × T
ΔR represents the optical path difference, Δn represents the birefringence, and T represents the thickness of the crystal.

Second harmonic generation

Polycrystalline SHG signals of the (NH4)2S2O3 samples were investigated by using the Kurtz–Perry method30 and measured by using a Q-switched Nd:YAG solid-state laser with a fundamental wavelength of 1064 nm. Owing to the significant correlation between the SHG efficiency and the particle size of the powder, the (NH4)2S2O3 crystals were ground and divided into several different particle size ranges including 25–45,45–62, 62–75, 75–109, 109–150, and 150–212 μm. As a reference, the KDP crystals were also ground and sieved following the above procedures. Then they were placed in a fixed position and irradiated by using a pulsed laser (λ = 1064 nm). The outputs of second harmonic intensity from the samples and KDP were collected by using a RIGOL DS1052E 50 MHz oscilloscope.

First-principles calculations

First-principles calculations of crystal (NH4)2S2O3 were performed by using a CASTEP package based on density functional theory (DFT) in Material Studio software. The exchange–correlation energy was described by Perdew–Burke–Ernzerhof (PBE) in the generalized gradient approximation (GGA).31 To model the effective interaction between atomic nuclei and valence electrons of all the elements in (NH4)2S2O3, the optimized norm-conserving pseudopotentials in the Kleinman–Bylander form were used.32 The valence configurations of (NH4)2S2O3 including N 2s22p3, H 1s1, S 3s23p4, and O 2s22p4 were observed using a relatively small basis set without compromising calculation accuracy. A high kinetic energy cut-off of 274 eV and 1 × 2 × 2 Monkhorst–Pack33k-point meshes were selected for the numerical integration calculation of (NH4)2S2O3 in the Brillouin zone. According to the transition from valence bands to conduction bands of the electron, the imaginary part of the dielectric function was calculated. Based on the Kramers–Kronig34 transform the real part of the dielectric function was obtained and the refractive index was determined. The calculation results of SHG coefficients dij were obtained using a formula developed by Lin's group.35

Results and discussion

Crystal structure

Single crystals of (NH4)2S2O3 crystallized in the NCS space group C2 (no. 5) of the monoclinic crystal system. The crystal structure of (NH4)2S2O3 was composed of isolated NH4+ and S2O32− anionic groups. Within an asymmetric unit, there were three N atoms, two S atoms and three O atoms. Every S2O3 group was formed by one central S6+ atom coordinating with three O atoms and one S2− atom, in which the S–O bond lengths ranged from 1.465(2) to 1.485(3) Å and one S[double bond, length as m-dash]S bond was 1.975(2) Å. It is clear from Fig. 1a that there are two different orientations of S2O3 in the structure, S2O3(1) group and S2O3(2) group, both of which are arranged neatly along the b-axis. Obviously, in the a-axis and c-axis, the polarity of the S2O3(1) group and S2O3(2) group canceled each other out, respectively. The polarity of both S2O3 groups in the b-axis was superimposed, which was conducive to increasing the SHG effect. Looking at the crystal structure from the ac plane (Fig. 1b), there were three kinds of ammonium ions, NH4(1), NH4(2), and NH4(3), next to the S2O3 groups, which were connected by hydrogen bonds, and the adjacent S2O3 groups were connected by hydrogen bonds of the NH4(1) group and NH4(3) group. In contrast, the NH4(2) groups were dispersed among those [S2O3NH4]2 structures. The NH4(2) groups were connected to the S2O3 groups above or below them by hydrogen bonds to form a three-dimensional structure of (NH4)2S2O3 (Fig. 1c).
image file: d3qi00172e-f1.tif
Fig. 1 (a) S2O3 group aligned along the b axis. (b) Crystal structure representations of (NH4)2S2O3 in the ac plane and (c) in the bc plane.

Thermal analysis

As shown in Fig. S3, the title compound decomposed around 170 °C, and there was only one-step decomposition of (NH4)2S2O3 in the range of 170–390 °C. The 99.23% (cal. 100%) weight loss can be assigned to the complete decomposition of (NH4)2S2O3 into NH3, SO2, S and H2O with the evaporation of S.

UV-vis diffuse reflectance spectra

As shown in the UV-vis diffuse reflectance spectra (Fig. S4), the absorption edge of (NH4)2S2O3 was down to 238 nm. According to the Kubelka–Munk function,29 the optical band gap of (NH4)2S2O3 was deduced to be 4.44 eV, which was much larger than the band gap of Na10Cd(NO3)4(SO3S)4 (Eg = 3.74 eV).36

Birefringence

The single crystal of (NH4)2S2O3 was chosen for the measurement of birefringence (Fig. S5). The retardation value of the (NH4)2S2O3 crystal which was measured using a polarizing microscope (ZEISS Axio Scope A1) at a wavelength of 546 nm is 808.1 nm. Meanwhile, the thickness of the measured sample was 10.5 μm. Based on the formula ΔR = Δn × T, the experimental birefringence of (NH4)2S2O3 was 0.077 at 546 nm. The theoretical birefringence value calculated through the first-principles method (Δn = |nxnz|) was 0.08 at 546 nm (Fig. 2b) which was perfectly matched with the experimental birefringence value. Compared to other NLO sulfates, such as Rb2Bi2(SO4)2Cl4 (calc. 0.047 at 1064 nm),37 (NH4)2Bi2(SO4)2Cl4 (calc. 0.055 at 1064 nm),37 K2Bi2(SO4)2Cl4 (calc. 0.056 at 1064 nm),37 and Nb2O3(IO3)2(SO4) (calc. 0.220 at 1064 nm),38 the birefringence of (NH4)2S2O3 was well suitable and could be more in line with the requirements of the phase matching ability of crystals.
image file: d3qi00172e-f2.tif
Fig. 2 (a) SHG intensity measurements of (NH4)2S2O3. (b) Dispersion curves of the refractive index of (NH4)2S2O3.

NLO properties

The powder SHG measurement was performed by the Kurtz–Perry method.30 As shown in Fig. 2a, with KDP of the same particle size as the reference, the SHG response of (NH4)2S2O3 was approximately 3.3 times that of KDP at 1064 nm. As shown in Fig. 2a, the SHG responses increased when the particle sizes became larger, showing that (NH4)2S2O3 was phase-matchable. It is notable that the SHG effect of (NH4)2S2O3 was superior to those of NLO sulfates containing non-polar SO4 groups, such as Cs4Mg6(SO4)8 (0.2 × KDP),39 Li8NaRb3(SO4)6·2H2O (0.5 × KDP),40 (NH4)2Na3Li9(SO4)7 (0.5 × KDP),41 NH4NaLi2(SO4)2 (1.1 × KDP),41 and Li9Na3Rb2(SO4)7 (1.3 × KDP).42 The stronger SHG response of (NH4)2S2O3 should be due to its polar S2O3 groups with enhanced microscopic second order polarizability. Furthermore, as depicted in Fig. 1a, S2O3 was neatly arranged, which was also beneficial for improving the SHG effect of the title compound.

In order to further investigate the origin of the SHG contributions of the (NH4)2S2O3 crystal, the local dipole moments for the NH4 and S2O3 groups were calculated. The calculated results are shown in the ESI (Table S5). It is worth noting that the S2O3 polar tetrahedron has a large distortion and shows a large dipole moment of 13.67 Debye. It is even more important that in a unit cell of the title compound, the dipole moment of four S2O3 groups adopts a superimposed arrangement mode on the y-axis which results in a large dipole moment of −33.44 Debye, which is the key component for the whole dipole moment (−34.88 Debye). In polar materials, a large net dipole moment usually implies a large SHG response.43,44 Therefore, the large SHG response of (NH4)2S2O3 can be attributed to the polar tetrahedron S2O3 groups with a large distortion and their superimposed dipole moments along the y-axis.

Theoretical calculations

In order to gain insight into the optical properties of (NH4)2S2O3, the first-principles calculations of (NH4)2S2O3 were carried out. The calculated band gap of (NH4)2S2O3 was about 4.106 eV (Fig. 3b), which was in good agreement with the experimental value of 4.44 eV. N 2s22p3, H 1s1, S 3s23p4, and O 2s22p4 were used for calculations of the total and partial densities of states (DOS and PDOS) of (NH4)2S2O3 as shown in Fig. 3a. The optical properties of the compounds depended largely on the states near the forbidden band; so only the valence band (VB) top and the conduction band (CB) bottom were analyzed. It was obvious that the VBs were mostly contributed by S 3p and O 2p and the CBs were mainly due to S 3p, O 2p, N 2P, and H 1s, respectively. From the above analysis, it was clear that the optical properties of (NH4)2S2O3 were attributed mainly to the S2O3 polar tetrahedral groups.
image file: d3qi00172e-f3.tif
Fig. 3 (a) Total and partial DOS curves for (NH4)2S2O3. (b) Calculated electronic band structure for (NH4)2S2O3.

According to the formula proposed by Lin et al.,35 the SHG coefficient dij of (NH4)2S2O3 was calculated. (NH4)2S2O3 crystallized in the 2 point group. Considering the Kleinman symmetry, four independent SHG tensors including d21, d22, d23 and d25 of (NH4)2S2O3 were calculated (Fig. S7). The largest one was d23 of −1.05 pm V−1, which was about 2.69 times that of KDP (0.39 pm V−1). It matched well with the experimental result. As shown in Fig. S6, the occupied and unoccupied states of SHG-densities for the largest tensor d23 were calculated to further confirm the contributions of the S2O3 group to the macroscopic hyperpolarizability. In Fig. S6a, it is obvious that the sources of SHG densities in the occupied states were mainly contributed by the S2O3 group. For the unoccupied states, the SHG weighted densities were mostly contributed by the S2O3 group and NH4 group (Fig. S6b). It could be concluded that the S2O3 groups made great contributions to the SHG response in (NH4)2S2O3. Atom contributions to SHG coefficients were calculated to investigate the origin of the SHG effect intuitively.48–50 The SHG-contribution of the S2O3 tetrahedron to the SHG response of (NH4)2S2O3 was 66.9%, which was about twice as much as that of NH4. This result further confirmed that the S2O3 group played a major role in the improvement of the SHG response.

Conclusions

In a word, we have successfully synthesized a high-performance NLO thiosulfate (NH4)2S2O3via a water solution method under mild conditions. It exhibited a short cut-off edge of 238 nm, a large SHG effect that is 3.3 times that of KDP and sufficient birefringence for phase-matching. According to the theoretical calculations, the S2O3 group shows a great improvement in polarizability anisotropy and hyperpolarizability compared to SO4, which indicates that the polar tetrahedral S2O3 group is an outstanding functional group for building NLO materials. These materials containing the polar tetrahedral S2O3 group might show great improvements in the SHG effect while retaining a short UV absorption edge. It is clear that the S2O3 group has great potential for the synthesis of high performance nonlinear optical crystals.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant No. 22222510, 21975255 and 21921001), the Foundation of Fujian Science & Technology Innovation Laboratory (2021ZR202) and the Youth Innovation Promotion Association CAS (2019303).

References

  1. P. Becker, Borate materials in nonlinear optics, Adv. Mater., 1998, 10, 979 CrossRef CAS.
  2. Y. N. Xia, C. T. Chen, D. Y. Tang and B. C. Wu, New nonlinear optical crystals for UV and VUV harmonic generation, Adv. Mater., 1995, 7, 79 CrossRef CAS.
  3. G. H. Zou, C. S. Lin, H. Jo, G. Nam, T. S. You and K. M. Ok, Pb2BO3Cl: A Tailor-Made Polar Lead Borate Chloride with Very Strong Second Harmonic Generation, Angew. Chem., Int. Ed., 2016, 55, 12078–12082 CrossRef CAS PubMed.
  4. X. L. Wang, L. K. Chen, W. Li, H. L. Huang, C. Liu, C. Chen, Y. H. Luo, Z. E. Su, D. Wu, Z. D. Li, H. Lu, Y. Hu, X. Jiang, C. Z. Peng, L. Li, N. L. Liu, Y. A. Chen, C. Y. Lu and J. W. Pan, Experimental Ten-Photon Entanglement, Phys. Rev. Lett., 2016, 117, 210502 CrossRef PubMed.
  5. H. Tian, C. S. Lin and X. Zhao, Design of a new ultraviolet nonlinear optical material KNO3SO3NH3 exhibiting an unexpected strong second harmonic generation response, Mater. Today Phys., 2022, 28, 100849 CrossRef CAS.
  6. X. Dong, L. Huang, C. Hu, H. Zeng, Z. Lin, X. Wang, K. M. Ok and G. Zou, CsSbF2SO4: An Excellent Ultraviolet Nonlinear Optical Sulfate with a KTiOPO4 (KTP)-type Structure, Angew. Chem., Int. Ed., 2019, 58, 6528–6534 CrossRef CAS PubMed.
  7. H. Tian, C. S. 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., 2023 DOI:10.31635/ccschem.022.202202582.
  8. C. T. Chen, A Localized Quantal Theoretical Treatment Based on an Anionic Coordination Polyhedron Model for the EO and SHG Effects in Crystals of the Mixed-Oxided Types, Sci. Sin., 1979, 22, 756–776 CAS.
  9. T. T. Tran, N. Z. Koocher, J. M. Rondinelli and P. S. Halasyamani, Beryllium-Free β-Rb2Al2B2O7 as a Possible Deep-Ultraviolet Nonlinear Optical Material Replacement for KBe2BO3F2, Angew. Chem., Int. Ed., 2017, 56, 2969–2973 CrossRef CAS PubMed.
  10. B. C. Wu, D. Y. Tang, N. Ye and C. T. Chen, Linear and nonlinear optical properties of the KBe2BO3F2 (KBBF) crystal, Opt. Mater., 1996, 5, 105–109 CrossRef CAS.
  11. C. T. Chen, Z. Y. Xu, D. Q. Deng, J. Zhang, G. K. L. Wong, B. C. Wu, N. Ye and D. Y. Tang, The vacuum ultraviolet phase-matching characteristics of nonlinear optical KBe2BO3F2 crystal, Appl. Phys. Lett., 1996, 68, 2930–2932 CrossRef CAS.
  12. Z. Z. Zhang, Y. Wang, B. B. Zhang, Z. H. Yang and S. L. Pan, CaB5O7F3: A Beryllium-Free Alkaline-Earth Fluorooxoborate Exhibiting Excellent Nonlinear Optical Performances, Inorg. Chem., 2018, 57, 4820–4823 CrossRef CAS PubMed.
  13. M. Mutailipu, M. Zhang, B. B. Zhang, L. Y. Wang, Z. L. Yang, X. Zhou and S. L. Pan, SrB5O7F3 Functionalized with [B5O9F3]6 Chromophores: Accelerating the Rational Design of Deep-Ultraviolet Nonlinear Optical Materials, Angew. Chem., Int. Ed., 2018, 57, 6095–6099 CrossRef CAS PubMed.
  14. G. Q. Shi, Y. Wang, F. F. Zhang, B. B. Zhang, Z. H. Yang, X. L. Hou, S. L. Pan and K. R. Poeppelmeier, Finding the Next Deep-Ultraviolet Nonlinear Optical Material: NH4B4O6F, J. Am. Chem. Soc., 2017, 139, 10645–10648 CrossRef CAS PubMed.
  15. T. T. Tran, J. Young, J. M. Rondinelli and P. S. Halasyamani, Mixed-Metal Carbonate Fluorides as Deep-Ultraviolet Nonlinear Optical Materials, J. Am. Chem. Soc., 2017, 139, 1285–1295 CrossRef CAS PubMed.
  16. T. T. Tran, J. G. He, J. M. Rondinelli and P. S. Halasyamani, RbMgCO3F: A New Beryllium-Free Deep-Ultraviolet Nonlinear Optical Material, J. Am. Chem. Soc., 2015, 137, 10504–10507 CrossRef CAS PubMed.
  17. G. H. Zou, N. Ye, L. Huang and X. S. Lin, Alkaline-Alkaline Earth Fluoride Carbonate Crystals ABCO3F (A = K, Rb, Cs; B = Ca, Sr, Ba) as Nonlinear Optical Materials, J. Am. Chem. Soc., 2011, 133, 20001–20007 CrossRef CAS PubMed.
  18. L. Huang, G. H. Zou, H. Q. Cai, S. C. Wang, C. S. Lin and N. Ye, Sr2(OH)3NO3: the first nitrate as a deep UV nonlinear optical material with large SHG responses, J. Mater. Chem. C, 2015, 3, 5268–5274 RSC.
  19. X. H. Dong, L. Huang, Q. Y. Liu, H. M. Zeng, Z. E. Lin, D. G. Xu and G. H. Zou, Perfect balance harmony in Ba2NO3(OH)3: a beryllium-free nitrate as a UV nonlinear optical material, Chem. Commun., 2018, 54, 5792–5795 RSC.
  20. 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, 37, 15978–15981 CrossRef PubMed.
  21. Y. Sun, C. Lin and S. F, K2(BeS)O4F2: a novel fluorosulfate with unprecedented 1D [(BeO3F)-(SO3F)] chains exhibiting large birefringence, Inorg. Chem. Front., 2022, 9, 6490–6497 RSC.
  22. B. B. Zhang, G. Q. Shi, Z. H. Yang, F. F. Zhang and S. L. Pan, Fluorooxoborates: Beryllium-Free Deep-Ultraviolet Nonlinear Optical Materials without Layered Growth, Angew. Chem., Int. Ed., 2017, 56, 3916–3919 CrossRef CAS PubMed.
  23. L. Xiong, J. Chen, J. Lu, C. Y. Pan and L. M. Wu, Mono-fluorophosphates: A New Source of Deep-Ultraviolet Nonlinear Optical Materials, Chem. Mater., 2018, 30, 7823–7830 CrossRef CAS.
  24. B. B. Zhang, G. P. Han, Y. Wang, X. L. Chen, Z. H. Yang and S. L. Pan, Expanding Frontiers of Ultraviolet Nonlinear Optical Materials with Fluorophosphates, Chem. Mater., 2018, 30, 5397–5403 CrossRef CAS.
  25. R. H. Blessing, An Empirical Correction for Absorption Anisotropy, Acta Crystallogr., Sect. A: Found. Adv., 1995, 51, 33–38 CrossRef PubMed.
  26. O. V. Dolomanov, L. J. Bourhis, R. J. Gildea, J. A. K. Howard and H. Puschmann, OLEX2: a complete structure solution, refinement and analysis program, J. Appl. Crystallogr., 2009, 42, 339–341 CrossRef CAS.
  27. A. L. Spek, Single-crystal structure validation with the program PLATON, J. Appl. Crystallogr., 2003, 36, 7–13 CrossRef CAS.
  28. J. Tauc, Absorption Edge and Internal Electric Fields in Amorphous Semiconductors, Mater. Res. Bull., 1970, 5, 721–729 CrossRef CAS.
  29. P. Kubelka and F. Munk, An article on optics of paint layers, Z. Tech. Phys., 1931, 12(1), 886–892 Search PubMed.
  30. S. K. Kurtz and T. T. Perry, A Powder Technique for Evaluation of Nonlinear Optical Materials, J. Appl. Phys., 1968, 39, 3798–3813 CrossRef CAS.
  31. J. P. Perdew, A. Ruzsinszky, G. I. Csonka, O. A. Vydrov, G. E. Scuseria, L. A. Constantin, X. Zhou and K. Burke, Restoring the Density-Gradient Expansion for Exchange in Solids and Surfaces, Phys. Rev. Lett., 2008, 100, 136406 CrossRef PubMed.
  32. A. M. Rappe, K. M. Rabe, E. Kaxiras and J. D. Joannopoulos, Optimized Pseudopotentials, Phys. Rev. B: Condens. Matter Mater. Phys., 1990, 41, 1227–1230 CrossRef PubMed.
  33. H. J. Monkhorst and J. D. Pack, Special Points for Brillouin-Zone Integrations, Phys. Rev. B: Solid State, 1976, 13, 5188–5192 CrossRef.
  34. E. D. Palik, Handbook of Optical Constants of Solids, Academic Press, 1985 Search PubMed.
  35. J. Lin, M. H. Lee, Z. P. Liu, C. T. 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, 13380–13389 CrossRef CAS; Z. S. Lin, X. X. Jiang, L. Kang, P. F. Gong, S. Y. Luo and M. H. Lee, Theoretical calculations and predictions of the nonlinear optical coefficients of borate crystals, J. Phys. D: Appl. Phys., 2014, 47, 253001 CrossRef.
  36. Y. Liu, Y. Liu, Z. Lin, Y. Li, Q. Ding, X. Chen, L. Li, S. Zhao, M. Hong and J. Luo, Nonpolar Na10Cd(NO3)4(SO3S)4 Exhibits a Large Second-Harmonic Generation, CCS Chem., 2021, 4, 526–531 CrossRef.
  37. K. C. Chen, Y. Yang, G. Peng, S. D. Yang, T. Yan, H. X. Fan, Z. S. Lin and N. Ye, A2Bi2(SO4)2Cl4 (A = NH4, K, Rb): achieving a subtle balance of the large second harmonic generation effect and sufficient birefringence in sulfate nonlinear optical materials, J. Mater. Chem. C, 2019, 7, 9900–9907 RSC.
  38. H. X. Tang, Y. X. Zhang, C. Zhuo, R. B. Fu, H. Lin, Z. J. Ma and X. T. Wu, A Niobium Oxyiodate Sulfate with a Strong Second Harmonic-Generation Response Built by Rational Multi-Component Design, Angew. Chem., Int. Ed., 2019, 58, 3824–3828 CrossRef CAS PubMed.
  39. Y. Q. Li, J. H. Luo, X. H. Ji and S. G. Zhao, A Short-wave UV Nonlinear Optical Sulfate of High Thermal Stability, Chin. J. Struct. Chem., 2020, 39(3), 485–492 CAS.
  40. Y. Q. Li, S. E. Zhao, P. Shan, X. F. Li, Q. R. Ding, S. Liu, Z. Y. Wu, S. S. Wang, L. N. Li and J. H. Luo, Li8NaRb3(SO4)6·2H2O as a new sulfate deep-ultraviolet nonlinear optical material, J. Mater. Chem. C, 2018, 6, 12240–12244 RSC.
  41. Y. Q. Li, F. Liang, S. G. Zhao, L. N. Li, Z. Y. Wu, Q. R. Ding, S. Liu, Z. S. Lin, M. C. Hong and J. H. Luo, Two Non-π-Conjugated Deep-UV Nonlinear Optical Sulfates, J. Am. Chem. Soc., 2019, 141, 3833–3837 CrossRef CAS PubMed.
  42. Y. Li, C. Yin, X. Yang, X. Kuang, J. Chen, L. He, Q. Ding, S. Zhao, M. Hong and J. Luo, Nonlinear Optical Switchable Sulfate of Ultrawide Bandgap, CCS Chem., 2021, 3, 2298–2306 CrossRef CAS.
  43. H. Yu, H. Wu, S. Pan, Z. Yang, X. Hou, X. Su, Q. Jing, K. R. Poeppelmeier and J. M. Rondinelli, Cs3Zn6B9O21: A Chemically Benign Member of the KBBF Family Exhibiting the Largest Second Harmonic Generation Response, J. Am. Chem. Soc., 2014, 136, 1264 CrossRef CAS PubMed.
  44. H. Wu, S. Pan, K. R. Poeppelmeier, H. Li, D. Jia, Z. Chen, X. Fan, Y. Yang, J. M. Rondinelli and H. Luo, K3B6O10Cl: A New Structure Analogous to Perovskite with a Large Second Harmonic Generation Response and Deep UV Absorption Edge, J. Am. Chem. Soc., 2011, 133, 7786 CrossRef CAS PubMed.
  45. M. Mutailipu, K. R. Poeppelmeier and S. Pan, Borates: A Rich Source for Optical Materials, Chem. Rev., 2021, 121(3), 1130–1202 CrossRef CAS PubMed.
  46. J. Lu, J. Yue, L. Xiong, W. Zhang, L. Chen and L. Wu, Uniform Alignment of Non-π-Conjugated Species Enhances Deep Ultraviolet Optical Nonlinearity, J. Am. Chem. Soc., 2019, 141, 8093–8097 CrossRef CAS PubMed.
  47. S. T. Teng, H. Fuess and J. W. Bats, Ammonium thiosulphate, Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem., 1979, 35, 1682–1684 CrossRef.
  48. J. E. Sipe and E. Ghahramani, Nonlinear-Optical Response Of Semiconductors In The Independent-Particle Approximation, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 48, 11705–11722 CrossRef CAS PubMed.
  49. A. Zhou, C. Lin, B. Li, W. Cheng, Z. Guo, Z. Hou, F. Yuan and G. Chai, Ba6In6Zn4Se19: A High Performance Infrared Nonlinear Optical Crystal with [InSe3]3− Trigonal Planar Functional Motifs, J. Mater. Chem. C, 2020, 8(23), 7947–7955 RSC.
  50. C. Lin, A. Zhou, W. Cheng, N. Ye and G. Chai, Atom-Resolved Analysis of Birefringence of Nonlinear Optical Crystals by Bader Charge Integration, J. Phys. Chem. C, 2019, 123, 31183–31189 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3qi00172e

This journal is © the Partner Organisations 2023
Click here to see how this site uses Cookies. View our privacy policy here.