A(NH2SO3) (A=Li, Na): Two ultraviolet transparent sulfamates exhibiting second harmonic generation response

Cuiwu MO Gangmin ZHANG Chao WU Zhipeng HUANG Chi ZHANG

Citation:  Cuiwu MO, Gangmin ZHANG, Chao WU, Zhipeng HUANG, Chi ZHANG. A(NH2SO3) (A=Li, Na): Two ultraviolet transparent sulfamates exhibiting second harmonic generation response[J]. Chinese Journal of Inorganic Chemistry, 2024, 40(7): 1387-1396. doi: 10.11862/CJIC.20240045 shu

两种具有二次谐波响应的紫外透过氨基磺酸盐A(NH2SO3)(A=Li、Na)

    通讯作者: 吴超, wuc@tongji.edu.cn
    张弛, chizhang@tongji.edu.cn
  • 基金项目:

    国家自然科学基金 51432006

    国家自然科学基金 52002276

    教育部长江学者创新团队计划 IRT14R23

    教育部和国家外国专家局111计划 B13025

摘要: 通过简便的蒸发方法得到了2种碱金属磺酸盐非线性光学(NLO)晶体,即Li(NH2SO3)和Na(NH2SO3)。Li(NH2SO3)以极性空间群Pca21(编号29)结晶。Li(NH2SO3)的结构可以描述为由[LiO4]7-多面体通过共角连接与NH2SO3-四面体相互连接而形成的三维网络。Na(NH2SO3)以极性空间群P212121(编号19)结晶。Na(NH2SO3)的结构可以描述为由扭曲的[NaO6]11-八面体通过共角连接与NH2SO3-四面体相互连接而形成的三维网络。紫外可见近红外光谱表明,Li(NH2SO3)和Na(NH2SO3)分别具有5.25和4.81eV的大光学带隙。粉末二次谐波发生(SHG)测量显示,Li(NH2SO3)和Na(NH2SO3)的SHG强度分别为KH2PO4的0.32倍和0.31倍。第一原理计算证实,非线性光学性能主要来自氨基磺酸阴离子和碱金属氧阴离子多面体的协同作用。

English

  • Nonlinear optical (NLO) crystals can convert laser frequency and have been widely used in semiconductor manufacturing, lithography, attosecond pulse generation, and advanced instruments[1-5]. So far, a large number of inorganics, organics, organometallics, inorganic-organic hybrid compounds, polymers, and inorganic glasses have been found to have second-harmonic generation (SHG) activity. Currently, only a few NLO crystals have been commercially applied, including KBe2BO3F2[6-8], β-BaB2O4[9], LiB3O5[10], CsB3O5[11], KH2PO4 (KDP)[12], KTiOPO4[13], and AgGaS2[14]. However, these crystals all have some inherent drawbacks and cannot be called “perfect” crystals. Therefore, it will still take a long time to explore more promising crystals.

    NLO crystals can be divided into four groups according to their transparent wavelength range: deep ultraviolet (DUV, < 200 nm), ultraviolet (UV, 200-400 nm), visible near-infrared (Vis-NIR, 400 nm-2.5 µm), middle and far-infrared (MFIR, 2.5-20 µm) [15-16]. UV NLO crystals are the key component of frequency conversion to generate short-wavelength laser light that is intrinsic in cutting-edge laser technology and funda-mental science[17]. For a long time, the exploration of UV NLO crystals has mainly focused on π-conjugated systems, such as BO33-[18], CO32-[19], and NO3-[20]. Based on these excellent π-conjugated NLO functional units, a large number of crystals with excellent NLO performance were reported, such as NH4B4O6F[21], NaB4O6F[22], RbMgCO3F[23], NaZnCO3(OH)[24], Sr2(OH)3NO3[25], and Rb2Na(NO3)3[26]. In recent years, UV NLO crystals based on non-π-conjugated systems (PO43- and SO42-) have gradually gained attention. Wherein, PO43- unit is usually conducive to the formation of short UV cut-off edges and tends to connect with other elements or units to form 0D to 3D crystal structures, such as chain, layer, and network structures. Many UV NLO phosphate crystals have been discovered successfully, including Ba3P3O10X[27], RbBa2(PO3)5[28], Rb2Ba3(P2O7)2[28], and LiCs2PO4[29]. Compared to the PO43- unit, theoretically speaking, the SO42- unit possesses larger microscopic polarizability and polarizability anisotropy, which is more conducive to generating large macroscopic NLO effects. In addition, SO42- has an optical band gap equivalent to PO43-, which can promote the transmission of sulfate NLO crystals in UV and even DUV. In recent years, NH4NaLi2(SO4)2[30], (NH4)2Na3Li9(SO4)7[30], Li2KRb(SO4)2[31], and MF2(SO4) (M=Zr, Hf) [32] were reported with excellent NLO performance. However, PO43- and SO42- are both isotropic tetrahedral building units with almost nonpolar symmetry, which limits their contribution to SHG coefficient and birefringence. In addition, it is reported that the heteroatomic tetrahedra have some obvious advantages over the regular tetrahedra in NLO hyperpolarization and optical anisotropy, which helps to achieve a comprehensive balance between the large band gap, high SHG response, and proper birefringence of NLO crystals. These heteroatomic tetrahedra include BOxF4-x (x=1-3)[22, 33], PO x F4-x (x= 1-3)[34-35], and SiOx F6-x (x=1-5)[36-37]. In 2021, Ye et al.[38] discovered a new non-π-conjugated unit NH2SO3- through first-principles calculation, and successfully synthesized two excellent NLO crystals Sr(NH2SO3)2 and Ba(NH2SO3)2. The calculation results show that NH2SO3- has a more obvious polarizability anisotropy, greater hyperpolarizability (about 1.92 times that of SO42-), and a wider optical band gap (8.17 eV) than SO42-. It provides a new way to explore non-π-conjugated systems.

    Based on the above considerations, we introduced alkali metal cations into the NH2SO3- unit, and successfully obtained two alkali-metal sulfamates NLO crystals, Li(NH2SO3) and Na(NH2SO3), through the facile evaporation method. Although the structures of these two compounds have been reported, the physicochemical properties of these two compounds have not been reported, so far. Herein, we provide a detailed report on the physicochemical and NLO properties of these two compounds. The structure-function relationships of these two crystals have also been studied by theoretical calculations.

    Sulfamic acid (NH2SO3H, 99.99%, Xiya Reagent), sodium carbonate (Na2CO3, 99.80%, Xiya Reagent), lithium carbonate (Li2CO3, 99.00%, Xiya Reagent) and deionized water were commercially available and used as received without further purification.

    Li2CO3 (1.0 mmol, 0.074 g) and NH2SO3H (2.0 mmol, 0.194 g) were mixed and poured into 3 mL deionized water, fully stirred at room temperature, and then placed in the fume hood. After 2-3 d, colorless rod-shaped crystals of Li(NH2SO3) could be obtained, with a yield of about 73% (based on the amount of Li used).

    The preparation of Na(NH2SO3) was the same as that of Li(NH2SO3) except that Li2CO3 was replaced by Na2CO3 (1.0 mmol, 0.106 g). Colorless rod-shaped crystals were obtained with a yield of about 70% (based on the amount of Na used).

    Single-crystal X-ray diffraction data collection of Li(NH2SO3) and Na(NH2SO3) was carried out on a Bruker D8 VENTURE CMOS X-ray source with Mo radiation (λ=0.071 073 nm) at room temperature. APEX Ⅱ software was applied to collect and reduce data. For Li(NH2SO3), in a range of 2.505° < θ < 27.469°, a total of 6 728 reflections were collected and 1548 were independent with Rint=0.039 8, of which 1 026 were observed with I > 2σ(I). For Na(NH2SO3), in a range of 2.598° < θ < 27.096°, a total of 9 539 reflections were collected and 2 372 were independent with Rint=0.030 2, of which 1 282 were observed with I > 2σ(I). Semi-empirical absorption corrections based on equivalent reflections were applied for both data sets using the APEX Ⅱ program. The structure was solved by direct methods and refined on F2 by full-matrix least-squares methods using the OLEX2 software package[39]. All non-hydrogen atoms were refined anisotropically. The detailed crystal data and structural refinement informa-tion for Li(NH2SO3) and Na(NH2SO3) are summarized in Table S1 (Supporting information). Selected bond distances (nm) and angles (°) are given in Tables S2 and S3, while atomic coordinates and equivalent isotropic displacement parameters as well as bond valence sums are collected in Tables S4 and S5. The powder X-ray diffraction data of each sample were collected on a Bruker D8 X-ray diffractometer equipped with Cu radiation (λ =0.154 18 nm) in a 2θ range of 5°-80° with a step size of 0.02° at room temperature, the working voltage and current were 40 kV and 40 mA respectively (Fig.S1).

    The IR spectra were recorded on a Nicolet iS10 Fourier transform IR spectrometer (resolution 4 cm-1, spectral range 400-4 000 cm-1). An optical diffuse reflectance spectrum was obtained using a Cary 5000 UV-Vis-NIR spectrophotometer at room temperature. A BaSO4 plate was used as a standard (100% reflectance). The absorption spectra of A(NH2SO3) (A=Li, Na) were calculated from the reflectance spectrum using the Kubelka-Munk function α/S = (1-R)2/(2R), where α is the absorption coefficient, S is the scattering coefficient that is essentially wavelength-independent when the particle size is larger than 5 µm, and R is the reflectance[40].

    A Netzsch STA 409PC thermal analyzer was used to analyze the thermal stabilities of A(NH2SO3) (A=Li, Na). The samples were heated from 30 to 800 ℃ with a heating rate of 20 ℃·min-1 in nitrogen gas.

    The SHG intensities of A(NH2SO3) (A=Li, Na) were measured according to the modified method of Kurtz and Perry[41]. A Q-switched Nd∶YAG laser with 1 064 nm radiation was employed for the visible SHG study. Because the SHG efficiency is related to the particle size, the polycrystalline samples of Li(NH2SO3) and Na(NH2SO3) were ground and sieved into several particle size ranges (0-26 µm, 26-50 µm, 50-74 µm, 74-105 µm, 105-150 µm, 150-200 µm). Crystalline KDP with the same particle size ranges was used as references.

    First-principles calculations on A(NH2SO3) (A=Li, Na) were performed using the CASTEP package[42], a total energy package based on pseudopotential density functional theory (DFT) [43]. The correlation-exchange terms in the Hamiltonian were described by the functional developed by Perdew, Burke, and Ernzerhof in the generalized gradient approximation form[44-45]. Optimized norm-conserving pseudopotentials[46] in the Kleinman-Bylander form were adopted to model the effective interaction between the valence electrons and atom cores, which allows the choice of a relatively small plane-wave basis set without compromising the computational accuracy. A kinetic energy cutoff of 850 eV and dense Monkhorst-Pack[47] k-point meshes spanning less than 0.000 015 nm3 in the Brillouin zone were chosen.

    Li(NH2SO3) crystallizes into an orthorhombic polar space group Pca21. There are sixteen independent atoms in one asymmetric unit of the crystal, including two Li atoms, two S atoms, two N atoms, six O atoms, and four H atoms (Fig. S2a). Each Li atom coordinates with four O atoms from four NH2SO3- units through corner-sharing, forming [LiO4]7- tetrahedra with Li—O distances in a range of 0.188 3(8)-0.196 0(9) nm. Each S atom forms four single bonds with three O atoms and one N atom to form NH2SO3- tetrahedra with the S—O and S—N bond lengths varying from 0.144 2(3) to 0.146 8(4) nm and from 0.161 6(6) to 0.162 4(6) nm, respectively, and the O—S—O bond angles are in a range of 110.1(2)°-113.9(2)°, the O—S—N bond angles are in a range of 103.8(3)°-112.1(2)°. The [LiO4]7- unit is connected with NH2SO3- units to form a 3D net-work structure. The Li atoms are connected by the O—S—O bonds of the three and four NH2SO3- units to form the [Li4O8S4] ring and the [Li3O6S3] ring, respectively (Fig. 1a and 1c). The [Li4O8S4] ring and the [Li3O6S3] ring grow infinitely in the ac and ab planes, respectively (Fig. 1b and 1d). The NH2SO3- units act as interlayer linkers connecting Li atoms, thus forming the 3D network structure (Fig. 1e).

    Figure 1

    Figure 1.  Crystal structure of Li(NH2SO3): (a)[Li4O8S4] ring; (b) [Li4O8S4] layer parallel to the ac plane; (c) [Li3O6S3] ring; (d) [Li3O6S3] layer parallel to the ab plane; (e) 3D framework structure of Li(NH2SO3)

    Although Na(NH2SO3) shares a similar chemical formula as Li(NH2SO3), it crystallizes into orthorhombic polar space group P212121. One asymmetric unit of Na(NH2SO3) contains twenty-four independent atoms, including three Na atoms, nine O atoms, three N atoms, three S atoms, and six H atoms (Fig. S2b). There are three different coordination modes of Na atoms, and all of them are six-coordinated with six O atoms from six NH2SO3- units through corner-sharing, forming three twisted [NaO6]11- octahedral with Na—O distances in a range of 0.230 7(6)-0.239 6(5) nm, and then three different [NaO6]11- units connect by corner-sharing, form-ing the [Na3 O11] unit (Fig. 2a and 2b). For NH2SO3- tetrahedra, the S—O and S—N bond lengths vary from 0.144 0(5) to 0.145 8(5) nm, and the O—S—O bond angles are in a range of 101.7(2)°-13.5(3)°. It can be observed that [Na3O11] units are tightly aligned together to form a chain structure by sharing O atoms (Fig. 2c). The NH2SO3- units act as interlayer linkers connecting Na atoms, thus forming the 3D network structure (Fig. 2d).

    Figure 2

    Figure 2.  (a) Arrangement of NH2SO3- units in Na(NH2SO3); (b) Coordination environment of Na atoms and the [Na3O11] unit; (c) Chain structure of [Na3O11] units; (d) 3D framework of Na(NH2SO3)

    Symmetry codes: #1: 0.5+x, 0.5-y, 1-z; #2:-0.5+x, 0.5-y, 1-z; #3: x, y, z; #4:-1+x, y, z; #5:-x, -0.5+y, 1.5-z.

    To investigate the thermal stability of Li(NH2SO3) and Na(NH2SO3), TGA was performed (Fig. 3). The TG curves showed that two crystalline samples Li(NH2SO3) and Na(NH2SO3) were stable up to 50 and 250 ℃, respectively. Li(NH2SO3) underwent two steps of weight loss when the temperature increased. The first step started at 50 ℃ and ended at 200 ℃, which corresponds to the loss of 0.5 molecules of N2 and 1 molecule of H2. The observed weight loss of 15.39% is very close to the calculated value of 15.53% for Li(NH2SO3). The second weight loss (84.86%) occurred between 200 and 670 ℃, which is attributable to the elimination of 0.5 SO2 and 0.5 Li2SO4 molecules (Calcd. 84.47%). The decomposition reaction is Li(NH2SO3)→ 0.5N2+H2 +0.5SO2 +0.5Li2SO4. Na(NH2SO3) also underwent two steps of weight loss when the temperature increased. The first step occurred between 250 and 400 ℃, corresponding to the loss of 0.5 molecules of N2 and 1 molecule of H2. The observed weight loss of 11.17% is close to the calculated 13.44%. The second step occurred between 400 and 450 ℃, corresponding to the loss of 0.5 SO2 molecules. The observed weight loss of 32.75% is consistent with the calculated 26.89%. The remaining part of the 0.5 Na2SO4 molecules gradually lost weight as the temperature increased until the sample lost all weight (due to the maximum temperature set for testing at 800 ℃, subsequent curves cannot be made). The decomposition reaction is Na(NH2SO3)→0.5N2+H2+0.5SO2+0.5Na2SO4.

    Figure 3

    Figure 3.  TG curves of (a) Li(NH2SO3) and (b) Na(NH2SO3)

    The IR spectra for the compounds Li(NH2SO3) and Na(NH2SO3) were measured in a wavenumber range of 400-4 000 cm-1 at room temperature (Fig. 4). The IR spectra of the two compounds were very similar. Taking Li(NH2SO3) as an example, the absorption bands occurring at 3 356 and 3 277 cm-1 are attributable to the stretching vibration of the N—H bond, and the absorption band occurring at 1 573 cm-1 is ascribed to the in-plane bending vibration of the N—H bond, the absorption bands occurring at 738 and 673 cm-1 are ascribed to the out of plane bending vibration of the N—H bond. The absorption bands observed at 1 241, 1 156, and 1 058 cm-1 are ascribed to S—O vibrations. The absorption band at 557 cm-1 is ascribed to the vibration of the Li—O bond.

    Figure 4

    Figure 4.  IR spectra of (a) Li(NH2SO3) and (b) Na(NH2SO3)

    The UV-Vis-NIR diffuse reflectance spectra of Li(NH2SO3) and Na(NH2SO3) are shown in Fig. 5. The spectra indicated that the optical band gaps for compounds Li(NH2SO3) and Na(NH2SO3) were ca. 5.25 and 4.81 eV, respectively, corresponding to absorption edg-es of 236 and 258 nm, respectively.

    Figure 5

    Figure 5.  UV-Vis-NIR diffuses reflectance spectra of (a) Li(NH2SO3) and (b) Na(NH2SO3)

    The inset shows the corresponding band gap.

    Based on the polar structure with a space group of Pca21 and P212121, the powder SHG response was measured by the Kurtz-Perry method. The SHG intensities increased with the increasing of particle size in a range of 26-280 µm (Fig. 6a), indicating both Li(NH2SO3) and Na(NH2SO3) are phase-match crystals at the 1 064 nm laser. As shown in Fig. 6b, Li(NH2SO3) and Na(NH2SO3) showed SHG intensities of about 0.32 times and 0.31 times that of KDP in the particle size range of 105-150 µm, respectively. The SHG response of crystals correlates with the induced dipole moments reflecting the ease of electron motion.

    Figure 6

    Figure 6.  (a) Phase-matching curve of Li(NH2SO3) and Na(NH2SO3) with 1064 nm laser radiation; (b) Oscilloscope traces of the SHG signals for powders of Li(NH2SO3) and Na(NH2SO3) and KDP in the same particle size range of 105-150 µm

    KDP was used as a reference for the SHG measurement at 1 064 nm.

    To further understand the structure-function relationship of Li(NH2SO3) and Na(NH2SO3), theoretical calculations using the CASTEP program based on DFT were performed. The band structure diagram (Fig. 7a and 7b) shows that the minimum conduction band (CBM) and maximum valence band (VBM) of Li(NH2SO3) and Na(NH2SO3) were located at the G point, with direct band gaps of 5.47 and 4.99 eV, respectively, which is greater than the experimental value of 5.25 and 4.81 eV. The discrepancy can be attributed to the limitations of the exchange-correlation function GGA-PBE[48-49]. Fig. 7c and 7d show the total density of states (DOS) and partial density of states (PDOS) diagrams of Li(NH2SO3) and Na(NH2SO3). Due to their similarity, we will take Li(NH2SO3) as an example to provide a detailed introduction. Observing the DOS of Li(NH2SO3), it is evident that the conduction band located between-10 and 0 eV is mainly occupied by O2p, N2p, and S3p orbitals, along with smaller contributions from Li1s, Li2p, H1s, and S3s orbitals. The valence band from 0 to 10 eV is mainly composed of S3p, Li2p, and O2p orbitals, as well as some Li1s, S3s, H1s, and N2s orbitals. The above calculations indicate that Li, O, N, and S atoms significantly affect the band structure and optical properties of Li(NH2SO3), which are mainly attributed to the synergistic effect of the [LiO4]7- polyhedra and the NH2SO3- tetrahedra. Compared with Na(NH2SO3), the difference in band gaps between the two crystals is mainly due to the difference between the [LiO4]7- polyhedra and the [NaO6]11- octahedra.

    Figure 7

    Figure 7.  Band structure diagram of (a) Li(NH2SO3) and (b) Na(NH2SO3); DOS and PDOS diagrams of (c) Li(NH2SO3) and (d) Na(NH2SO3)

    In summary, we have successfully synthesized two alkali-metal sulfamates nonlinear optical crystals, Li(NH2SO3) and Na(NH2SO3), through the facile evaporation method. Li(NH2SO3) and Na(NH2SO3) crystallize in the space group Pca21 and P212121, respectively. The UV-Vis-NIR spectra demonstrated that Li(NH2SO3) and Na(NH2SO3) possessed large optical band gaps of 5.25 and 4.81 eV, respectively. Powder SHG measurements showed that the SHG intensity of Li(NH2SO3) and Na(NH2SO3) were 0.32 times and 0.31 times that of KDP, respectively. This research demonstrates that Li(NH2SO3) and Na(NH2SO3) are two promising UV NLO materials.

    Supporting information is available at http://www.wjhxxb.cn


    1. [1]

      Cyranoski D. Materials science: China's crystal cache[J]. Nature, 2009, 457(7232):  953-955. doi: 10.1038/457953a

    2. [2]

      Mutailipu M, Zhang M, Yang Z H, Pan S L. Targeting the next generation of deep-ultraviolet nonlinear optical materials: Expanding from borates to borate fluorides to fluorooxoborates[J]. Acc. Chem. Res., 2019, 52(3):  791-801. doi: 10.1021/acs.accounts.8b00649

    3. [3]

      Wu H P, Pan S L, Poeppelmeier K R, Li Y H, Jia D Z, Chen Z C, Fan X Y, Yang Y, Rondinelli J M, Luo S H. K3B6O10Cl: A new structure analogous to perovskite with a large second harmonic generation response and deep UV absorption edge[J]. J. Am. Chem. Soc., 2011, 133(20):  7786-7790. doi: 10.1021/ja111083x

    4. [4]

      Wu H P, Yu H W, Pan S L, Huang Z J, Yang Z H, Su X, Poeppelmeier K R. Cs2B4SiO9: A deep-ultraviolet nonlinear optical crystal. Angew[J]. Chem., Int. Ed., 2013, 52(12):  3406-3410. doi: 10.1002/anie.201209151

    5. [5]

      Zhao S G, Gong B F, Bai L, Xu X, Zhang S Q, Sun Z H, Lin Z S, Hong M C, Chen C T, Luo J H. Beryllium-free Li4Sr(BO3)2 for deep-ultraviolet nonlinear optical applications[J]. Nat. Commun., 2014, 5(1):  4019. doi: 10.1038/ncomms5019

    6. [6]

      Wu B, Tang D, Ye N, Chen C T. Linear and nonlinear optical properties of the KBe2BO3F2(KBBF) crystal[J]. Opt. Mater., 1996, 5(5/6):  105109.

    7. [7]

      Chen C T. Recent advances in deep and vacuum-UV harmonic generation with KBBF crystal[J]. Opt. Mater., 2004, 26(4):  425-429. doi: 10.1016/j.optmat.2004.02.007

    8. [8]

      Chen C T, Wang G L, Wang X Y, Zhu Y, Xu Z Y, Kanai T, Watanabe S. Improved Sellmeier equations and phase-matching characteristics in deep-ultraviolet region of KBe2BO3F2 crystal[J]. IEEE. J. Quantum Electron., 2008, 44(7):  617-621. doi: 10.1109/JQE.2008.920324

    9. [9]

      Chen C T, Wu B C, Jiang A D, You G M. A new-type ultraviolet SHG crystal β-BaB2O4[J]. Scientia Sinica Series B-Chemical Biological Agricultural Medical & Earth Science, 1985, 28(3):  235-243.

    10. [10]

      Chen C T, Wu Y C, Jiang A D, Wu B C, You G M, Li R K, Lin S J. New nonlinear-optical crystal: LiB3O5[J]. J. Opt. Soc. Am. B-Opt. Phys., 1989, 6(4):  616-621. doi: 10.1364/JOSAB.6.000616

    11. [11]

      Wu Y C, Sasaki T, Nakai N, Yokotani A H, Tang H G, Chen C T. CsB3O5: A new nonlinear optical crystal[J]. Appl. Phys. Lett., 1993, 62(21):  2614-2615. doi: 10.1063/1.109262

    12. [12]

      Ward J F, Franken P A. Structure of nonlinear optical phenomena in potassium dihydrogen phosphate[J]. Phys. Rev., 1964, 133(1A):  183-190. doi: 10.1103/PhysRev.133.A183

    13. [13]

      Lavrovskaya O I, Pavlova N I, Tarasov A V. Second harmonic generation of light from an AlG: Nd3+ laser in an optically biaxial crystal of KTiOPO4[J]. Sov. Phys. Crystallogr., 1986, 31:  1145-1151.

    14. [14]

      Okorogu A O, Mirov S B, Lee W, Crouthamel D I, Jenkins N, Dergachev A Y, Vodopyanov K L, Badikov V V. Tunable middle infrared down conversion in GaSe and AgGaS2[J]. Opt. Commun., 1998, 155(4/5/6):  307-312.

    15. [15]

      Guo S P, Chi Y, Guo G C. Recent achievements on middle and farinfrared second-order nonlinear optical materials[J]. Coord. Chem. Rev., 2017, 335:  44-57. doi: 10.1016/j.ccr.2016.12.013

    16. [16]

      Tran T T, Yu H W, Rondinelli J M, Poeppelmeier K R, Halasyamani P S. Deep ultraviolet nonlinear optical materials[J]. Chem. Mater., 2016, 28(15):  5238-5258. doi: 10.1021/acs.chemmater.6b02366

    17. [17]

      Cavalieri A L, Muller N, Uphues T, Yakovlev V X, Baltuska A, Horvath B, Schmidt B, Blumel L, Holzwarth R, Hendel S, Drescher M, Kleineberg U, Echenique P M, Kienberger R, Krausz F, Heinzmann U. Attosecond spectroscopy in condensed matter[J]. Nature, 2007, 449(7165):  1029-1032. doi: 10.1038/nature06229

    18. [18]

      Yao W J, He R, Wang X Y, Lin Z S, Chen C T. Analysis of deep-UV nonlinear optical borates: approaching the end[J]. Adv. Opt. Mater., 2014, 2(5):  411-417. doi: 10.1002/adom.201300535

    19. [19]

      Liu Y C, Shen Y G, Zhao S G, Luo J H. Structure-property relationship in nonlinear optical materials with π-conjugated CO3 triangles[J]. Coord. Chem. Rev., 2020, 407:  213152. doi: 10.1016/j.ccr.2019.213152

    20. [20]

      Liu X M, Gong P F, Yang Y, Song G M, Lin Z S. Nitrate nonlinear optical crystals: A survey on structure-performance relationships[J]. Coord. Chem. Rev., 2019, 400:  213045. doi: 10.1016/j.ccr.2019.213045

    21. [21]

      Shi G Q, Wang Y, Zhang F F, Zhang B B, Yang Z H, Hou X L, Pan S L, Poeppelmeier K R. Finding the next deep-ultraviolet nonlinear optical material: NH4B4O6F[J]. J. Am. Chem. Soc., 2017, 139(31):  10645-10648. doi: 10.1021/jacs.7b05943

    22. [22]

      Zhang Z Z, Wang Y, Zhang B B, Yang Z H, Pan S L. Polar fluoro-oxoborate, NaB4O6F: A promising material for ionic conduction and nonlinear optics[J]. Angew. Chem. Int. Ed., 2018, 57(22):  6577-6581. doi: 10.1002/anie.201803392

    23. [23]

      Tran T T, He J G, Rondinelli J M, Halasyamani P S. RbMgCO3F: A new deep-ultraviolet nonlinear optical material[J]. J. Am. Chem. Soc., 2015, 137(33):  10504-10507. doi: 10.1021/jacs.5b06519

    24. [24]

      Peng G, Lin C S, Ye N. NaZnCO3(OH): A high-performance carbonate ultraviolet nonlinear optical crystal derived from KBe2BO3F2[J]. J. Am. Chem. Soc., 2020, 142(49):  20542-20546. doi: 10.1021/jacs.0c09866

    25. [25]

      Huang L, Zou G H, Cai H Q, Wang S C, Lin C S, Ye N. Sr2(OH)3NO3: The first nitrate as a deep UV nonlinear optical material with large SHG responses[J]. J. Mater. Chem. C, 2015, 3(20):  5268-5274. doi: 10.1039/C5TC00344J

    26. [26]

      Zou G H, Lin C S, Kim G H, Jo H, Ok K M. Rb2Na (NO3)3: A congruently melting UV-NLO crystal with a very strong second-harmonic generation response[J]. Crystals, 2016, 6(4):  42. doi: 10.3390/cryst6040042

    27. [27]

      Yu P, Wu L M, Zhou L J, Chen L. Deep-ultraviolet nonlinear optical crystals: Ba3P3O10X (X=Cl, Br)[J]. J. Am. Chem. Soc., 2014, 136(1):  480-487. doi: 10.1021/ja411272y

    28. [28]

      Zhao S G, Gong P F, S , Luo S Y, Bai L, Lin Z S, Ji C M, Chen T L, Hong M C, Luo J H. Deep-ultraviolet transparent phosphates RbBa2(PO3)5 and Rb2Ba3(P2O7)2 show nonlinear optical activity from condensation of[PO4]3- units[J]. J. Am. Chem. Soc., 2014, 136(24):  8560-8563. doi: 10.1021/ja504319x

    29. [29]

      Li L, Wang Y, Lei B H, Han S J, Yang Z H, Poeppelmeier K R, Pan S L. A new deep-ultraviolet transparent orthophosphate LiCs2PO4 with large second harmonic generation response[J]. J. Am. Chem. Soc., 2016, 138(29):  9101-9104. doi: 10.1021/jacs.6b06053

    30. [30]

      Li Y Q, Liang F, Zhao S G, Li L N, Wu Z Y, Ding Q R, Liu S, Lin Z S, Hong M C, Luo J H. Two non-π-conjugated deep-UV nonlinear optical sulfates[J]. J. Am. Chem. Soc., 2019, 141(9):  3833-3837. doi: 10.1021/jacs.9b00138

    31. [31]

      Sha H Y, Xu J X, Xiong Z Y, Wang Z J, Su R B, He C, Yang X M, Long X F, Liu Y. An optimized KBe2BO3F2-Like structure: The unity of deep-ultraviolet transparency, nonlinear optical property, and ferroelectricity[J]. Adv. Opt. Mater., 2022, 10(15):  2200228. doi: 10.1002/adom.202200228

    32. [32]

      Wu C, Jiang C B, Wei G F, Jiang X X, Wang Z J, Lin Z S, Huang Z P, Humphrey M G, Zhang C. Toward large second-harmonic generation and deep-UV transparency in strongly electropositive transition metal sulfates[J]. J. Am. Chem. Soc., 2022, 145(5):  3040-3046.

    33. [33]

      Wang Y, Zhang B B, Yang Z H, Pan S L. Cation-tuned synthesis of fluorooxoborates: Towards optimal Deep-ultraviolet nonlinear optical materials[J]. Angew. Chem. Int. Ed., 2018, 130(8):  2150-2154.

    34. [34]

      Zhang B B, Han G P, Wang Y, Chen X L, Yang Z H, Pan S L. Expanding frontiers of ultraviolet nonlinear optical materials with fluorophosphates[J]. Chem. Mater., 2018, 30(15):  5397-5403. doi: 10.1021/acs.chemmater.8b02223

    35. [35]

      Wu M F, Feng J W, Xie W C, Tudi A, Chu D D, Lu J J, Pan S L, Yang Z H. From phosphate fluoride to fluorophosphate: Design of novel ultraviolet/deep-ultraviolet nonlinear optical materials for BePO3F with optical property enhancement[J]. ACS Appl. Mater. Interfaces, 2022, 14(34):  39081-39090. doi: 10.1021/acsami.2c12001

    36. [36]

      Ding Q R, Liu X M, Zhao S G, Wang Y S, Li Y Q, Li L N, Liu S, Lin Z S, Hong M C, Luo J H. Designing a deep-UV nonlinear optical fluorooxosilicophosphate[J]. J. Am. Chem. Soc., 2020, 142(14):  6472-6476. doi: 10.1021/jacs.0c00060

    37. [37]

      Han G P, Lei B H, Yang Z H, Wang Y, Pan S L. A fluorooxosilicophosphate with an unprecedented SiO2F4 species[J]. Angew. Chem. Int. Ed., 2018, 57(31):  9828-9832. doi: 10.1002/anie.201805759

    38. [38]

      Hao X, Luo M, Lin C S, Peng G, Xu F, Ye N. M (NH2SO3)2(M=Sr, Ba): Two deep-ultraviolet transparent sulfamates exhibiting strong second harmonic generation responses and moderate birefringence[J]. Angew. Chem. Int. Ed., 2021, 60(14):  7621-7625. doi: 10.1002/anie.202016372

    39. [39]

      Sheldrick G M. SHELXS-97: Program for the solution of crystal structures. University of Göttingen, Germany, 1997.

    40. [40]

      Johnson T J, Bernacki B E, Redding R L, Su Y F, Brauer C S, Myers T L, Stephan E G. Intensity-value corrections for integrating sphere measurements of solid samples measured behind glass[J]. Appl. Spectrosc., 2014, 68(11):  1224-1234. doi: 10.1366/13-07322

    41. [41]

      Kurtz S K, Perry T T. A powder technique for the evaluation of non-linear optical materials[J]. J. Appl. Phys., 1968, 39(8):  3798-3813. doi: 10.1063/1.1656857

    42. [42]

      Clark S J, Segall M, Pickard C J, Hasnip P J, Probert M, Refson K, Payne M C. First principles methods using CASTEP[J]. Z. Kristall., 2005, 220(5/6):  567-570.

    43. [43]

      Payne M C, Teter M P, Allan D C, Arias T A, Joannopoulos J D. Iterative minimization techniques for ab initio total-energy calculations: Molecular dynamics and conjugate gradients[J]. Rev. Mod. Phys., 1992, 64(4):  1045-1097. doi: 10.1103/RevModPhys.64.1045

    44. [44]

      Perdew J P, Burke K, Ernzerhof M. Generalized gradient approximation made simple[J]. Phys. Rev. Lett., 1998, 77(18):  3865-3868.

    45. [45]

      Perdew J P, Wang Y.. Pair-distribution function and its coupling-constant average for the spin-polarized electron gas[J]. Phys. Rev. B, 1992, 46(20):  12947-12954. doi: 10.1103/PhysRevB.46.12947

    46. [46]

      Hamann D R, Schlüter M, Chiang C. Norm-conserving pseudopoten-tials[J]. Phys. Rev. Lett., 1979, 43(20):  1494-1497. doi: 10.1103/PhysRevLett.43.1494

    47. [47]

      Monkhorst H J, Pack J D. Special points for Brillouin-zone integrations[J]. Phys. Rev. B, 1976, 13(12):  5188-5192. doi: 10.1103/PhysRevB.13.5188

    48. [48]

      Godby R W, Schlüter M, Sham L J. Trends in self-energy operators and their corresponding exchange-correlation potentials[J]. Phys. Rev. B, 1987, 36(12):  6497-6500. doi: 10.1103/PhysRevB.36.6497

    49. [49]

      Okoye C M I. Theoretical study of the electronic structure, chemical bonding and optical properties of KNbO3 in the paraelectric cubic phase[J]. J. Phys.-Condes. Matter, 2003, 15(35):  5945-5958. doi: 10.1088/0953-8984/15/35/304

  • Figure 1  Crystal structure of Li(NH2SO3): (a)[Li4O8S4] ring; (b) [Li4O8S4] layer parallel to the ac plane; (c) [Li3O6S3] ring; (d) [Li3O6S3] layer parallel to the ab plane; (e) 3D framework structure of Li(NH2SO3)

    Figure 2  (a) Arrangement of NH2SO3- units in Na(NH2SO3); (b) Coordination environment of Na atoms and the [Na3O11] unit; (c) Chain structure of [Na3O11] units; (d) 3D framework of Na(NH2SO3)

    Symmetry codes: #1: 0.5+x, 0.5-y, 1-z; #2:-0.5+x, 0.5-y, 1-z; #3: x, y, z; #4:-1+x, y, z; #5:-x, -0.5+y, 1.5-z.

    Figure 3  TG curves of (a) Li(NH2SO3) and (b) Na(NH2SO3)

    Figure 4  IR spectra of (a) Li(NH2SO3) and (b) Na(NH2SO3)

    Figure 5  UV-Vis-NIR diffuses reflectance spectra of (a) Li(NH2SO3) and (b) Na(NH2SO3)

    The inset shows the corresponding band gap.

    Figure 6  (a) Phase-matching curve of Li(NH2SO3) and Na(NH2SO3) with 1064 nm laser radiation; (b) Oscilloscope traces of the SHG signals for powders of Li(NH2SO3) and Na(NH2SO3) and KDP in the same particle size range of 105-150 µm

    KDP was used as a reference for the SHG measurement at 1 064 nm.

    Figure 7  Band structure diagram of (a) Li(NH2SO3) and (b) Na(NH2SO3); DOS and PDOS diagrams of (c) Li(NH2SO3) and (d) Na(NH2SO3)

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  265
  • HTML全文浏览量:  12
文章相关
  • 发布日期:  2024-07-10
  • 收稿日期:  2024-02-02
  • 修回日期:  2024-03-26
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章