Regulating the key performance parameters for Hg-based IR NLO chalcogenides via bandgap engineering strategy

A-Yang Wang Sheng-Hua Zhou Mao-Yin Ran Xin-Tao Wu Hua Lin Qi-Long Zhu

Citation:  A-Yang Wang, Sheng-Hua Zhou, Mao-Yin Ran, Xin-Tao Wu, Hua Lin, Qi-Long Zhu. Regulating the key performance parameters for Hg-based IR NLO chalcogenides via bandgap engineering strategy[J]. Chinese Chemical Letters, 2024, 35(10): 109377. doi: 10.1016/j.cclet.2023.109377 shu

Regulating the key performance parameters for Hg-based IR NLO chalcogenides via bandgap engineering strategy

English

  • Due to the unique capability of infrared (IR) nonlinear optical (NLO) materials to convert frequencies and produce different wavelengths based on intrinsic light, they are utilized as an essential component in solid-state laser devices within military and civilian fields [1-5]. So far, commercial IR-NLO crystals have been limited in their applications due to inherent defects that lead to low utilization efficiency. For instance, AgGaS2 [6] has a low laser-induced damage threshold (LIDT), AgGaSe2 [7] cannot achieve phase matching (PM) behavior and also has a small LIDT, and ZnGeP2 [8] exhibits severe two-photon absorption. The aforementioned limitations are primarily caused by the narrow band gaps (Eg) of these materials. In addition, there is an incompatible relationship between a wide Eg and a big second–harmonic-generation (SHG) coefficient (deff) [9-14]. Therefore, it is of great significance and importance to explore new materials that exhibit a well-balanced performance in terms of IR-NLO properties.

    In recent decades, extensive research has been conducted on transition metal-based chalcogenides, establishing them as leading candidates for IR-NLO applications [15-19]. Among these chalcogenides, Hg-based ones have gained significant popularity compared to their counterparts, such as Zn and Cd [20]. This popularity is attributed to their unique structural advantages, which include diverse coordination methods like linear [HgQ2], trigonal-planar [HgQ3], and tetrahedral [HgQ4]. The highly polarized Hg-Q bonds present in these materials are beneficial for achieving strong effective deff and significant refractive index variations (Δn). Moreover, Hg-based chalcogenides offer a wider range of IR transmission, covering two crucial atmospheric windows at 3–5 µm and 8–12 µm. So far, all reported Hg-based IR-NLO materials have focused on multivariate systems, such as the chalcogenide system of AHgPS4 [21] and AEHgMQ4 [22-25] (A = alkali-metal elements; AE = alkaline-earth-metal elements; M = group 14 elements; Q = chalcogen), as well as the salt-inclusion chalcogenide system of [Ba4Cl2][HgGa4S10] [26] and [AX][Hg3P2S8] (X = halogen) [27]. However, research on ternary systems is relatively scarce. Recently, the ternary transition-metal-rich M4MQ6 family system has caught our attention, particularly compound Hg4GeS6 [28,29]. Despite being reported as early as 1968, it has not yet undergone thorough investigation regarding its NLO performance [30]. Although theoretical calculations indicate a significant deff of approximately 50 pm/V, preliminary experimental characterization suggests that the SHG signal at of this material is relatively weak under 2050 nm frequency doubling light conditions. This weakness can primarily be attributed to the strong absorption of the both fundamental and frequency-doubled light, which is caused by the material's low optical Eg (ca. 1.3 eV). On the other hand, Hg4GeS6 possesses a dense 3D tetrahedral-stacking framework. However, due to the slight anisotropy of this structure, its birefringence index is exceedingly small (Δn < 0.01), rendering PM unattainable within the working band. Consequently, this limitation hinders its potential for further applications.

    In this work, we have implemented a bandgap engineering strategy that involves alkali metal introduction and Hg/Ge ratio regulation to address the limitations of Hg4GeS6 mentioned above. The rationale behind this approach is based on the following considerations. Firstly, the introduction of alkali metals with high electronegativity is helpful in obtaining higher Eg. Additionally, it helps in disrupting the original compact structure, leading to the formation of a lower-dimensional structure. This alteration in structure enhances its anisotropy, which in turn results in higher Δn values. Furthermore, by reducing the Hg/Ge ratio, i.e., increasing the Ge content in the component, which can also achieve higher Eg. After multiple attempts, we have successfully obtained two new non-centrosymmetric (NCS) quaternary Hg-based chalcogenides, namely Rb2HgGe3S8 and Cs2HgGe3S8, using the high-temperature solid-phase method. As anticipated, both compounds exhibit excellent IR-NLO performances including sufficient deff (0.55–0.70 × AgGaS2), wide Eg (3.27 and 3.41 eV), huge LIDTs (17.4–19.7 × AgGaS2), broad optical transmission intervals (0.32–17.5 µm), and suitable Δn (0.069–0.086@2050 nm). These findings indicate that these compounds hold great potential as candidates for IR-NLO applications. In this report, we will provide details on their synthesis, crystal structures, optical properties, and theoretical calculation analysis.

    The crystal structure of Hg4GeS6 is significantly complicated, consisting of 4 independent [HgS4] tetrahedra and 1 [GeS4] tetrahedra. The three-dimensional (3D) framework can be described as a combination of 3D Hg–Ge–S stacks and 2D Hg–S parts (Fig. 1a, and a more detailed structure is shown in Fig. S1 in Supporting information). In the 3D Hg–Ge–S parts, the [Hg2GeS7] layers are connected to each other by sharing S atoms, stacking in an ABABAB manner. Within each [Hg2GeS7] layer, interesting 12-membered rings (12-MRs) [Hg2GeS9] and 6-membered rings (6-MRs) [Hg2GeS7] are formed through vertex-sharing [Hg2S4], [Hg4S4], and [GeS4]. Each 12-MR is linked to four neighboring 12-MRs and four 6-MRs through Hg/Ge–S bonds. The 2D Hg–S parts are formed by stacking [Hg2S5] layers along the a-axis. Within these layers, zigzag 12-MRs [HgS3] are formed through vertex-sharing [Hg1S4] and [Hg3S4]. Unfortunately, due to the slight anisotropy of this dense 3D tetrahedral-stacking framework, the Δn index of Hg4GeS6 is exceedingly small (< 0.01). Therefore, achieving perfect PM feature within the working band is not possible. On the other hand, the high Hg/Ge ratio in the Hg4GeS6 results in a low optical Eg (ca. 1.3 eV). This low Eg presents a challenge in obtaining a high LIDT, thereby restricting its application in the high-energy laser band. In order to address the limitations of the ternary compound Hg4GeS6, we have successfully synthesized two quaternary NCS Hg-based chalcogenides, namely Rb2HgGe3S8 and Cs2HgGe3S8. This achievement was accomplished by employing a bandgap engineering approach, which included the introduction of alkali metals and regulation of the Hg/Ge ratio.

    Figure 1

    Figure 1.  Structure transformation from (a) 3D Hg4GeS6 to (b) 2D A2HgGe3S8; (c) A 2D [HgGe3S8]2– layer viewed along the bc plane with the coordination environment of (d) [HgGeS7] polyhedron and (e) [Ge2S6] dimer marked.

    The single X-ray diffraction data accurately correlates to the orthorhombic system [space group: P212121 (No. 19)]; Pearson symbol: oP60; idealized Wyckoff sequence: a14] for A2HgGe3S8 (A = Rb, Cs) is accurately correlated. Detailed crystallographic parameters for both compounds are listed in Table S1 (Supporting information). In each unit, there are two crystallographically independent A atoms, one Hg, three Ge, and eight S atoms, all occupying the 4a Wyckoff site (Table S2 in Supporting information). As shown in Fig. 1b, A2HgGe3S8 (A = Rb, Cs) compounds exhibit a 2D [HgGe3S8]2– layer, which is aligned along the c axis in the form of ABABAB fashion, with A+ cations inserted to maintain charge balance. Fig. 1c illustrates that each 2D [HgGe3S8]2– layer consists of 1D chains composed of [HgGeS6]6–. These 1D chains are linked by [HgGeS7] tetrahedra (Fig. 1d), alternating via corner-sharing along the a direction. Moreover, the neighboring chains are linked together into a layer through [Ge2S6]4– dimers (Fig. 1e), which share corners along the c direction (Fig. 1c). As listed in Tables S3 and S4 (Supporting information), the Hg and Ge atoms are in distorted tetrahedral coordination with 4 S atoms, and the bond lengths of Hg–S and Ge–S are in the ranges of 2.499(3)–2.554(3) Å and 2.156(3)–2.281(3) Å, respectively, which are comparable to those of reported Hg-based and Ge-based chalcogenides [31-38]. In addition, the A+ cations are coordinated by 8 and 11 S atoms to form [AS8] and [AS11] polyhedra respectively. The Rb–S bond distances range from 3.373 Å to 4.281 Å, while the Cs–S bond lengths range from 3.467 Å to 4.501 Å (Fig. S2 in Supporting information). These measurements are consistent with the bond distances observed in other alkali-metal sulfides [39-43].

    The high-yield bulk crystal A2HgGe3S8 (A = Rb, Cs) is synthesized by fully mixing of HgS, Ge, S and RbBr/CsBr at high temperature of 1073 K (experimental details can be found in Supporting information). The experimental powder XRD pattern of A2HgGe3S8 (A = Rb, Cs) matches the simulated pattern, indicating that the synthesis process described above can achieve a high phase purity of the crystals (Fig. S3 in Supporting information). SEM and EDX spectrum analysis were conducted, and the results showed that the compounds had a uniform distribution of A/Hg/Ge/S ratio, with an approximate ratio of 2:1:3:8, which is consistent with the theoretical element content ratio of the title compounds (Fig. S4 in Supporting information). As indicated by UV–vis-NIR absorption spectroscopy, the experimental Eg values are 3.41 and 3.27 eV for Rb2HgGe3S8 and Cs2HgGe3S8, respectively, which correspond to their crystal colours (the inset in Figs. 2a and b). Comparing the Eg values greater than 2.33 eV (Table S3 in Supporting information) of typical Hg-based IR-NLO chalcogenides in Fig. 2c, only Zn2HgP3S8 [44], AEHgGeS4 (AE = Sr, Ba) [25] and SrHgSiS4 [24] have been able to break through the Eg of 3.0 eV. The A2HgGe3S8 (A = Rb, Cs) not only surpassed the 3.0 eV Eg, but Rb2HgGe3S8 also possesses the largest Eg (3.41 eV) among all reported Hg-based IR-NLO materials. It is worth noting that in the A2MM3Q8 system (A = alkali-metal elements; M = divalent transition-metal elements, M = group 14 elements, Q = chalcogen) with space groups of P21, P21/c and P212121, there is a strong correlation between the experimental Eg (measured in eV) and the cell volume (measured in Å3) of the reported compounds (Fig. S5 in Supporting information). This correlation can be calculated using the formula Eg = –0.0041 × V + 9.5448 (with an R2 value of 0.878). In addition, the IR spectra (Fig. S6 in Supporting information) indicate that there are no noticeable absorption peaks in the range of 4000–600 cm–1, suggesting that A2HgGe3S8 (A = Rb, Cs) possess a wide IR transmission region. Specifically, Rb2HgGe3S8 has a transmission region of 17.2 µm, while Cs2HgGe3S8 has a transmission region of 17.5 µm. This enables coverage of two important atmospheric windows, which are the 3–5 µm and 8–12 µm ranges. As shown in Fig. S7 (Supporting information), the TG-DSC curves of A2HgGe3S8 (A = Rb, Cs) indicate no significant weight loss until reaching 800 K. However, as the temperature increases further, the title compounds undergo decomposition, which can be confirmed by powder XRD results (Fig. S8 in Supporting information).

    Figure 2

    Figure 2.  (a, b) UV–visible–NIR diffuse reflectance spectra of A2HgGe3S8 (A = Rb, Cs) based on the powder samples (inset: photographs of the crystals). (c) Comparison of reported IR-NLO Hg-based materials with Eg > 2.33 eV. (d) SHG intensity vs. particle size of compounds under 2050 nm laser irradiation.

    The powder SHG measurements of Rb2HgGe3S8 and Cs2HgGe3S8 were carried out under laser irradiation at 2050 nm using the modified Kurtz–Perry method [45]. It was observed that there is a positive correlation between the particle size and SHG intensity, indicating that both compounds exhibit type-Ⅰ PM behavior (Fig. 2d). In the particle size range of 150–210 µm, the SHG intensities of Rb2HgGe3S8 and Cs2HgGe3S8 are approximately 0.55 and 0.70 times that of the benchmark AgGaS2, respectively. To the best of our knowledge, both compounds are the first examples of NLO activity among all reported A2MM3Q8 systems with a space group of P212121, breaking the trade-off relationship between large Eg (> 3.0 eV) and sufficient deff (> 0.5 × AgGaS2) in Hg-based IR-NLO chalcogenides (Table S4 in Supporting information). The Eg is a major factor that affects the LIDT, and a wide Eg is beneficial for achieving high LIDT [46-49]. The measured power LIDTs [50] of Rb2HgGe3S8 and Cs2HgGe3S8 are 49.1 and 55.4 MW/cm2, respectively, which is approximately 17.4 times and 19.7 times that of AgGaS2 (2.82 MW/cm2) [51-55].

    To investigate the relationship between NCS structure and optical performance, we conducted a study on the electronic structure of the title compounds using first-principles computational methods [56,57]. The electron band structures reveal that the minimum value of the conduction band (CB) and the maximum value of the valence band (VB) occur at different points, indicating an indirect Eg for the product. The calculated Eg values are 2.16 eV and 2.19 eV, respectively (Fig. S9 in Supporting information) and the first Brillouin zone with high symmetry points was provided in Fig. S10 (Supporting information). However, it is important to note that due to the discontinuity of the exchange-related energy of the GGA functionals, the theoretical calculated Eg value will be lower than the experimentally measured one [58,59]. We also utilized the PBE approach to calculate the partial densities of states (PDOS) of A2HgGe3S8 (A = Rb, Cs) in the energy range from –10 eV to 10 eV. These results are displayed in Figs. 3a and b. According to the PDOS, it can be observed that the contribution of VB maximum is minimum S-3p and Hg-5d orbitals, with a small contribution from Hg-6p orbitals. Conversely, the contribution of CB minimum is mainly arises from Ge-4s and S-3p orbitals, with a small contribution from Ge-4p and Hg-6s orbitals. Based on these findings, we can deduce that the optical band gap primarily consists of Ge-4s, S-3p, Hg-5d orbitals. Therefore, the optical band gap of A2HgGe3S8 (A = Rb, Cs) is primarily determined by the functional units [GeS4] and [HgS4].

    Figure 3

    Figure 3.  Theoretical calculation results of A2HgGe3S8 (A = Rb, Cs): (a, b) PDOS; (c) frequency-dependent SHG coefficient d14; (d) theoretical birefringence (Δn); (e, f) the calculated shortest type-Ⅰ PM cut-off wavelength.

    Due to the restriction of Kleinman symmetry [60], A2HgGe3S8 (A = Rb, Cs) belongs to the orthogonal space group, with only one independent nonzero second-order nonlinear coefficient, i.e., d14 = d25 = d36. The calculated results for Rb2HgGe3S8 and Cs2HgGe3S8 are 0.721 pm/V and 0.684 pm/V at the wavelength of 2050 nm, respectively (Fig. 3c), which is consistent with the trend obtained from the experiment. Additionally, the theoretically calculated Δn values for Rb2HgGe3S8 and Cs2HgGe3S8 at 2050 nm are 0.086 and 0.069 (Fig. 3d), respectively. The moderate Δn index is conducive to PM, and the experimental results also confirmed that the title compounds exhibit PM behaviour. According to the condition of type-Ⅰ phase matching [ne(2ω) = no(ω)], where ne and no represent the refractive index of extraordinary light and ordinary light, respectively, it can be concluded that the shortest PM cut-off edges for Rb2HgGe3S8 and Cs2HgGe3S8 are 650 nm and 570 nm, respectively (Figs. 3e and f) [61-64].

    In order to further investigate the source of the SHG effect, we conducted a detailed analysis of the cut-off energy dependent static d14 based on the length specification [65,66], as shown in Fig. 4. The results indicate that d14 exhibits an upward trend in the VB-1 (−3.8~0.0 eV), CB-1 (2.1–3.0 eV), CB-3 (5.5–7.7 eV), and CB-5 (8.4–9.3 eV) regions. This implies that these four regions play a crucial role in the SHG response. By considering the charge density of PDOS and related components, we can conclude that the SHG response is a result of the synergistic effect between the NLO-activity [HgS4] and [GeS4] BBUs, specifically within the 2D [HgGe3S8]2– layer.

    Figure 4

    Figure 4.  Theoretical analysis of the SHG source: (a) Cut-off-energy-dependent variation of static coefficient d14; (b) charge density maps in the VB-1, CB-1, CB-3, and CB-5 regions along the bc plane.

    In conclusion, the introduction of alkali metals and regulation of the Hg/Ge ratio in the ternary narrow-band-gap parent Hg4GeS6 has led to the successful discovery and systematic characterization of two new Hg-based IR-NLO chalcogenides, namely Rb2HgGe3S8 and Cs2HgGe3S8. These compounds belong to the NCS P212121 space group and exhibit interesting 2D layered structures composed of [HgS4] and [GeS4] tetrahedra. They demonstrate excellent performance balance in terms of PM deff (0.55–0.70 × AaGgS2), Eg (3.27–3.41 eV), optical transparency window (0.32–17.5 µm), and LIDTs (17.4–19.7 × AaGgS2). Furthermore, detailed theoretical studies support the notion that tetrahedral [HgS4] and [GeS4] motifs play a pivotal role in the enhancing deff and Eg of these materials. This research expands our understanding of the structure-activity relationship of Hg-based chalcogenides and contributes to the design of high-performance IR-NLO candidates.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    This work was supported by the National Natural Science Foundation of China (Nos. 22175175 and 22193043), Natural Science Foundation of Fujian Province (Nos. 2022L3092 and 2023H0041), Fujian Science & Technology Innovation Laboratory for Optoelectronic Information of China (No. 2021ZR118), and the Youth Innovation Promotion Association CAS (No. 2022303). The authors thank Prof. Bing-Xuan Li at FJIRSM for helping with the NLO measurements and Prof. Yong-Fan Zhang at Fuzhou University for helping with the DFT calculations.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2023.109377.


    1. [1]

      F.J. Duarte, Chapters 2, 9 and 12, Tunable Laser Applications, CRC Press, Boca Raton, FL, 2008.

    2. [2]

      V. Petrov, Prog. Quantum Electron. 44 (2015) 1–106. doi: 10.1016/j.pquantelec.2015.08.002

    3. [3]

      V.A. Serebryakov, E.V. Boiko, N.N. Petrishchev, et al., J. Opt.Technol. 77 (2010) 6–17. doi: 10.1364/JOT.77.000006

    4. [4]

      X.T. Wu, L. Chen, Struct. Bonding 145 (2012) 1–42.

    5. [5]

      N.L.B. Sayson, T. Bi, V. Ng, et al., Nat. Photonics 13 (2019) 701–706. doi: 10.1038/s41566-019-0485-4

    6. [6]

      G.D. Boyd, E. Buehler, F.G. Storz, Appl. Phys. Lett. 18 (1971) 301–304. doi: 10.1063/1.1653673

    7. [7]

      G.C. Catella, L.R. Shiozawa, J.R. Hietanen, et al., Appl. Opt. 32 (1993) 3948–3951. doi: 10.1364/AO.32.003948

    8. [8]

      A. Harasaki, K.J. Kato, Appl. Phys. 36 (1997) 700–703. doi: 10.1143/JJAP.36.700

    9. [9]

      L. Kang, M.L. Zhou, J.Y. Yao, et al., J. Am. Chem. Soc. 137 (2015) 13049–13059. doi: 10.1021/jacs.5b07920

    10. [10]

      S.P. Guo, Y. Chi, G.C. Guo, Coord. Chem. Rev. 335 (2017) 44–57. doi: 10.1016/j.ccr.2016.12.013

    11. [11]

      K. Wu, S.L. Pan, Coord. Chem. Rev. 377 (2018) 191–208. doi: 10.1016/j.ccr.2018.09.002

    12. [12]

      H. Lin, W.B. Wei, H. Chen, et al., Coord. Chem. Rev. 406 (2020) 213150. doi: 10.1016/j.ccr.2019.213150

    13. [13]

      W.K. Wang, D.J. Mei, F. Liang, et al., Coord. Chem. Rev. 421 (2020) 213444. doi: 10.1016/j.ccr.2020.213444

    14. [14]

      M.Y. Ran, A.Y. Wang, W.B. Wei, et al., Coord. Chem. Rev. 481 (2023) 215059. doi: 10.1016/j.ccr.2023.215059

    15. [15]

      H. Chen, W.B. Wei, H. Lin, et al., Coord. Chem. Rev. 448 (2021) 214154. doi: 10.1016/j.ccr.2021.214154

    16. [16]

      H. Chen, M.Y. Ran, W.B. Wei, et al., Coord. Chem. Rev. 470 (2022) 214706. doi: 10.1016/j.ccr.2022.214706

    17. [17]

      W. Zhou, J. Wu, W. Liu, et al., Coord. Chem. Rev. 477 (2023) 214950. doi: 10.1016/j.ccr.2022.214950

    18. [18]

      F. Hou, D. M, Y. Zhang, et al., J. Alloys Compd. 904 (2022) 163944. doi: 10.1016/j.jallcom.2022.163944

    19. [19]

      M. Ma, J. Dang, Y. Wu, et al., Inorg. Chem. 62 (2023) 6549–6553. doi: 10.1021/acs.inorgchem.3c00683

    20. [20]

      C.X. Li, X.H. Meng, Z. Li, et al., Coord. Chem. Rev. 453 (2022) 214328. doi: 10.1016/j.ccr.2021.214328

    21. [21]

      W. Xing, C. Tang, P. Gong, et al., Inorg. Chem. 60 (2021) 18370–18378. doi: 10.1021/acs.inorgchem.1c02965

    22. [22]

      Y. Guo, F. Liang, Z. Li, et al., Inorg. Chem. 58 (2019) 10390–10398. doi: 10.1021/acs.inorgchem.9b01572

    23. [23]

      Y. Guo, F. Liang, W. Yin, et al., Chem. Mater. 31 (2019) 3034–3040. doi: 10.1021/acs.chemmater.9b01023

    24. [24]

      X.Y. Zhang, H.P. Wu, Z.G. Hu, et al., Adv. Optical Mater. 11 (2023) 2301735.

    25. [25]

      M.Y. Ran, S.H. Zhou, W.B. Wei, et al., Small 20 (2024) 2304563. doi: 10.1002/smll.202304563

    26. [26]

      Y. Zhang, H. Wu, Z. Hu, et al., Inorg. Chem. Front. 9 (2022) 4075–4080. doi: 10.1039/D2QI00937D

    27. [27]

      W. Xing, C. Tang, N. Wang, B. Kang, et al., Adv. Opt. Mater. 9 (2021) 2100563. doi: 10.1002/adom.202100563

    28. [28]

      M.Y. Li, B.X. Li, H. Lin, et al., Inorg. Chem. 57 (2018) 8730–8734. doi: 10.1021/acs.inorgchem.8b01682

    29. [29]

      J.D. Chen, C.S. Lin, S.D. Yang, et al., Cryst. Growth Des. 20 (2020) 2489–2496. doi: 10.1021/acs.cgd.9b01649

    30. [30]

      J. Serment, G. Perez, P. Hagenmuller, Bull. Soc. Chim. Fr. 2 (1968) 561–566.

    31. [31]

      K. Wu, Z. Yang, S. Pan, Chem. Mater. 28 (2016) 2795–2801. doi: 10.1021/acs.chemmater.6b00683

    32. [32]

      K. Wu, Z. Yang, S. Pan, Chem. Commun. 53 (2017) 3010–3013. doi: 10.1039/C6CC09565H

    33. [33]

      W. Xing, C. Tang, N. Wang, et al., Inorg. Chem. 59 (2020) 18452–18460. doi: 10.1021/acs.inorgchem.0c03176

    34. [34]

      Z. Yang, Y. Yang, Y. Guo, et al., Chem. Mater. 31 (2019) 1110–1117. doi: 10.1109/apap47170.2019.9224671

    35. [35]

      M.Y. Li, Z.J. Ma, B.X. Li, et al., Chem. Mater. 32 (2020) 4331–4339. doi: 10.1021/acs.chemmater.0c01258

    36. [36]

      M. Yan, Z.D. Sun, W.D. Yao, et al., Inorg. Chem. Front. 7 (2020) 2451–2458. doi: 10.1039/d0qi00266f

    37. [37]

      D. Mei, W. Cao, N. Wang, et al., Mater. Horiz. 8 (2021) 2330. doi: 10.1039/d1mh00562f

    38. [38]

      W. Wang, D. Mei, S. Wen, et al., Chin. Chem. Lett. 33 (2022) 2301–2315. doi: 10.1016/j.cclet.2021.11.089

    39. [39]

      H. Lin, L.J. Zhou, L. Chen, Chem. Mater. 24 (2012) 3406–3414. doi: 10.1021/cm301550a

    40. [40]

      H. Lin, H. Chen, Y.J. Zheng, et al., Chem. Eur. J. 23 (2017) 10407–10412. doi: 10.1002/chem.201701679

    41. [41]

      M.M. Chen, S.H. Zhou, W.B. Wei, et al., Adv. Optical Mater. 10 (2022) 2102123. doi: 10.1002/adom.202102123

    42. [42]

      M.M. Chen, S.H. Zhou, W.B. Wei, et al., ACS Mater. Lett. 4 (2022) 1264–1269. doi: 10.1021/acsmaterialslett.2c00409

    43. [43]

      H. Chen, M.Y. Ran, S.H. Zhou, et al., Chin. Chem. Lett. 34 (2023) 107838. doi: 10.1016/j.cclet.2022.107838

    44. [44]

      Y. Chu, H.S. Wang, T.D. Abutukadi, et al., Small 20 (2024) 2305074.

    45. [45]

      S.K. Kurtz, T.T. Perry, J. Appl. Phys. 39 (1968) 3798–3813. doi: 10.1063/1.1656857

    46. [46]

      M.Y. Ran, S.H. Zhou, W.B. Wei, et al., Small 19 (2023) 2300248. doi: 10.1002/smll.202300248

    47. [47]

      H.D. Yang, M.Y. Ran, W.B. Wei, et al., Mater. Today Phys. 35 (2023) 101127. doi: 10.1016/j.mtphys.2023.101127

    48. [48]

      M.Y. Ran, Z. Ma, H. Chen, et al., Chem. Mater. 32 (2020) 5890–5896. doi: 10.1021/acs.chemmater.0c02011

    49. [49]

      J.X. Zhang, M.Y. Ran, X.T. Wu, et al., Inorg. Chem. Front. 10 (2023) 5244–5257. doi: 10.1039/d3qi01144e

    50. [50]

      M.J. Zhang, X.M. Jiang, L.J. Zhou, et al., J. Mater. Chem. C 1 (2013) 4754–4760. doi: 10.1039/c3tc30808a

    51. [51]

      M.Y. Li, B.X. Li, H. Lin, et al., Chem. Mater. 31 (2019) 6268–6275. doi: 10.1021/acs.chemmater.9b02389

    52. [52]

      H. Chen, Y.Y. Li, B.X. Li, et al., Chem. Mater. 32 (2020) 8012–8019. doi: 10.1021/acs.chemmater.0c03008

    53. [53]

      H.D. Yang, M.Y. Ran, S.H. Zhou, et al., Chem. Sci. 13 (2022) 10725–10733. doi: 10.1039/d2sc03760b

    54. [54]

      Y.F. Shi, Z. Ma, B.X. Li, et al., Mater. Chem. Front. 6 (2022) 3054–3061. doi: 10.1039/d2qm00621a

    55. [55]

      M.Y. Ran, S.H. Zhou, B.X. Li, et al., Chem. Mater. 34 (2022) 3853–3861. doi: 10.1021/acs.chemmater.2c00385

    56. [56]

      S.N. Rashkeev, W.R.L. Lambrecht, B. Segall, Phys. Rev. B 57 (1998) 3905–3919.

    57. [57]

      Z. Fang, J. Lin, R. Liu, et al., CrystEngComm 16 (2014) 10569–10580. doi: 10.1039/C4CE01606H

    58. [58]

      K.J. Burke, Chem. Phys. 136 (2012) 150901.

    59. [59]

      K. Govaerts, R. Saniz, B. Partoens, et al., Phys. Rev. B: Condens. Matter Mater. Phys. 87 (2013) 235210. doi: 10.1103/PhysRevB.87.235210

    60. [60]

      D.A. Kleinman, Phys. Rev. 126 (1962) 1977. doi: 10.1103/PhysRev.126.1977

    61. [61]

      M.M. Chen, Z. Ma, B.X. Li, et al., J. Mater. Chem. C 9 (2021) 1156. doi: 10.1039/d0tc05952h

    62. [62]

      Y. Xiao, M.M. Chen, Y.Y. Shen, et al., Inorg. Chem. Front. 8 (2021) 2835. doi: 10.1039/d1qi00214g

    63. [63]

      C. Liu, S.H. Zhou, Y. Xiao, et al., J. Mater. Chem. C 9 (2021) 15407. doi: 10.1039/d1tc04498b

    64. [64]

      H. Lin, Y.Y. Li, M.Y. Li, et al., J. Mater. Chem. C 7 (2019) 4638–4643. doi: 10.1039/c9tc00647h

    65. [65]

      C. Aversa, J.E. Sipe, Phys. Rev. B 52 (1995) 14636. doi: 10.1103/PhysRevB.52.14636

    66. [66]

      S.N. Rashkeev, W.R.L. Lambrecht, B. Segall, Phys. Rev. B. 57 (1998) 3905.

  • Figure 1  Structure transformation from (a) 3D Hg4GeS6 to (b) 2D A2HgGe3S8; (c) A 2D [HgGe3S8]2– layer viewed along the bc plane with the coordination environment of (d) [HgGeS7] polyhedron and (e) [Ge2S6] dimer marked.

    Figure 2  (a, b) UV–visible–NIR diffuse reflectance spectra of A2HgGe3S8 (A = Rb, Cs) based on the powder samples (inset: photographs of the crystals). (c) Comparison of reported IR-NLO Hg-based materials with Eg > 2.33 eV. (d) SHG intensity vs. particle size of compounds under 2050 nm laser irradiation.

    Figure 3  Theoretical calculation results of A2HgGe3S8 (A = Rb, Cs): (a, b) PDOS; (c) frequency-dependent SHG coefficient d14; (d) theoretical birefringence (Δn); (e, f) the calculated shortest type-Ⅰ PM cut-off wavelength.

    Figure 4  Theoretical analysis of the SHG source: (a) Cut-off-energy-dependent variation of static coefficient d14; (b) charge density maps in the VB-1, CB-1, CB-3, and CB-5 regions along the bc plane.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  39
  • HTML全文浏览量:  5
文章相关
  • 发布日期:  2024-10-15
  • 收稿日期:  2023-11-09
  • 接受日期:  2023-12-06
  • 修回日期:  2023-12-04
  • 网络出版日期:  2023-12-12
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