Superatomic Ag58 nanoclusters incorporating a [MS4@Ag12]2+ (M = Mo or W) kernel show aggregation-induced emission

Jun-Jie Fang Zheng Liu Yun-Peng Xie Xing Lu

Citation:  Jun-Jie Fang, Zheng Liu, Yun-Peng Xie, Xing Lu. Superatomic Ag58 nanoclusters incorporating a [MS4@Ag12]2+ (M = Mo or W) kernel show aggregation-induced emission[J]. Chinese Chemical Letters, 2024, 35(10): 109345. doi: 10.1016/j.cclet.2023.109345 shu

Superatomic Ag58 nanoclusters incorporating a [MS4@Ag12]2+ (M = Mo or W) kernel show aggregation-induced emission

English

  • Atomically precise metal nanoclusters (NCs) have garnered significant attention in both fundamental and applied sciences, owing to their exceptional physical and chemical properties. These NCs have been extensively studied in areas such as catalysis [1], optics [2] and magnetism [3]. The precise atomic arrangement within these clusters offers a unique platform for investigating the intricate relationship between structure and properties [4]. However, achieving precise control over the properties and functionalities of these NCs by manipulating atoms and molecules within their crystal structures remains a significant challenge [5]. In recent years, template-assisted synthesis of NCs has emerged as a promising approach, primarily due to its potential for achieving structural predictability [68]. Unlike conventional surface-protecting ligands such as thiols [9,10], phosphines [11,12], alkynyls [13,14], and calixarenes [15,16], templates enable the aggregation of multiple metal cations, leading to the formation of cluster species with captivating structures and diverse properties [17]. This avenue opens up exciting opportunities for exploring novel NC structures and uncovering their associated properties.

    In previous studies, a diverse set of templates has been explored for the construction of high-nuclear silver clusters. This includes silver sulfide units [18], halogens [19,20], rare earth ions [21], acid anions [22,23], and oxometalates [24]. Among them, polyoxometalates, characterized by an abundance of terminal oxygen atoms and high charge densities, have emerged as particularly effective templates for facilitating the aggregation of additional [6,7]. In a significant contribution, Wang et al. demonstrated that the addition of PhCO2H triggers the transformation of [Mo8O28@Ag50] into [Mo6O22@Ag44] cluster [25]. Similarly, Shen et al. highlighted the role of anionic templates in tuning the photoluminescence (PL) properties of Ag33 clusters [24].

    Conversely, there has been mounting research interest in mixed-valence Ag(0)/Ag(Ⅰ) silver nanoclusters due to their intriguing structures and unique physicochemical properties. However, these clusters often suffer from poor stability attributed to the presence of partial Ag(0) characters [10,26,27]. To address this, improved synthetic approaches involving co-protection of various ligands, especially thiolates and phosphines, have proven effective in enhancing the stability of silver clusters such as Ag29 [12], Ag40 [28], Ag46 [29], Ag50 [30], Ag67 [11], Ag78 [31].

    Despite these advancements, there remains a dearth of investigations focusing on the incorporation of templates in mixed-valence Ag(0)/Ag(Ⅰ) clusters. Wang et al. observed that the formation of octahedral Ag64+ kernel and various molybdates can be disrupted by the presence of N,N-dimethylformamide (DMF) solvent [32]. Furthermore, Duan et al. identified the crucial roles played by oxometalates and phosphines in the formation of twenty-electron silver nanoclusters [33]. Additionally, Liu et al. reported the utilization of Ag6S unit as a template in the synthesis of Ag46 clusters [34]. Jin et al. uncovered the presence of Ag14S12 kernel in Ag62S12 clusters and Ag14S13 kernel in Ag62S13 clusters, leading to different numbers of valence electrons and subsequent variations in their PL properties [35,36]. Lastly, Wang et al. successfully synthesized Ag6Z4@Ag36 (Z = S or Se) clusters [37] through the controlled introduction of Ph3CSH or Ph3PSe, which gradually release S2− or Se2−.

    We have devised innovative strategies for the synthesis of a novel class of Ag NCs, utilizing NaBH4 as the reducing agent. In our synthetic approach, (NH4)2MS4 (M = Mo or W) serves a dual role as both a template and a slow-release reagent for S2− ions, while dppb acts as protective ligands, imparting stability to the NCs. Through meticulous isolation and characterization efforts, we have successfully identified and confirmed the formulas of four distinct silver nanoclusters: [(MoS4)@Ag12@Ag46S24(dppb)12] (1: C2/c space group; 3: I4 space group), [(WS4)@Ag12@Ag46S24(dppb)12] (2: C2/c space group; 4: I4 space group). In contrast to most other superatomic nanoclusters where the zero-valent Ag atoms occupy central positions, the distinctive feature of these eight-electron superatoms lies in the MS42− (M = Mo or W) template positioned at the center. The zero-valent Ag atoms of these Ag58 NCs are situated within the inner [MoS4@Ag12]2+ kernel. Notably, intermolecular C−H···π interactions between adjacent clusters can induce a space group transformation from C2/c to I4, further elucidating the structural intricacies. An intriguing observation is the aggregation-induced emission enhancement (AIEE) exhibited by these silver nanoclusters. By reducing the polarity of the solvent or introducing compounds containing phenyl groups, we achieve AIEE, amplifying their luminescent properties.

    The novel silver NCs 14 were successfully synthesized using a convenient one-pot method, and detailed experimental procedures can be found in Supporting information. Single-crystal X-ray diffraction (SCXRD) analysis revealed that the main structural motifs of 1 and 2 were analogous to those of NCs 3 and 4, albeit with different space group symmetry; specifically, 1 and 2 exhibited the C2/c space group, while 3 and 4 adopted the I4 space group. Comprehensive crystallographic data and structure refinement details for 14 are provided in Tables S1 and S2 (Supporting information).

    Crystals of 14 exhibit high solubility in MeOH and CH2Cl2, enabling a detailed investigation of their solution behaviors through electrospray ionization mass spectrometry (ESI-MS). In the positive ion mode ESI-MS analysis of 1, a distinct peak at m/z = 4128.50 corresponding to the three-charge state [1b]3+ was observed. This peak can be confidently attributed to the species [(MoS4)@Ag12@Ag46S24(dppb)11 + 3Ag+ + CH2Cl2 + CH3OH]3+ (sim. m/z = 4128.39). Furthermore, the peak at m/z = 3123.53, corresponding to the four-charge state [1a]4+, was identified as [(MoS4)@Ag12@Ag46S24(dppb)11 + 4Ag+ + CH2Cl2 + CH3OH]4+ (sim. m/z = 3123.52). Similarly, in the positive ion mode ESI-MS analysis of 2, prominent peaks at m/z = 3061.67 ([2a]4+), 4118.54 ([2b]3+), and 4129.19 ([2c]3+) were observed. These peaks correspond to the species [(WS4)@Ag12@Ag46S24(dppb)11 + 2Ag+ + 2H+]4+ (sim. m/z = 3062.84), [(WS4)@Ag12@Ag46S24(dppb)11 + 3Ag+]3+ (sim. m/z = 4118.41), and [(WS4)@Ag12@Ag46S24(dppb)11 + 3Ag+ + CH3OH]3+ (sim. m/z = 4129.09), respectively (Fig. 1 and Fig. S1 in Supporting information). Importantly, these species arise from the incorporation of three Ag+ ions and solvent molecules subsequent to the removal of a single dppb ligand from the pristine cluster. These assignments are consistent with the data obtained from single-crystal X-ray diffraction analysis. Other characterizations including elemental analysis (EA), powder XRD, thermogravimetric (TG) analysis, Fourier-transform infrared (FT-IR) spectra, and X-ray photoelectron spectroscopy (XPS) were analyzed in Figs. S2–S5 (Supporting information).

    Figure 1

    Figure 1.  Experimental isotope distributions of selected peaks for (a) 1 and (b) 2 (positive ion mode, dissolved in CH3OH and CH3CN).

    Single-crystal X-ray analysis revealed that NCs 14 are neutral silver nanoclusters consisting of 58 Ag atoms. They exhibit a tetrahedral structural motif, comprising a central MS42− (M = Mo or W) template, twenty-four bridging S2− ions, and twelve dppb ligands. The MS4 tetrahedron containing the Mo or W atom exhibits disorder, indicating significant vibrational motion within the tetrahedral framework at low temperatures. Although the structures of all four NCs are similar, a detailed description will be provided for 1 for clarity. In order to enhance visual understanding, Fig. 2 depicts a schematic representation of the crystal structure, with the average positions of the Mo atoms.

    Figure 2

    Figure 2.  (a) MoS42− anion template. (b) Cup-like [MoS4@Ag12]2+ kernel (MoS4 tetrahedra are highlighted in pink). (c) "Wind blade"-like Ag9S5 unit of Ag46S24 shell. (d) "Windmill" shaped Ag46S24 shell. (e) Multishelled core architecture of [MoS4@Ag12@Ag46S24] skeleton. (f) Overview of the total structure of [MoS4@Ag12@Ag46S24(dppb)12]. Color codes: Mo, brown; Ag, green and dark blue; S, orange and yellow; P, purple; C, gray; Hydrogen atoms are omitted for clarity.

    As depicted in Fig. 2, NC 1 features a multi-shell core architecture encapsulating a [MoS4@Ag12@Ag46S24(dppb)12] motif. Notably, the encapsulated MoS42− anion template is housed within an irregular Ag12 shell, distinct from the conventional M12 (M = Au or Ag) icosahedral structures reported in previous literature [27,38]. This observation highlights the substantial structural influence exerted by the introduction of templates on the core architecture. Each of the four S atoms from the template is coordinated to three Ag atoms, leading to the formation of the MoS4@Ag12 kernel, characterized by argentophilic Ag−Ag bond distances ranging from 2.870 Å to 3.217 Å. The observed Ag−SMo (SMo = sulfur atom of the MoS42− anion template) bond lengths and average Mo−S bond lengths are 2.552−2.920 and 2.530 Å, respectively. As depicted in Figs. 2d and e, the MoS4@Ag12 kernel is further surrounded by an Ag46S24 shell resembling a "windmill" shape. The Ag−Ag and Ag−S bond lengths between the MoS4@Ag12 kernel and Ag46S24 shell span from 2.8893 Å to 3.350 Å and from 2.427 Å to 2.870 Å, respectively. The "windmill"-shaped Ag46S24 shell consists of four "wind blade"-like Ag9S5 units, which are interconnected through two distinct bridging coordination modes (Fig. S6 in Supporting information). Each of the six S2− ions adopts a μ4 ligation mode, bridging four surrounding Ag atoms, while the remaining eighteen S2− ions coordinate to five Ag atoms through a μ5 ligation mode, with Ag−S distances ranging from 2.308 Å to 2.943 Å (Fig. S7 in Supporting information). Notably, argentophilic Ag−Ag interactions are observed, with Ag−Ag distances ranging from 2.992 Å to 3.266 Å.

    In the context of the multishelled core architecture [MoS4@Ag12@Ag46S24], it is observed that each of the twelve dppb ligands coordinates to an Ag atom on the Ag46S24 shell through Ag−P bonds, with bond lengths ranging from 2.393 Å to 2.457 Å. The bidentate phosphine ligands, characterized by differing lengths, exhibit distinct coordination modes on the surface of Ag clusters [30,39]. As illustrated in Fig. S8 (Supporting information), four distinct staple motifs involving dppb ligands, Ag atoms, and S atoms were identified in 1, namely Ph2P−Ag−Ag−Ag−PPh2, Ph2P−Ag−S−Ag−PPh2, Ph2P−Ag−Ag−S−Ag−PPh2, and Ph2P−Ag−S−Ag−S−Ag−PPh2.

    These specific coordination modes of dppb ligands contribute to the protection of the cluster surface, reduction of surface energy, and enhancement of stability. According to the "superatom" theory [40], the total electron count for the [(MS4)@Ag58S24(dppb)12] (M = Mo or W) cluster is calculated to be 8, which arises from [(58 × 1) − (24 × 2)−2] = 8. Notably, 8 corresponds to a magic number based on the electronic cluster shell model [41], following an Aufbau rule of 1S2|1P6|. This shell closure typically exhibits a significant energy gap between the highest occupied molecular orbitals (HOMO) and the lowest unoccupied molecular orbitals (LUMO), thereby contributing to the enhanced stability of ligand-protected metal clusters [27,33,38].

    The surfaces of NCs 14 are abundantly decorated with phenyl groups, giving rise to a diverse array of C−H···π intermolecular interactions between adjacent nanoclusters. In 1, these interactions are observed within a range of 2.996−4.033 Å, while in 2, they span from 2.929−3.975 Å. Similarly, in 3, the range is 2.947−4.086 Å, and in 4, it is 2.912−4.039 Å (Fig. 3a and Fig. S9 in Supporting information).

    Figure 3

    Figure 3.  (a) The intermolecular C−H···π interactions (Å) between dppb ligands of adjacent clusters in 1 (phenyl groups with interactions are highlighted in pink). (b, c) Cluster packing patterns along the b axis of 1 (C2/c) and 3 (I4). Color codes: Ag, dark blue; S, yellow; P, purple; C, gray; Hydrogen atoms are omitted for clarity.

    In an effort to synthesize 3 and 4, which possess the same cluster structure as 1 and 2 but belong to different space groups, PhCO2H, PhCO2K, PhPO3H2, or PhPO3Na2 were introduced to the synthesis system of 1 and 2. The successful synthesis of 3 and 4 with higher yields suggests that the addition of these compounds containing phenyl groups facilitates the aggregation and crystallization of the nanoclusters during the assembly process. This phenomenon can be attributed to the favorable π···π or C−H···π interactions between these compounds and the surface ligands of the nanoclusters, effectively restricting the mobility of the nanoclusters and enhancing their crystallization propensity. Furthermore, these interactions induce a transition in the cluster's space group from C2/c to I4 (Figs. 3b and c, and Fig. S10 in Supporting information).

    Notably, despite the similarities in structures among 14, there exist variations in average bond lengths attributable to distinct core structures and space groups. In 2, for instance, the W−S average bond length within the WS4 tetrahedra is measured at 2.593 Å, slightly longer than the Mo−S bond length (2.530 Å) observed in 1. As for 3, the average bond lengths of Ag−Ag and Ag−S are recorded as 2.913 Å and 2.448 Å, respectively. These values indicate a reduction compared to the corresponding bond lengths in 1, namely Ag−Ag (3.068 Å) and Ag−S (2.553 Å). Likewise, 4 exhibits shorter average bond lengths of Ag−Ag (3.044 Å) and Ag−S (2.469 Å) compared to 2, where the corresponding lengths are Ag−Ag (3.057 Å) and Ag−S (2.524 Å). These observations suggest that the transition from the C2/c to I4 space group leads to a contraction in average bond lengths. It is worth noting that differences in core structure and space group do not induce alterations in the coordination patterns of S2− ions and dppb ligands on the surface.

    The UV-vis absorption spectra of 14 exhibit two distinct peaks in the ranges of 475−520 and 600−650 nm (Figs. 4a and b). The presence of these similar absorption characteristic peaks indicates their comparable electronic structures [23]. Luminescence, a critical property of silver clusters, holds significant potential for applications in molecular sensing, detection, and imaging [4244]. Therefore, we conducted an extensive investigation into the emission properties of these Ag58 NCs in solution. When excited with 480 nm light at room temperature, 1 dissolved in methanol demonstrates a broad emission band centered at 625 nm, while 2 exhibits an emission peak at 630 nm (Figs. 4c and d, Figs. S11 and S12 in Supporting information). Although 1 and 2 display similar PL emission spectra, the emission peak of 2 is red-shifted by 5 nm compared to that of 1. This phenomenon can be attributed to the differential templates encapsulated within the identical silver-thiolate framework [45,46].

    Figure 4

    Figure 4.  (a, b) UV-vis spectra of 14. Emission spectra of (c) 1 and (d) 2 (dissolved in THF and MeOH, and the volume fraction of THF is 0%, 25%, 50%, and 75%, respectively; all excited by 480 nm light). (e, f) Emission spectra of 2 (added different quality of PPh3) and 2 (added different quality of PhCO2H).

    To gain further insight into the PL mechanism, we performed density functional theory (DFT) calculations employing the ORCA 5.0.3 program [47] and Multiwfn 3.8 program [48]. In these calculations, the phenyl groups present in the dppb ligands were selectively replaced with methyl groups to reduce computational costs. Fig. 5 and Fig. S13 (Supporting information) provide the results obtained, revealing HOMO−LUMO band gaps of 0.97 eV for 1 and 1.05 eV for 2, consistent with the value of 1.2 eV reported for the eight-electron Au25(SMe)18 cluster [49]. Comparative analysis further indicates that 1 and 2 exhibit similar arrangements of electron clouds in their Kohn-Sham orbitals, as well as contributions from different components to the LUMO and HOMO. The PL mechanism is attributed to the MS4 (M = Mo or W) → Ag charge transfer process, which is modified by the presence of outer S2− ions and dppb ligands within the crystal structure.

    Figure 5

    Figure 5.  (a, b) Composition analysis of frontier Kohn–Sham orbitals of 1 and 2. Electronic density diagrams of the (c) LUMO, (d) LUMO + 1, (e) HOMO, (f) HOMO-1 of 1 (isovalue = 0.01 a.u.). The phenyl groups present in the dppb ligands were selectively replaced with methyl groups, abbreviated as d(Me)(Me)b.

    The surfaces of Ag58 nanoclusters are adorned with numerous phenyl groups, establishing a foundation for exploring their AIE activities. To examine these behaviors, the nanoclusters were investigated in a MeOH/THF (tetrahydrofuran) solvent system. In each experiment, 5 mg of freshly prepared crystals were dissolved in 10 mL of solvent. By modulating the polarity of the mixed solvent system, the aggregation degree could be controlled. Specifically, increasing the volume fraction of THF (f = volTHF/volMeOH + THF) led to decreased polarity and a higher degree of aggregation. The observed results, depicted in Figs. 4c and d, were achieved through deliberate manipulation of these variables. With an increase in f, the PL intensities of 1 and 2 exhibited corresponding increments, accompanied by blue shifts of the emission peaks by 3−5 nm, suggesting a progressive enlargement of the assembly size under the influence of the solvent [50].

    Furthermore, the addition of PPh3 or PhCO2H to MeOH solvent of 2 yielded similar results of AIEE (Figs. 4e and f). For each experiment, 5 mg of freshly prepared crystals were dissolved in 10 mL of MeOH, followed by thorough stirring upon addition of PPh3 or PhCO2H. The phenyl groups present in PPh3 and PhCO2H facilitated abundant π··π and C−H···π intermolecular interactions with the numerous dppb ligands on the cluster's surface. Consequently, these interactions induced aggregation of the Ag58 clusters in the solvent, leading to enhanced PL intensity. It can be inferred that the C−H···π intermolecular interactions between adjacent molecules effectively curtail intramolecular rotations and vibrations. This constraint significantly enhances the radiative transition of the aggregated state, ultimately culminating in the observed AIE behavior. Overall, this AIE property emerges from the restricted intramolecular rotation (RIR) mechanism [51].

    In summary, we have successfully synthesized and characterized four novel silver thiolate nanoclusters, employing template-assisted strategies. These strategies involve the utilization of (NH4)2MoS4 or (NH4)2WS4 as both templates and slow chalcogen-release reagents. The resulting nanoclusters exhibit an eight-electron superatom nature and possess a multishelled core architecture denoted as [MS4@Ag12@Ag46S24(dppb)12] (M = Mo or W). Our study demonstrates the achievement of the AIEE through intermolecular C−H···π interactions present on the surfaces of nanoclusters 14. This effect can be realized by reducing the polarity of the solvent or introducing phenyl-containing compounds. Importantly, we highlight the significant role played by templates and dppb ligands in directing the construction of silver thiolate clusters with distinct optical properties. Furthermore, this work paves the way for exploring the intricate relationship between cluster structures and their associated properties, thereby opening new avenues for future investigations in this field.

    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.

    We gratefully acknowledge financial support from the National Natural Science Foundation of China (Nos. 21771071, 22171094, 21925104, and 92261204) and the Hubei Provincial Natural Science Foundation of China (No. 2021CFA020). We thank the staff in the Analytical and Testing Center at the Huazhong University of Science and Technology for all related measurements. We thank the staff at the BL17B beamline of the National Center for Protein Sciences Shanghai (NCPSS) at Shanghai Synchrotron Radiation Facility for the assistance with data collection.

    The X-ray crystallographic data for 14 have been deposited at the Cambridge Crystallographic Data Centre (CCDC), under deposition number 2193459, 2193586, 2193591, and 2193473.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2023.109345.


    1. [1]

      S.S. Wang, G.Y. Yang, Chem. Rev. 115 (2015) 4893–4962. doi: 10.1021/cr500390v

    2. [2]

      P. Yuan, R. Zhang, E. Selenius, et al., Nat. Commun. 11 (2020) 2229. doi: 10.1038/s41467-020-16062-6

    3. [3]

      S. Bao, L. Zheng, Coord. Chem. Rev. 319 (2016) 63–85. doi: 10.1016/j.ccr.2016.05.002

    4. [4]

      D.A. Tomalia, S.N. Khanna, Chem. Rev. 116 (2016) 2705–2774. doi: 10.1021/acs.chemrev.5b00367

    5. [5]

      H. Liu, C. Song, R. Huang, et al., Angew. Chem. Int. Ed. 55 (2016) 3699–3703. doi: 10.1002/anie.201511765

    6. [6]

      Y. Xie, T.C.W. Mak, J. Am. Chem. Soc. 133 (2011) 3760–3763. doi: 10.1021/ja1112035

    7. [7]

      Y. Xie, T.C.W. Mak, Angew. Chem. Int. Ed. 51 (2012) 8783–8786. doi: 10.1002/anie.201202195

    8. [8]

      Z. Wang, H. Su, Y. Gong, et al., Nat. Commun. 11 (2020) 308. doi: 10.1038/s41467-019-13682-5

    9. [9]

      C.P. Joshi, M.S. Bootharaju, M.J. Alhilaly, et al., J. Am. Chem. Soc. 137 (2015) 11578–11581. doi: 10.1021/jacs.5b07088

    10. [10]

      R.S. Dhayal, J. Liao, Y. Liu, et al., Angew. Chem. Int. Ed. 54 (2015) 3702–3706. doi: 10.1002/anie.201410332

    11. [11]

      M.J. Alhilaly, M.S. Bootharaju, C.P. Joshi, et al., J. Am. Chem. Soc. 138 (2016) 14727–14732. doi: 10.1021/jacs.6b09007

    12. [12]

      L.G. AbdulHalim, M.S. Bootharaju, Q. Tang, et al., J. Am. Chem. Soc. 137 (2015) 11970–11975. doi: 10.1021/jacs.5b04547

    13. [13]

      G. Duan, L. Tian, J. Wen, et al., Nanoscale 10 (2018) 18915–18919. doi: 10.1039/C8NR06399K

    14. [14]

      M. Qu, H. Li, L. Xie, et al., J. Am. Chem. Soc. 139 (2017) 12346–12349. doi: 10.1021/jacs.7b05243

    15. [15]

      D. Geng, X. Han, Y. Bi, et al., Chem. Sci. 9 (2018) 8535–8541. doi: 10.1039/C8SC03193B

    16. [16]

      Y. Bi, S. Du, W. Liao, Coord. Chem. Rev. 276 (2014) 61–72. doi: 10.1016/j.ccr.2014.06.011

    17. [17]

      J. Liu, H. Su, Z. Wang, et al., Chem. Commun. 54 (2018) 4461–4464. doi: 10.1039/C8CC01767K

    18. [18]

      Y. Xie, J. Jin, X. Lu, et al., Angew. Chem. Int. Ed. 54 (2015) 15176–15180. doi: 10.1002/anie.201507512

    19. [19]

      Z. Qin, D. Zhao, L. Zhao, et al., Nanoscale Adv. 1 (2019) 2529–2536. doi: 10.1039/C9NA00234K

    20. [20]

      L. Ren, P. Yuan, H. Su, et al., J. Am. Chem. Soc. 139 (2017) 13288–13291. doi: 10.1021/jacs.7b07926

    21. [21]

      G. Duan, Y. Xie, J. Jin, et al., Chem. Eur. J. 24 (2018) 6762–6768. doi: 10.1002/chem.201705906

    22. [22]

      M. Zhao, S. Huang, Q. Fu, et al., Angew. Chem. Int. Ed. 59 (2020) 20031–20036. doi: 10.1002/anie.202007122

    23. [23]

      J.S. Yang, Z. Han, X.Y. Dong, et al., Angew. Chem. Int. Ed. 59 (2020) 11898–11902. doi: 10.1002/anie.202004268

    24. [24]

      Y.L. Shen, J.L. Jin, G.X. Duan, et al., Chem. Eur. J. 27 (2021) 1122–1126. doi: 10.1002/chem.202004055

    25. [25]

      Z. Wang, H. Su, C. Tung, et al., Nat. Commun. 9 (2018) 4407. doi: 10.1038/s41467-018-06755-4

    26. [26]

      H. Yang, Y. Wang, X. Chen, et al., Nat. Commun. 7 (2016) 12809. doi: 10.1038/ncomms12809

    27. [27]

      A. Desireddy, B.E. Conn, J. Guo, et al., Nature 501 (2013) 399–402. doi: 10.1038/nature12523

    28. [28]

      X. Yuan, C. Sun, X. Li, et al., J. Am. Chem. Soc. 141 (2019) 11905–11911. doi: 10.1021/jacs.9b03009

    29. [29]

      M. Bodiuzzaman, A. Ghosh, K.S. Sugi, et al., Angew. Chem. Int. Ed. 58 (2019) 189–194. doi: 10.1002/anie.201809469

    30. [30]

      W. Du, S. Jin, L. Xiong, et al., J. Am. Chem. Soc. 139 (2017) 1618–1624. doi: 10.1021/jacs.6b11681

    31. [31]

      H. Yang, J. Yan, Y. Wang, et al., J. Am. Chem. Soc. 139 (2017) 16113–16116. doi: 10.1021/jacs.7b10448

    32. [32]

      Z. Wang, H. Su, M. Kurmoo, et al., Nat. Commun. 9 (2018) 2094. doi: 10.1038/s41467-018-04499-9

    33. [33]

      G. Duan, J. Han, B. Yang, et al., Nanoscale 12 (2020) 1617–1622. doi: 10.1039/C9NR07779K

    34. [34]

      X. Liu, J. Chen, J. Yuan, et al., Angew. Chem. Int. Ed. 57 (2018) 11273–11277. doi: 10.1002/anie.201805594

    35. [35]

      S. Jin, S. Wang, L. Xiong, et al., Chem. Mater. 28 (2016) 7905–7911. doi: 10.1021/acs.chemmater.6b03472

    36. [36]

      S. Jin, S. Wang, Y. Song, et al., J. Am. Chem. Soc. 136 (2014) 15559–15565. doi: 10.1021/ja506773d

    37. [37]

      Z. Wang, J. Liu, H. Su, et al., J. Am. Chem. Soc 141 (2019) 17884–17890. doi: 10.1021/jacs.9b09460

    38. [38]

      H. Yang, Y. Wang, H. Huang, et al., Nat. Commun. 4 (2013) 2422. doi: 10.1038/ncomms3422

    39. [39]

      F. Tian, R. Chen, J. Am. Chem. Soc. 141 (2019) 7107–7114. doi: 10.1021/jacs.9b02162

    40. [40]

      M. Walter, J. Akola, O. Lopez-Acevedo, et al., Proc. Natl. Acad. Sci. U. S. A. 105 (2008) 9157–9162. doi: 10.1073/pnas.0801001105

    41. [41]

      B. Yoon, P. Koskinen, B. Huber, et al., ChemPhysChem 8 (2007) 157–161. doi: 10.1002/cphc.200600524

    42. [42]

      T. Chen, S. Yang, J. Chai, et al., Sci. Adv. 3 (2017) e1700956. doi: 10.1126/sciadv.1700956

    43. [43]

      N. Goswami, Q. Yao, Z. Luo, et al., J. Phys. Chem. Lett. 7 (2016) 962–975. doi: 10.1021/acs.jpclett.5b02765

    44. [44]

      Y. Xie, Y. Shen, G. Duan, et al., Mater. Chem. Front. 4 (2020) 2015–2222.

    45. [45]

      X. Kang, M. Zhu, Chem. Soc. Rev. 48 (2019) 2422–2457. doi: 10.1039/C8CS00800K

    46. [46]

      Z. Chen, N. Zhao, Y. Fan, et al., Coord. Chem. Rev. 253 (2009) 1–20. doi: 10.1016/j.ccr.2007.11.015

    47. [47]

      F. Neese, F. Wennmohs, U. Becker, et al., J. Chem. Phys. 152 (2020) 224108. doi: 10.1063/5.0004608

    48. [48]

      T. Lu, F. Chen, J. Comput. Chem. 33 (2012) 580–592. doi: 10.1002/jcc.22885

    49. [49]

      J. Akola, M. Walter, R.L. Whetten, et al., J. Am. Chem. Soc. 130 (2008) 3756–3757. doi: 10.1021/ja800594p

    50. [50]

      Y. Zhang, D. Li, Y. Li, et al., Chem. Sci. 5 (2014) 2710. doi: 10.1039/c4sc00721b

    51. [51]

      Q. Zhang, D. Li, X. Li, et al., J. Am. Chem. Soc. 138 (2016) 13541–13550. doi: 10.1021/jacs.6b04776

  • Figure 1  Experimental isotope distributions of selected peaks for (a) 1 and (b) 2 (positive ion mode, dissolved in CH3OH and CH3CN).

    Figure 2  (a) MoS42− anion template. (b) Cup-like [MoS4@Ag12]2+ kernel (MoS4 tetrahedra are highlighted in pink). (c) "Wind blade"-like Ag9S5 unit of Ag46S24 shell. (d) "Windmill" shaped Ag46S24 shell. (e) Multishelled core architecture of [MoS4@Ag12@Ag46S24] skeleton. (f) Overview of the total structure of [MoS4@Ag12@Ag46S24(dppb)12]. Color codes: Mo, brown; Ag, green and dark blue; S, orange and yellow; P, purple; C, gray; Hydrogen atoms are omitted for clarity.

    Figure 3  (a) The intermolecular C−H···π interactions (Å) between dppb ligands of adjacent clusters in 1 (phenyl groups with interactions are highlighted in pink). (b, c) Cluster packing patterns along the b axis of 1 (C2/c) and 3 (I4). Color codes: Ag, dark blue; S, yellow; P, purple; C, gray; Hydrogen atoms are omitted for clarity.

    Figure 4  (a, b) UV-vis spectra of 14. Emission spectra of (c) 1 and (d) 2 (dissolved in THF and MeOH, and the volume fraction of THF is 0%, 25%, 50%, and 75%, respectively; all excited by 480 nm light). (e, f) Emission spectra of 2 (added different quality of PPh3) and 2 (added different quality of PhCO2H).

    Figure 5  (a, b) Composition analysis of frontier Kohn–Sham orbitals of 1 and 2. Electronic density diagrams of the (c) LUMO, (d) LUMO + 1, (e) HOMO, (f) HOMO-1 of 1 (isovalue = 0.01 a.u.). The phenyl groups present in the dppb ligands were selectively replaced with methyl groups, abbreviated as d(Me)(Me)b.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  47
  • HTML全文浏览量:  6
文章相关
  • 发布日期:  2024-10-15
  • 收稿日期:  2023-10-16
  • 接受日期:  2023-11-27
  • 修回日期:  2023-11-14
  • 网络出版日期:  2023-12-02
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