Artigo Acesso aberto Revisado por pares

Breakdown of Valence Shell Electron Pair Repulsion Theory in an H-Bond-Stabilized Linear sp-Hybridized Sulfur

2020; Chinese Chemical Society; Volume: 3; Issue: 10 Linguagem: Inglês

10.31635/ccschem.020.202000471

ISSN

2096-5745

Autores

Jin Wu, Bo Jin, Xiang Wang, Yayun Ding, Xiaoli Wang, Dandan Tang, Xiaohong Li, Jie Shu, Dong‐Sheng Li, Qipu Lin, Yan‐Bo Wu, Tao Wu,

Tópico(s)

Ammonia Synthesis and Nitrogen Reduction

Resumo

Open AccessCCS ChemistryRESEARCH ARTICLE1 Oct 2021Breakdown of Valence Shell Electron Pair Repulsion Theory in an H-Bond-Stabilized Linear sp-Hybridized Sulfur Jin Wu†, Bo Jin†, Xiang Wang, Yayun Ding, Xiao-Li Wang, Dandan Tang, Xiaohong Li, Jie Shu, Dong-Sheng Li, Qipu Lin, Yan-Bo Wu and Tao Wu Jin Wu† College of Chemistry, Chemical Engineering and Material Sciences, Soochow University, Suzhou 215123 , Bo Jin† The Key Laboratory of the Materials for Energy Storage and Conversion of Shanxi Province, Institute of Molecular Science, Shanxi University, Taiyuan 030006 , Xiang Wang College of Chemistry, Chemical Engineering and Material Sciences, Soochow University, Suzhou 215123 , Yayun Ding College of Chemistry, Chemical Engineering and Material Sciences, Soochow University, Suzhou 215123 , Xiao-Li Wang College of Chemistry, Chemical Engineering and Material Sciences, Soochow University, Suzhou 215123 , Dandan Tang College of Chemistry, Chemical Engineering and Material Sciences, Soochow University, Suzhou 215123 , Xiaohong Li College of Chemistry, Chemical Engineering and Material Sciences, Soochow University, Suzhou 215123 , Jie Shu Testing and Analysis Centre, Soochow University, Suzhou 215123 , Dong-Sheng Li College of Materials and Chemical Engineering, Hubei Provincial Collaborative Innovation Center for New Energy Microgrid, Key Laboratory of Inorganic Nonmetallic Crystalline and Energy Conversion Materials, China Three Gorges University, Yichang 443002 , Qipu Lin State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002 , Yan-Bo Wu The Key Laboratory of the Materials for Energy Storage and Conversion of Shanxi Province, Institute of Molecular Science, Shanxi University, Taiyuan 030006 and Tao Wu *Corresponding author: E-mail Address: [email protected] College of Chemistry, Chemical Engineering and Material Sciences, Soochow University, Suzhou 215123 https://doi.org/10.31635/ccschem.020.202000471 SectionsSupplemental MaterialAboutAbstractPDF ToolsAdd to favoritesDownload CitationsTrack Citations ShareFacebookTwitterLinked InEmail Exploring the unusual orbital hybridization types of atoms and their new connection modes contributes to the development of chemical bond theory and can inspire compounds with unique molecular configurations. Dicoordinated sulfur (S) atoms (or anions) with sp3 hybridization in a bent-bridging mode are commonly observed in many inorganic and organic compounds. However, sp-hybridized S species have, thus far, been extremely rare, and the linearly bridging mode has only been "forcibly" achieved with the aid of metal–S multiple bonds and/or significant steric hindrance from the surrounding organic ligands. In this study, we showed for the first time that X-ray crystallography could be employed to observe an sp-hybridized, linearly coordinated S, unprecedentedly stabilized only by hydrogen-bonding interactions in an indium (In)–S dimeric cluster. The H-bond-assisted stabilizing mechanism was confirmed by 1H NMR characterization. The analyses based on quantum chemistry calculations revealed that the concerned S atom possessed two valence lone pairs and formed two single bonds with In atoms. However, such an S atom adopted a linear rather than the bent bonding geometry, violating the famous valence shell electron pair repulsion (VSEPR) theory. These results may contribute to opening of a new branch of chemistry concerning weak-interaction-determined unique hybridization of nonmetallic atoms. Download figure Download PowerPoint Introduction Atomic orbital hybridization types and their connection modes have always been topics of interest in the field of theoretical chemistry.1–5 The well-established valence shell electron pair repulsion (VSEPR) theory is constructive in predicting the molecular geometry. According to this theory, the valence electron pairs surrounding an atom and including σ-bonding pairs and valence lone pairs tend to repel each other, and thus, will adopt an arrangement that minimizes the repulsion. This will, in turn, decrease the bond energy, increase stability, and finally determine the molecular geometry. In theoretical chemistry, atomic orbital hybridization and their connection modes are closely related to VSEPR theory, in that the orbital hybridization of the central atom in a molecular compound is related to the configuration of its outer-shell valence electrons and has a significant effect on its mode of connection to the surrounding atoms, which accordingly dominates the final structure and properties of the compound. Specifically, an atom with two, three, and four valence electron pairs will generally adopt the linear sp, planar trigonal sp2, and spatial tetrahedral sp3 hybridization, respectively. As is well-known, a sulfur (S) atom can take various hybridization configurations (sp2, sp3, sp3d, and sp3d2) in different molecular structures, and each hybridization corresponds to different connection modes of the central atom ( Supporting Information Table S1). Among the various configurations, the most common type is sp3-hybridized S, which usually adopts the dicoordinated bent (or angular) connection mode ( B-μ2-S) with a bond angle that is adjustable within a specific range (Figure 1a). Notably, Mealli et al.6 created a dicoordinated linearly bridging sulfide anion ( L-μ2-S) in a Ni–S complex in 1975. As expected, the subsequent theoretical calculations demonstrated that the sulfide anion in this rare connection mode adopts sp hybridization and forms d–p multiple bonds between the S2− and Ni2+ ions (Figure 1b).7 After this work, other M–S complexes with similar structural features were reported successively, most of which involved transition metal ions (such as V, U, Mo, W, Co, Fe, and Cr; see Supporting Information Table S2 and Figure S1), facilitating the formation of metal–S multiple bonds. However, though such type of L-μ2-S is rare, it is not too curious because such S atoms are involved in bonding with the σ-donating and π-accepting transition metals, where the lone pairs of S participate in the π-back-bonding and are no longer counted as valence electron pairs according to the VSEPR theory, so the total number of valence electron pairs are two, representing the classical sp hybridization. In contrast, there have been only two cases in which the central L-μ2-S anion coordinates to the metal ions via a metal–S single bond (M = Cu2+ or Pb4+) (Figure 1c).8–22 These results broke through the traditional notion that sulfide anions only adopt the B-μ2-S mode. Nevertheless, note that the space volume of organic ligands around the metal ions in these M–S complexes with the L-μ2-S mode is typically quite large. Therefore, it can be assumed that the sulfide anions are all "forced" to adopt a linearly bridging mode because the bent-bridging mode might lead to significant steric hindrance between adjacent ligands, which would cause an unstable molecular system with high total energy. Previous studies showed that large ligands' assistance with significant steric hindrance and the participation of multiple bonds involving transition metals are the two essential factors in forcing sulfide anions into the sp-hybridized linearly bridging mode. Nonetheless, in this work, we created the first sp-hybridized L-μ2-S anion possessing two S–In σ-bonding pairs and two valence lone pairs, which violated the famous VSEPR rule. This structure is unprecedented in that the linear S was not stabilized by the metal–S multiple bonds or significant steric hindrance from the ligand, but by the three N–H⋯S hydrogen bonds that form between the two lone pairs on the central sulfide anion and three organic imine molecules (Figure 1d). Figure 1 | Schematic presentations of different connection modes of a central S2− anion in M–S complexes. (a) Classical B-μ2-S with sp3 hybridization. (b) L-μ2-S with sp hybridization, stabilized by both metal–S multiple bonds and large steric hindrance from ligands bonded to transition metal ions. (c) L-μ2-S with sp hybridization only stabilized by a large steric hindrance from ligands bonded to the main group metal ions. (d) L-μ2-S with sp hybridization only stabilized by H-bonds (this work). Notes: gray areas represent the space volume of ligands bonded to M; the metal–S multiple bonds (rb) are much shorter than the single bonds (ra, rc, rd) ( Supporting Information Table S2); L in H–N(L) represents an organic imine molecule. Download figure Download PowerPoint Experimental Methods All chemicals were analytical grade and purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Indium nitrate hydrate [In(NO3)3·4.5H2O, 99.5%], indium powder (In, 99.99%), thiourea (CH4N2S, 99%), S (99.5%), acetonitrile (CH3CN, 99.5%), (±)-2-amino-1-butanol (2-AB, 98%), 1,5-diazabicyclo[4.3.0]non-5-ene (DBN, 98%), 2,6-dimethylpyridine (2,6-DMP, 98%), 2,6-dihydroxybenzoic acid (2,6-DOC, 98%), and 3-amino-1-propanol (3-AP, 98%) were used without further purification. Synthesis of 1 ({[(In20S33)(DBN)6]·(DBN-H+)6}) Route A In (114 mg, 1.00 mmol), S (96 mg, 3.00 mmol), DBN (2.00 mL, 16.18 mmol), 2,6-DMP (2.00 mL, 17.17 mmol), and 2-AB (2.00 mL, 21.57 mmol) were mixed in a 25 mL Teflon lining stainless steel and stirred for 30 min, then heated to 180 ℃ for 16 days. After cooling to room temperature, rhombic transparent crystals were obtained with a yield of 12.2% based on In. C/H/N elemental analysis: Anal. Calcd for Compound 1 (wt %): C, 20.80; N, 6.93; H. 3.12. Found: C, 22.36; N, 6.896; H, 3.525. Notes: the as-synthesized microcrystals are sensitive to solvents (water, CH3CN, and ethanol) and lose their crystallinity after solvents' treatment ( Supporting Information Figure S2); no impurities like nonmetallic halide and transitional metal ions were found in 1 ( Supporting Information Figure S3). Route B (Synthesis of 1) In(NO3)3·4.5H2O (73 mg, 0.19 mmol), CH4N2S (85 mg, 1.12 mmol), 2,6-DOC (47 mg, 0.28 mmol), DBN (2 mL, 16.18 mmol), 3-AP (1 mL, 13.14 mmol), and CH3CN (1 mL, 19.24 mmol) were mixed in a 20 mL glass bottle with ultrasonic treatment, then heated to 80 °C for 21 days. After cooling to room temperature, several rhombic transparent crystals were obtained with a yield of 16.8% based on In. C/H/N elemental analysis: Anal. Calcd for Compound 1 (wt%): C, 20.80; N, 6.93; H, 3.12. Found: C, 20.72; N, 7.13; H, 3.307. Single-crystal X-ray diffraction characterization The single-crystal X-ray diffraction measurements of 1 were performed using a Bruker Smart CPAD area diffractometer (Bruker Co. Ltd, Germany) with nitrogen flow temperature controller containing graphite-monochromated MoKα (λ = 0.71073 Å) radiation at 120 K. The structure was solved by the direct method using SHELXS-2014 and the refinement against all reflections of the compound was performed using SHELXL-2014. In these structures, some cations and free solvent molecules were highly disordered and could not be located. The diffuse electron densities resulting from these residual cations and solvent molecules were removed from the dataset using the SQUEEZE routine of PLATON and refined further by performing least-squares refinement cycles based on the data generated. 1H NMR characterization Proton nuclear magnetic resonance (1H NMR) spectra were recorded using a Bruker instrument (400 MHz) and internally referenced to tetramethylsilane or residual solvent signals. The crystals picked out from the mother liquor were cleaned with weighing paper to remove the solvent molecules on the crystal surface and used directly for the NMR analysis. The crystal could be ground for the NMR assay, but the process should be completed as soon as possible because the structure of 1 could be destroyed by solvent due to its poor stability. Then the corresponding signal could be detected using d6-dimethyl sulfoxide (DMSO) as a solvent. We presumed that the signal originated from the incompletely dissolved solid suspended in the solvent. Results and Discussion The rare sp-hybridized L-μ2-S reported here was observed in the hybrid In–S molecular compound 1, synthesized by the solvothermal reaction.23 According to the results of X-ray photoelectron spectroscopy (XPS) ( Supporting Information Figure S4) and electron paramagnetic resonance (EPR) spectra ( Supporting Information Figure S5), the S components in compound 1 were determined as having a valence state of −2, rather than radicals. Single-crystal X-ray diffraction analysis showed that compound 1 consisted of organic imine template molecules and hybrid dimeric supertetrahedral T3-InS clusters, in which three corners of each T3 cluster were coordinated by nitrogen atoms from three neutral imine molecules of DBN, and the fourth corners of the two clusters were connected via the L-μ2-S linker (Figure 2a). Since the two corner-shared T3 clusters adopted a staggered conformation like ethane, the whole cluster-based molecule with a point group close to D3d had overall low system energy. Notably, three highly disordered protonated DBN molecules (denoted as DBN–H+) undoubtedly surrounded the L-μ2-S symmetric center on the equatorial plane (perpendicular to the C3 axis). Detailed structural information on selected bond lengths and bond angles is shown in Figures 2b and 2c. The In–( L-μ2-S) connection is a typical single bond with a bond length of ∼2.491 Å. Intuitively, it is rational for the two adjacent T3 clusters with a negative charge (−6) to adopt a linear connection mode owing to electrostatic repulsion. However, it should be noted that no similar cases had been reported before this work, despite the creation of many supertetrahedral-cluster-based open frameworks. Instead, the adjacent supertetrahedral Tn clusters (n = 2–6) with a high negative charge were bridged via corner-shared sp3-hybridized bent μ2-S2−, trigonal pyramidal μ3-S2−, or tetrahedral μ4-S2− ( Supporting Information Figure S6) regardless of the cluster size.25–31 Given this fact, we tentatively inferred that the three N–H···S hydrogen-bonding interactions stabilized the L-μ2-S mode in compound 1. Figure 2 | (a) Dimeric T3-InS cluster in compound 1 with H-bond-stabilized L-μ2-S. (b) Close-up of the L-μ2-S linker with symbols representing the bond length and angle. (c) Detailed comparison of bond lengths and bond angles (d: data from ref 24). Download figure Download PowerPoint We verified that the linear-bonding structure of the bridging S was independent of crystalline packing but potentially stabilized by the three H-bond donors by performing theoretical calculations on the simplified model structures, M1 and M2 (in −2 charge state), as well as M1–L3 and M2–L3 (in +1 charge state). M1 and M2 denote isolated dimeric T1-InS and T2-InS clusters, respectively, while L represents the DBN–H+ molecule. As shown in Figure 3, the model structures M1 and M2 are both second-order saddle points with two degenerate imaginary frequencies without the H-bond donor. Adjusting the structure of M in terms of the vibrational modes of these frequencies leads to energy minima ( M1′ and M2′) with In–S–In angle of 113.1° and 114.3°, respectively. In contrast, in the presence of three H-bond donors ( L), the model molecules M1–L3 and M2–L3 had equilibrium structures with In–S–In angle of 180° without imaginary frequencies. This shows that the linear S anion found in our study is independent of crystalline packing but dependent on H-bond donors, that is, it was stabilized by H-bonds. Figure 3 | Optimized geometries of model structures. NImag denotes the number of imaginary frequencies. Download figure Download PowerPoint Further, we experimentally verified the hydrogen-bonding interactions between L-μ2-S and DBN–H+ at the equatorial plane using 1H NMR characterization to measure the target hydrogen atom signals. First, the spectrum was computationally simulated using M2–L3 as an example. The results showed that the chemical shift of a hydrogen atom bonded to the nitrogen atom from an isolated DBN–H+ was ∼5.554 ppm ( Supporting Information Figure S7). Once the target hydrogen atom in DBN–H+ interacted with the L-μ2-S linker in the dimeric T2-InS cluster (i.e., M2–L3), the 1H NMR signal displayed a high chemical shift of 9.968 ppm ( Supporting Information Figure S8). Experimentally, we could not obtain a clear solid-state 1H NMR signal of the hydrogen atoms participating in the N–H···S hydrogen bonds ( Supporting Information Figure S9), possibly because the target hydrogen atoms represented a relatively small fraction of the total hydrogen atoms in the DBN molecules. Nevertheless, the desired signal was obtained successfully by using liquid 1H NMR characterization. As shown in Figure 4, a very weak signal at 9.868 ppm was observed when using d6-DMSO as a solvent, consistent with the theoretical value (9.968 ppm). Notably, this chemical shift is much greater than the theoretical value for an isolated DBN–H+ (δ = 5.554 ppm) and is even higher than that observed in a previously reported sample ( ISC-3) (δ = 6.499, Supporting Information Figure S10) with normal N–H···S hydrogen bonds between DBN–H+ template molecular ions and S2− from T3-InS clusters. Thus, the above results clearly demonstrated that the DBN–H+ molecules participated in the formation of N–H···S hydrogen-bonding interactions with L-μ2-S in compound 1. Unfortunately, these N–H···S hydrogen-bond interactions were not detected readily through infrared (IR) and Raman characterizations. Figure 4 | 1H NMR spectrum of compound 1 dispersed in d6-DMSO solvent. Notes: the three strong signals in the range of 4.0–0 ppm correspond to water, DMSO residue, and TMS, respectively; the insets show local coordination and hydrogen bonds around the central L-μ2-S (upper left), a close-up of a hydrogen atom polarized by L-μ2-S (upper right), and the amplified signal at δ = 10.0 ppm (lower). Download figure Download PowerPoint To better understand the electronic structure of 1, the natural bond orbital (NBO)32 and adaptive natural density partitioning (AdNDP)33 analyses were performed using M1–L3 (L = DBN–H+) as an example. The NBO analyses revealed that the Wiberg bond order between the L-μ2-S and two In atoms were 0.48 and 0.49, respectively, which was significantly lower than that of the standard electron-sharing covalent single bond (close to 1.00), indicating its dative bond nature. Such bond nature could be verified by orbital coefficients because the contribution from S comprises 87.23% and 87.57% of the two S–In bonds, respectively. Besides, the s and p components comprise 49.88% and 50.09% for the contribution from S in one of such two bonds, and they comprised 49.60% and 50.38% for the other bond. Obviously, the ratio of s∶p was very close to 1∶1, suggesting sp hybridization. The AdNDP-partitioned electron pairs concerning the bridging S atom are shown in Figure 5. Of the four pairs of valence electrons on the linear S atom, two were involved in the axial bonding, forming two S–In two-center–two-electron (2c–2e) σ-bonds with occupation numbers (ONs) of 1.95 and 1.97 |e|. We have demonstrated the role of the H-bonds by the two AdNDP schemes presented in Figure 5, depicting the radial bonding with regard to the remaining two valence pairs of the S atom. In scheme 1, the radial bonding is described as two 1c–2e lone pairs, both with a low ON of 1.88 |e|, which agrees with the NBO analyses that defined such orbital as the lone pairs both with ON of 1.88 |e|. Notably, neither pair involved the H atoms. However, in scheme 2, the electron pairs are described as H-bond-influenced S lone pairs (4c–2e), both with a higher ON of 1.93 |e|. Furthermore, as shown in the top view, the three H atoms are involved in electron clouds of these electron pairs. Hence, scheme 2 is more reasonable than scheme 1 for describing the radial bonding due to the higher ONs and more rational electron cloud distribution. Accordingly, we inferred from these findings that three evenly distributed H-bonds play a crucial role in converting two highly polarized lone pairs on sp3-hybridized S atom into two nonpolarized lone pairs on an sp-hybridized S atom as observed in compound 1. Therefore, this is an extremely rare case where weak interactions influenced bonding geometry dominantly. Figure 5 | AdNDP views of chemical bonding in model molecule M1–L3 (L = DBN–H+). For the electron pairs distributed perpendicularly to the In–S–In the axis, two bonding schemes were proposed. The two phases in the schemes of 1c–2e S lone pairs and H-bond-influenced S lone pairs (4c–2e) are displayed in green/white and green/yellow, respectively. Download figure Download PowerPoint Conclusion We created a unique case of an sp-hybridized linearly bridging sulfide anion that is stabilized by hydrogen-bonding interactions, a mechanism that is different from the common bent-bridging S with sp3 hybridization and rare sp-hybridized via linear-bridging S stabilization by large steric hindrance or multiple bonds. The unique structure involving hydrogen-bonding interactions was characterized experimentally using crystallography and 1H NMR spectroscopy. The H-bond-assisted stabilizing mechanism was confirmed further and rationalized by theoretical calculations. Such a linear-bonding geometry of S is unprecedented in that two valence lone pairs of S atom did not influence the arrangement of the two σ-bonds, violating the well-established VSEPR theory. Importantly, we believe that the current results would spur a new research avenue toward discovering unique hybridizations of nonmetallic atoms assisted by weak interactions. Supporting Information Supporting Information is available. Conflict of Interest The authors declare no conflict of interest. Acknowledgments The authors acknowledge financial support from the National Natural Science Foundation of China (nos. 21671142, 21875150, and 21720102006), the 111 Project (no. D20015), the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD), Shanxi Natural Science Foundation (no. 201901D111018), and the OIT program of Shanxi Higher Education Institutions. References 1. Gillespie R. J.Improving Our Understanding of Molecular Geometry and the VSEPR Model through the Ligand Close-Packing Model and the Analysis of Electron Density Distributions.Coord. Chem. Rev.2000, 197, 51–69. Google Scholar 2. Guidez E. B.; Gordon M. S.; Ruedenberg K.Why Is Si2H2 Not Linear? An Intrinsic Quasi-Atomic Bonding Analysis.J. Am. Chem. Soc.2020, 142, 13729–13742. Google Scholar 3. Bard N. C.Molecular Geometry Predictions Using Simple MO Theory.J. Chem. Educ.1978, 55, 412–417. Google Scholar 4. Liu F.; Kuliksingle H. J.Impact of Approximate DFT Density Delocalization Error on Potential Energy Surfaces in Transition Metal Chemistry.J. Chem. Theory Comput.2020, 16, 264–277. Google Scholar 5. Cotton F. A.; Cowley A. H.; Feng X.The Use of Density Functional Theory to Understand and Predict Structures and Bonding in Main Group Compounds with Multiple Bonds.J. Am. Chem. Soc.1998, 120, 1795–1799. Google Scholar 6. Mealli C.; Midollini S.; Sacconi L.Preparation and X-ray Structure of a Dimeric Diamagnetic Complex of Nickel (II) with Distorted Tetrahedral Co-Ordination.J. Chem. Soc. Chem. Comm.1975, 19, 765–766. Google Scholar 7. Mealli C.; Sacconi L.MO Nature of the Collinear M-S-M Bridged Linkages in Dinuclear Thiometal Complexes.Inorg. Chem.1982, 21, 2870–2874. Google Scholar 8. Drew M. G. B.; Mitchell P. C. H.; Pygall C. F.Reaction between Molybdate (VI), Cyanide, and Hydrogen Sulfide.Angew. Chem. Int. Ed.1976, 15, 784–785. Google Scholar 9. Mealli C.; Midollini S.; Sacconi L.Transition-Metal Complexes with Sulfur Atom as Ligand. 2.1. Synthesis, Properties, and Structural Characterization of Thio, Mercapto, and Methylthio Complexes of Cobalt(I) and Nickel(I) and -(II) with Poly(tertiary phosphines).Inorg. Chem.1978, 17, 632–637. Google Scholar 10. Greenhough T. J.; Kolthammer B. W. S.; Legzdins P.; Trotter J.Crystal and Molecular Structures of Bis[(η5-cyclopentadienyl)-dicarbonylchromium] Sulfide, a Novel Organometallic Complex Possessing a Cr≡S≡Cr Linkages.Inorg. Chem.1979, 18, 3543–3548. Google Scholar 11. Schiemann J.; Hübener P.; Weiss E.μ2-E[V(CO)3diphos]2 (E = S, Se, Te): Complexes with Linear VEV Multiple-Bond Systems.Angew. Chem. Int. Ed.1983, 22, 980–981. Google Scholar 12. Goh L. Y.; Hambley T. W.; Robertson G. B.Sulphur Chromium Complexes: Syntheses and Crystal Structures of (η5-C5H5)2Cr2(CO)4S and (η5-C5H5)2Cr2(CO)5S2.J. Chem. Soc. Chem. Comm.1983, 23, 1458–1460. Google Scholar 13. Lincoin S.; Soong S.-L.; Koch S. A.; Sato M.; Enemark J. H.Polypyrazolylborate Complexes Containing the [MoSMo]2+ and [MoSeMo]2+ Units.Inorg. Chem.1985, 24, 1355–1359. Google Scholar 14. Thomas S.; Tiekink E. R. T.; Young C. G. (μ-Thio)ditungsten(II) Complexes: Syntheses and Crystal Structures of [{HB(Me2C3N2H)3}W(CO)2]2(μ-S) and [{HB(PriC3N2H2)3}W(CO)2]2(μ-S). Inorg. Chem.1994, 33, 1416–1420. Google Scholar 15. Avens L. R.; Barnhart D. M.; Burns C. J.; McKee S. D.; Smith W. H.Oxidation Chemistry of a Uranium (III) Aryloxide.Inorg. Chem.1994, 33, 4245–4254. Google Scholar 16. Rodriguez M. M.; Stubbert B. D.; Scarborough C. C.; Brennessel W. W.; Bill E.; Holland P. L.Isolation and Characterization of Stable Iron(I) Sulfide Complexes.Angew. Chem. Int. Ed.2012, 51, 8247–8250. Google Scholar 17. Holze P.; Horn B.; Limberg C.; Matlachowski C.; Mebs S.The Activation of Sulfur Hexafluoride at Highly Reduced Low-Coordinate Nickel Dinitrogen Complexes.Angew. Chem. Int. Ed.2014, 53, 2750–2753. Google Scholar 18. Yoshida T.; Adachi T.; Matsumura K.; Baba K.[Mo2(S)2(μ-S)(syn-Me8[16]aneS4)2)]2+, a Novel Sulfidomolybdenum Complex Containing a Linear S=Mo-S-Mo=S Linkages.Angew. Chem. Int. Ed.1993, 32, 1621–1623. Google Scholar 19. Cho J.; Heuvelen K. M. V.; Yap G. P. A.; Brunold T. C.; Riordan C. G.New Synthetic Routes to a Disulfidodinickel(II) Complex: Characterization and Reactivity of a Ni2(μ-η2:η2-S2) Core.Inorg. Chem.2008, 47, 3931–3933. Google Scholar 20. Delgado S.; Miguel P. J. S.; Priego J. L.; Jimenez-Aparicios R.; Gomez-García C. J.; Zamora F.A Conducting Coordination Polymer Based On Assembled Cu9 Cages.Inorg. Chem.2008, 47, 9128–9130. Google Scholar 21. Lam O. P.; Heinemann F. W.; Meyer K.Activation of Elemental S, Se and Te with Uranium(III): Bridging U-E-U (E = S, Se) and Diamond-Core Complexes U-(E)2-U (E = O, S, Se, Te).Chem. Sci.2011, 2, 1538–1547. Google Scholar 22. Yu L. C.; Peng H. M.Synthesis and Crystal Structure of a Pb Complex with Ciprofloxacin.Russ. J. Coord. Chem.2012, 38, 392–395. Google Scholar 23. Wu T.; Bu X.; Liao P.; Wang L.; Zheng S. T.; Ma R.; Feng P.Superbase Route to Supertetrahedral Chalcogenide Clusters.J. Am. Chem. Soc.2012, 134, 3619–3622. Google Scholar 24. Platts J. A.; Howard S. T.; Bracke B. R. F.Directionality of Hydrogen Bonds to Sulfur and Oxygen.J. Am. Chem. Soc.1996, 118, 2726–2733. Google Scholar 25. Xu X.; Wang W.; Liu D.; Hu D.; Wu T.; Bu X.; Feng P.Pushing Up the Size Limit of Metal Chalcogenide Supertetrahedral Nanocluster.J. Am. Chem. Soc.2018, 140, 888–891. Google Scholar 26. Bu X.; Zheng N.; Li Y.; Feng P.Templated Assembly of Sulfide Nanoclusters into Cubic-C3N4 Type Framework.J. Am. Chem. Soc.2003, 125, 6024–6025. Google Scholar 27. Zhang L.; Xue C.; Wang W.; Hu D.; Lv J.; Li D; Wu T.Stable Supersupertetrahedron with Infinite Order via the Assembly of Supertetrahedral T4 Zinc-Indium Sulfide Clusters.Inorg. Chem.2018, 57, 10485–10488. Google Scholar 28. Feng P.; Bu X.; Zheng N.The Interface Chemistry between Chalcogenide Clusters and Open Framework Chalcogenides.Acc. Chem. Res.2005, 38, 293–303. Google Scholar 29. Zheng N.; Bu X.; Lu H.; Zhang Q.; Feng P.Crystalline Superlattices from Single-Sized Quantum Dots.J. Am. Chem. Soc.2005, 127, 11963–11965. Google Scholar 30. Zhang Q.; Bu X.; Zhang J.; Wu T.; Feng P.Chiral Semiconductor Frameworks from Cadmium Sulfide Clusters.J. Am. Chem. Soc.2007, 129, 8412–8413. Google Scholar 31. Zhang Q.; Bu X.; Lin Z.; Wu T.; Feng P.Organization of Tetrahedral Chalcogenide Clusters Using a Tetrahedral Quadridentate Linker.Inorg. Chem.2008, 47, 9724–9726. Google Scholar 32. Reed A. E.; Curtiss L. A.; Weinhold F.Intermolecular Interactions from a Natural Bond Orbital, Donor-Acceptor Viewpoint.Chem. Rev.1988, 88, 899–926. Google Scholar 33. Zubarev D. Y.; Boldyrev A. I.Developing Paradigms of Chemical Bonding: Adaptive Natural Density Partitioning.Phys. Chem. Chem. Phys.2008, 10, 5207–5217. Google Scholar Previous article FiguresReferencesRelatedDetails Issue AssignmentVolume 3Issue 10Page: 2584-2590Supporting Information Copyright & Permissions© 2020 Chinese Chemical SocietyKeywordshydrogen bondVSEPR theoryindium sulfidesp-hybridized sulfurlinear coordinationAcknowledgmentsThe authors acknowledge financial support from the National Natural Science Foundation of China (nos. 21671142, 21875150, and 21720102006), the 111 Project (no. D20015), the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD), Shanxi Natural Science Foundation (no. 201901D111018), and the OIT program of Shanxi Higher Education Institutions. Downloaded 976 times PDF downloadLoading ...

Referência(s)
Altmetric
PlumX