ABSTRACT
Aromatic metallacycles are of considerable current interest. Reported aromatic metallacycles are mainly those with carbon, nitrogen, oxygen and sulfur. In this work, we report the synthesis and characterization of aromatic chloroosmacyclopentatrienes, which represent the first structurally confirmed metallaaromatic with a chlorine atom in its framework. Single-crystal X-ray diffraction studies show that these planar chloroosmacyclopentatrienes possess a very short Os–ClC distance suggesting M=ClC bond character.
Keywords: metallacycles, chlorometallacyclopentatriene, osmium, o-ethynylphenyl alkynes, halocarbon complex
Metal-chlorine bonds of halocarbon complexes are usually longer than metal-halide bonds. By contrast, this work demonstrates that complexes LnOsCl(ClR) have the Os−ClR bond shorter than the Os−Cl bond.
INTRODUCTION
The chemical bond between an s-block metal and chlorine (such as LiCl or NaCl) is usually described as an ionic bond in textbooks [1]. Compared to s-block metals, d-block transition metals can have more types of bonding with chlorine. In addition to well-known metal halide complexes, which contain ionic or covalent σ M–Cl bonds (type I in Fig. 1), d-block transition metals can also form complexes with chlorocarbons (ClR). In coordination complexes with chlorocarbons, a dative M(η1-ClC) bond (type II in Fig. 1), in which the chlorocarbon serves as a 2e donating ligand, is commonly assumed [2]. A chlorine atom in chlorocarbons could also interact with d-block transition metals via other types of bonds, for example, the covalent type III in chloronium form [3] and the type IV containing a metal-chlorine double bond. The organic counterparts of the covalent type III bonding were reported a long time ago [4]. For example, electrophilic addition reactions of alkenes with Cl2 are known to occur via cyclochloronium intermediates. The covalent type IV bonding is closely related to the bonding in inorganic compounds such as FCl=O and Cl3CCl=O [5]. However, examples of organometallic versions of M(η1-Cl+C) (type III) and M=ClC (type IV) bonds are rarely reported.
Figure 1.

Types of bonding between transition metals and (I) chloride, (II–IV) chlorocarbons.
Recently, the chemistry of compounds with an unusual carbon-halogen multiple bond is attracting attention both experimentally and theoretically. Novel compounds reported recently include compounds with a terminal C=Cl or C≡Cl bond [6], iodabenzene [7] and an η3-phenyldichloromethyl (η3-PhCCl2) complex [8]. Herein, we report the isolation and characterization of planar chlorometallacyclopentatrienes, in which the bond distance between the osmium atom and the ClC unit is very short, and the Os–ClC bond can be viewed as the M=ClC bond of type IV, possibly stabilized by the aromaticity of the metallacycle [9–24].
RESULTS AND DISCUSSION
The first example of a chlorometallacyclopentatriene was isolated during our investigation of the reactivity of OsCl2(PPh3)3 (1) with the o-ethynylphenyl alkynes [25–28]. Thus, treatment of OsCl2(PPh3)3 (1) with 1-ethynyl-2-(phenylethynyl)benzene (2) and excess HCl in dichloromethane at room temperature for 3 h produced a green solution, from which the complex 3 was isolated as a green solid in 62.0% yield (Scheme 1).
Scheme 1.

Osmium-mediated cyclization of an enediyne in the presence of HCl.
We propose that complex 3 might be formed via intermediates A and B as shown in Scheme 1. The reaction of OsCl2(PPh3)3 (1) with 2 could initially generate the vinylidene intermediate A. Reactions of OsCl2(PPh3)3 (1) with terminal alkynes (HC≡CR) to form osmium vinylidene complexes OsCl2(=C=CHR)(PPh3)2 are well-documented reactions [29–31]. It has also been reported that osmium vinylidene complexes can be protonated at their β-carbon to give osmium carbyne complexes [29–31]. Thus, the intermediate A could be protonated at the C2 position to give the intermediate B, which may undergo an electrophilic cyclization followed by the addition of Cl− to give complex 3. Consistent with the proposed mechanism, the reaction in the presence of DCl produced the partially deuterated complex 3D (Fig. 2) with a CHD methylene group (see Figs S11–S13 for characterization data).
Figure 2.

The structures of 3D and 4–8.
The structure of complex 3 has been determined by single-crystal X-ray diffraction studies, which confirmed that complex 3 is a tricyclic metallacycle (Fig. 3). The five-membered osmacycle is almost planar, as reflected by the mean deviation (0.021 Å) from the least-squares plane of Os1, C1, C5, C6 and Cl1. The Os–C1 bond length of 1.903(4) Å is in the range of reported Os=CHR lengths (1.810–2.142 Å, based on a search of the Cambridge Structural Database, CSD version 2020 (Nov 2019)) [32–36], which suggests its double-bond character. The most fascinating structural feature in metallacycle 3 is that the distance of Os–Cl1 (2.359 Å) is even shorter (by 0.040 Å) than that of the terminal Os–Cl3 bond (2.399 Å), indicating the unusual bonding (beyond a σ-bonding) between Os and Cl. It is noted that there are only two X-ray single-crystal structures (complexes 4 and 5 shown in Fig. 2) containing short Os···ClC contacts in CSD version 2020 (Nov 2019). In these two examples, the bond lengths of Os–ClAr are 2.574(1) Å for 4 [37] and 2.539(6) Å for 5 [38], which are typical dative bond lengths. Thus, complex 3 represents an interesting example that contains an unusually short Os–ClC bond. Notably, all the CSD crystal structures that contain both M–ClC and terminal M–Cl (M: transition metal) bonds have a distance of an M–ClC longer than that of a terminal M–Cl.
Figure 3.

Single-crystal X-ray structure of complex 3 with thermal ellipsoids at the 50% probability level (phenyl groups in PPh3 are omitted for clarity). Selected bond distances (Å) and angles (°) in 3: Os1–Cl3 2.3993(11), Os1–Cl2 2.5135(10), Os1–Cl1 2.3594(10), Os1–C1 1.903(4), C1-C2 1.515(6), C2-C3 1.495(6), C3-C4 1.397(7), C4-C5 1.480(6), C5-C6 1.336(6), C6-Cl1 1.781(4), C1-C5 1.498(6); Os1-C1-C5 121.7(3), C1-C5-C6 121.1(4), C5-C6-Cl1 113.2(3), C6-Cl1-Os1 101.28(15), Cl1-Os1-C1 82.61(13), C1-C2-C3 105.6(4), C2-C3-C4 111.3(4), C3-C4-C5 108.4(4), C4-C5-C1 108.5(4), C5-C1-C2 106.1(3).
Furthermore, the distance of Cl1–C6 (1.781(4) Å) is longer than those of Cl–C(sp2) bonds in dative ClC unit (1.739 Å for 4 and 1.754 Å for 5) and longer than those of organic Cl–C(sp2) bonds (e.g. 1.712–1.747 Å for 4 and 5). Moreover, the bond lengths of C5–C6 (1.336(6) Å) and C1–C5 (1.498(6) Å) in 3 are in the range of C–C single (1.54 Å) and double (1.33 Å) bond distances. These results suggest that a weak delocalization is present in the chlorometallacycle ring of complex 3. Previously reported structurally characterized complexes that are closely related to complex 3 include [Cp*Fe(CO){k2-C(OMe)C6H4-o-Cl}]OTf (6) and RuCl2(k2-CH-Ar-o-X)(IMes) (7, X = Br; 8, X = I) [39–42]. The five-membered chlorometallacycle in 6 does not however adopt a planar structure [39]. The complexes 7 (X = Br) and 8 (X = I) have longer Ru–XAr (X = Br, I) bond distances (longer than 2.5007 Å) [40–42], which are typical dative bonds. Thus, the structural features of complex 3 are completely different from those of complexes 6–8.
The nuclear magnetic resonance (NMR) spectroscopy and high-resolution mass spectroscopy (HRMS) data of 3 are consistent with the solid structure. For instance, complex 3 shows the signal of proton (C2H) at 0.57 ppm in the 1H NMR spectrum and the signal of C2 at 62.0 ppm in the 13C{1H} NMR spectrum, confirming that C2 has an sp3 hybridized character. The signal of metal-bonded C1 is located at 271.6 ppm in the 13C{1H} NMR spectrum, supporting the metal-carbene character of C1. Only one sharp signal appeared at -19.4 ppm in the 31P{1H} NMR spectrum, indicating the equivalent chemical environment of the two trans PPh3 ligands, which confirms the planar structure of fused rings. The molecular formula of complex 3 was also confirmed by HRMS data (m/z = 989.1647 for [3-Cl]+).
As shown in scheme 1, there are three possible resonance forms that could contribute to the overall structure of complex 3: the neutral metallacycle 3, the zwitterionic form 3’ and the chloride dative form 3”. The resonance form 3” cannot well account for the fact that the distance of Os–Cl1 (2.359 Å) in the metallacycle is much shorter than those of dative Os–ClAr bonds (2.574 Å for 4 and 2.538 Å for 5) and even shorter than that of Os–Cl3 (2.399 Å, a typical metal chloride σ-bond). The short Os–ClC distance in 3 suggests the M–ClC has a double-bond character, which can be attributed to the contribution from the resonance form 3. Complex 3 can be regarded as a chlorometallacyclopentatriene, which is a metal analog of chlorophenium ([C4H4Cl]+). Chlorophenium ([C4H4Cl]+) is closely related electronically to five-membered heteroaromatics such as thiophene or furan.
To check whether the chloroosmacyclopentatriene is aromatic or not, we calculated the nucleus-independent chemical shift (NICS) value [43], an index used to evaluate aromaticity or anti-aromaticity. The structure of 3 was optimized at PBE/def2-TZVP level [44,45] and the value of NICS(1)zz of the chlorometallacycle ring in complex 3 was calculated (at PBE/x2c-TZVPall level [44,46]) to be -9.97 (Fig. 4), suggesting that chlorometallacycle 3 has an aromatic character.
Figure 4.

NICS(0) and NICS(1)zz values for complexes 3, 3H and 3M; the isomerization stabilization energy of 3M from 3H; key occupied π-MOs of 3M (the isosurface value is 0.02).
To further understand the aromaticity of 3, we have performed a theoretical study of the model complex OsCl2(PH3)2(k2-ClC3H2Me) (3M, see Fig. 4 for its structure) at PBE/def2-TZVP level [44,45]. The optimized structure of 3M well reproduced the key structural feature of 3. The aromatic character of the chloroosmacyclopentatriene 3M is indicated by the calculated isomerization stabilization energy of -11.3 kcal/mol (the energy involved in the isomerization of 3H to form 3M, see Fig. 4). Moreover, the negative NICS(1)zz value (−7.59, calculated at PBE/x2c-TZVPall level [44,46] based on the optimized structure) of 3M also supports the aromaticity of the chloroosmacyclopentatriene.
As shown in Fig. 4, there are a total of four key occupied π-type molecular orbitals (HOMO-5, HOMO-7, HOMO-10, HOMO-13, see Fig. 4) for the metallacycle of 3M, indicating that the chlorometallacycle has eight π electrons. Thus, this planar chlorometallacyclopentatriene 3 is a unique Möbius [17,19,22,47–50] aromatic complex. It is worth noting that the π-bonding molecular orbitals of chlorometallacycle are generated by interactions of two d orbitals of an OsX2L2 fragment with the π-type molecular orbitals of a ClC3H2Me fragment. In other words, osmium can form π-bonds with the carbon-bound chlorine atom by back-donation from the osmium d orbitals to the π-type molecular orbitals of the ClC3H2Me fragment. As a result, the unusually short M–ClC bond in 3 can be partially related to a dπ-pπ π-interaction between the osmium atom and the ClC unit, as indicated by the resonance form 3. Thus, chloroosmacyclopentatriene 3 represents the first structurally confirmed aromatic chlorometallacycle with an M=Cl–C character.
The unusual structural feature of complex 3 prompted us to synthesize its analogues. By reacting o-ethynylphenyl alkynes 9, 11 and 13 with OsCl2(PPh3)3 (1), the chloroosmacyclopentatriene derivatives 10, 12 and 14 with a tBu, p-Tol or Si(iPr)3 (TIPS) substituent, respectively, were obtained (Scheme 2).
Scheme 2.

The preparation of chloroosmacyclopentatrienes.
The complexes 10, 12 and 14 were fully characterized by NMR spectroscopy and X-ray diffraction studies (Fig. 5). They all have structural features similar to those in 3, in which the M–ClR bond is systematically shorter than the corresponding M–Cl(terminal) bond. Subtle differences in the structures of complexes 3, 10, 12 and 14 are noted (Fig. 5). The Os–ClR bond in p-Tol-substituted osmium complex 12 (2.3624(8) Å) is similar to that of Ph-substituted osmium complex 3 (2.3594(10) Å), and an appreciably shorter Os–ClR bond is observed for tBu-substituted complex 10 (2.315(3) Å) and TIPS-substituted osmium complex 14 (2.3384(8) Å). The difference in the Os–Cl1(chlorocarbon) and Os–Cl3(terminal) bond distances in tBu-substituted complex 10 is as large as 0.091 Å.
Figure 5.
Single-crystal X-ray structures of 10, 12 and 14 with thermal ellipsoids at 50% probability level (the phenyl groups in PPh3 and the isopropyl groups in the Si(iPr)3 are omitted for clarity), and selected bond distances (Å) for complexes 3, 10, 12 and 14.
CONCLUSION
In summary, OsCl2(PPh3)3 was found to react with o-ethynylphenyl alkynes in the presence of HCl to give chlorometallacyclopentatriene complexes. These complexes are unusual in that the M–ClC(halocarbon) bond is appreciably shorter than the M–Cl(chloride) bond, and they can be regarded as compounds possessing an M=ClC bond. The novel metallacycles exhibit aromaticity, excellent planarity and thermal stability. This study not only enriches the family of metallacyclic chemistry, but also provides the synthesis of novel metallacycles containing an M=ClC bond.
Supplementary Material
ACKNOWLEDGEMENTS
We thank Prof. Zhenyang Lin for providing detailed comments regarding structure and bonding aspect.
Contributor Information
Zhenwei Chu, Department of Materials Science and Engineering, College of Materials, Xiamen University, Xiamen 361005, China.
Guomei He, Department of Materials Science and Engineering, College of Materials, Xiamen University, Xiamen 361005, China.
Chuan Shi, Department of Chemistry, The Hong Kong University of Science and Technology, Hong Kong, China.
Yuhui Hua, Department of Chemistry, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, China.
Yaxi Huang, Department of Materials Science and Engineering, College of Materials, Xiamen University, Xiamen 361005, China.
Jiangxi Chen, Department of Materials Science and Engineering, College of Materials, Xiamen University, Xiamen 361005, China.
Hujun Xie, Department of Applied Chemistry, Zhejiang Gongshang University, Hangzhou 310018, China.
Guochen Jia, Department of Chemistry, The Hong Kong University of Science and Technology, Hong Kong, China.
FUNDING
This work was supported by the National Natural Science Foundation of China (21772160) and the Hong Kong Research Grants Council (16308719).
AUTHOR CONTRIBUTIONS
J.C., H.X. and G.J. conceived the project. Z.C. and G.H. performed all the experimental works and analyzed the experimental data. Z.C. and H.X. conceived the theoretical work. Z.C., C.S. and H.X. analyzed and interpreted the computational data. Z.C., Y.-H.H. and Y.-X.H. analyzed and interpreted the single-crystal X-ray data. J.C. and G.J. drafted the paper. All of the authors discussed the results and contributed to the preparation of the final manuscript.
Conflict of interest statement
None declared.
REFERENCES
- 1. Atkins P, Overton T, Rourke Jet al. Inorganic Chemistry, 4th edn. Oxford: Oxford University Press, 2006, 400. [Google Scholar]
- 2. Kulawiec RJ, Crabtree RH. Coordination chemistry of halocarbons. Coord Chem Rev 1990; 99: 89–115. 10.1016/0010-8545(90)80061-W [DOI] [Google Scholar]
- 3. Butts MD, Scott BL, Kubas GJ. Syntheses and structures of alkyl and aryl halide complexes of the type [(PiPr3)2PtH(η1-XR)]BArf and analogues with Et2O, THF, and H2 ligands. Halide-to-metal π bonding in halocarbon complexes. J Am Chem Soc 1996; 118: 11831–43. 10.1021/ja961836y [DOI] [Google Scholar]
- 4. Sandin RB, Hay AS. Stable bromonium and chloronium salts. J Am Chem Soc 1952; 74: 274–5. 10.1021/ja01121a524 [DOI] [Google Scholar]
- 5. Kalemos A, Mavridis A. Myths and reality of hypervalent molecules. The electronic structure of FClOx, x = 1–3, Cl3PO, Cl3PCH2, Cl3CClO, and C(ClO)4. J Phys Chem A 2011; 115: 2378–84. 10.1021/jp111330y [DOI] [PubMed] [Google Scholar]
- 6. Fantuzzi F, Rudek B, Wolff Wet al. Doubly and triply charged species formed from chlorobenzene reveal unusual C–Cl multiple bonding. J Am Chem Soc 2018; 140: 4288–92. 10.1021/jacs.7b12749 [DOI] [PubMed] [Google Scholar]
- 7. Rawashdeh AM, Chakkingal PP, Zeng Tet al. An iodabenzene story. J Am Chem Soc 2017; 139: 7124–9. 10.1021/jacs.7b03388 [DOI] [PubMed] [Google Scholar]
- 8. Yeh W-Y. Unusual thermal reactivity of [W(≡CPh)(NCMe)(η2-C60)(η5-C5 Ph5)] in chlorobenzene involving activation of all ligands. Angew Chem Int Ed 2011; 50: 12046–9. 10.1002/anie.201105542 [DOI] [PubMed] [Google Scholar]
- 9. Zhang Y, Yu C, Huang Zet al. Metalla-aromatics: planar, nonplanar, and spiro. Acc Chem Res 2021; 54: 2323–33. 10.1021/acs.accounts.1c00146 [DOI] [PubMed] [Google Scholar]
- 10. Chen D, Hua Y, Xia H. Metallaaromatic chemistry: history and development. Chem Rev 2020; 120: 12994–3086. 10.1021/acs.chemrev.0c00392 [DOI] [PubMed] [Google Scholar]
- 11. Chen D, Xie Q, Zhu J. Unconventional aromaticity in organometallics: the power of transition metals. Acc Chem Res 2019; 52: 1449–60. 10.1021/acs.accounts.9b00092 [DOI] [PubMed] [Google Scholar]
- 12. Zhu C, Xia H. Carbolong chemistry: a story of carbon chain ligands and transition metals. Acc Chem Res 2018; 51: 1691–700. 10.1021/acs.accounts.8b00175 [DOI] [PubMed] [Google Scholar]
- 13. Frogley BJ, Wright LJ. Recent advances in metallaaromatic chemistry. Chem Eur J 2018; 24: 2025–38. 10.1002/chem.201704888 [DOI] [PubMed] [Google Scholar]
- 14. Ma W, Yu C, Chen Tet al. Metallacyclopentadienes: synthesis, structure and reactivity. Chem Soc Rev 2017; 46: 1160–92. 10.1039/C6CS00525J [DOI] [PubMed] [Google Scholar]
- 15. Frogley BJ, Wright LJ. Fused-ring metallabenzenes. Coord Chem Rev 2014; 270–1: 151–66. 10.1016/j.ccr.2014.01.019 [DOI] [Google Scholar]
- 16. Herges R. Topology in chemistry: designing Möbius molecules. Chem Rev 2006; 106: 4820–42. 10.1021/cr0505425 [DOI] [PubMed] [Google Scholar]
- 17. Huang Z, Zhang Y, Zhang W-Xet al. A tris-spiro metalla-aromatic system featuring craig-möbius aromaticity. Nat Commun 2021; 12: 1319. 10.1038/s41467-021-21648-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Zhang Y, Wei J, Zhu Met al. Tetralithio metalla-aromatics with two independent perpendicular dilithio aromatic rings spiro-fused by one manganese atom. Angew Chem Int Ed 2019; 58: 9625–31. 10.1002/anie.201904681 [DOI] [PubMed] [Google Scholar]
- 19. An K, Shen T, Zhu J. Craig-type möbius aromaticity and antiaromaticity in dimetalla[10]annulenes: a metal-induced Yin-and-Yang pair. Organometallics 2017; 36: 3199–204. 10.1021/acs.organomet.7b00341 [DOI] [Google Scholar]
- 20. Zhang Y, Wei J, Chi Yet al. Spiro metalla-aromatics of Pd, Pt, and Rh: synthesis and characterization. J Am Chem Soc 2017; 139: 5039–42. 10.1021/jacs.7b02039 [DOI] [PubMed] [Google Scholar]
- 21. Liu L, Zhu M, Yu HTet al. Organocopper(III) spiro complexes: synthesis, structural characterization, and redox transformation. J Am Chem Soc 2017; 139: 13688–91. 10.1021/jacs.7b08803 [DOI] [PubMed] [Google Scholar]
- 22. Zhu C, Li S, Luo Met al. Stabilization of anti-aromatic and strained five-membered rings with a transition metal. Nat Chem 2013; 5: 698–703. 10.1038/nchem.1690 [DOI] [PubMed] [Google Scholar]
- 23. Wei J, Zhang Y, Chi Yet al. Aromatic dicupra[10]annulenes. J Am Chem Soc 2016; 138: 60–3. 10.1021/jacs.5b11317 [DOI] [PubMed] [Google Scholar]
- 24. Batuecas M, Castro-Rodrigo R, Esteruelas MAet al. Aromatic osmacyclopropenefuran bicycles and their relevance for the metal-mediated hydration of functionalized allenes. Angew Chem Int Ed 2016; 55: 13749–53. 10.1002/anie.201606839 [DOI] [PubMed] [Google Scholar]
- 25. Chu Z, He G, Cheng Xet al. Synthesis and characterization of cyclopropaosmanaphthalenes containing a fused σ-aromatic metallacyclopropene unit. Angew Chem Int Ed 2019; 58: 9174–8. 10.1002/anie.201904815 [DOI] [PubMed] [Google Scholar]
- 26. Esteruelas MA, Oñate E, Paz Set al. Repercussion of a 1,3-hydrogen shift in a hydride-osmium-allenylidene complex. Organometallics 2021; 40: 1523–37. 10.1021/acs.organomet.1c00176 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Álvarez-Pérez A, González-Rodríguez C, García-Yebra Cet al. Catalytic cyclization of o-alkynyl phenethylamines via osmacyclopropene intermediates: direct access to dopaminergic 3-benzazepines. Angew Chem Int Ed 2015; 54: 13357–61. 10.1002/anie.201505782 [DOI] [PubMed] [Google Scholar]
- 28. Zhuo Q, Lin J, Hua Yet al. Multiyne chains chelating osmium via three metal-carbon σ bonds. Nat Commun 2017; 8: 1912. 10.1038/s41467-017-02120-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Wen TB, Hung WY, Zhou ZYet al. Synthesis and characterization of [OsCl2(=C=CHR)(PPh3)2] and related complexes. Eur J Inorg Chem 2004; 2837–46. 10.1002/ejic.200400003 [DOI] [Google Scholar]
- 30. Asensio A, Buil ML, Esteruelas MAet al. A four-electron π-alkyne complex as precursor for allenylidene derivatives: preparation, structure, and reactivity of [Os(η5-C5H5)(CCCPh2)L(PiPr3)]PF6 (L = CO, PHPh2). Organometallics 2004; 23: 5787–98. 10.1021/om049479k [DOI] [Google Scholar]
- 31. Bolaño T, Collado A, Esteruelas MAet al. Selectivity of allenylidene versus butadienyl protonation in an osmium−bisphosphine system. Organometallics 2009; 28: 2107–11. 10.1021/om900059g [DOI] [Google Scholar]
- 32. Buil ML, Cardo JJF, Esteruelas MAet al. Square-planar alkylidyne–osmium and five-coordinate alkylidene–osmium complexes: controlling the transformation from hydride-alkylidyne to alkylidene. J Am Chem Soc 2016; 138: 9720–8. 10.1021/jacs.6b05825 [DOI] [PubMed] [Google Scholar]
- 33. Zhu C, Wu J, Li Set al. Synthesis and characterization of a metallacyclic framework with three fused five-membered rings. Angew Chem Int Ed 2017; 56: 9067–71. 10.1002/anie.201704057 [DOI] [PubMed] [Google Scholar]
- 34. Frogley BJ, Wright LJ. A metallaanthracene and derived metallaanthraquinone. Angew Chem Int Ed 2017; 56: 143–7. 10.1002/anie.201608500 [DOI] [PubMed] [Google Scholar]
- 35. Bohle DS, Clark GR, Rickard CEFet al. Terminal methylene complexes of ruthenium(II) and osmium(II) and intramolecular methylene and acyl ligand combination to form metallaoxetenes: the crystal structures of [OsCl(η2-C[O]-o-tolyl)(=CH2)(PPh3)2] and [Ru(=C[Ph]OCH2)(CN-p-tolyl)2(PPh3)2]ClO4. J Chem Soc, Chem Commun 1987; 563–5. 10.1039/C39870000563 [DOI] [Google Scholar]
- 36. Rickard CEF, Roper WR, Woodgate SDet al. Electrophilic aromatic substitution reactions of a metallabenzene: nitration and halogenation of the osmabenzene [Os{C(SMe)CHCHCHCH}I(CO)(PPh3)2]. Angew Chem Int Ed 2000; 39: 750–2. [DOI] [PubMed] [Google Scholar]
- 37. Tena MA, Nürnberg O, Werner H. Fünf- und sechsfach koordinierte hydrido(phenolat)- und hydrido(thiophenolat)-komplexe von ruthenium und osmium. Chem Ber 1993; 126: 1597–602. 10.1002/cber.19931260715 [DOI] [Google Scholar]
- 38. Chen J, Sung HH-Y, Williams IDet al. Hydrolysis of osmium and ruthenium carbyne complexes. Organometallics 2011; 30: 6159–65. 10.1021/om200685v [DOI] [Google Scholar]
- 39. Poignant G, Nlate S, Guerchais Vet al. Synthesis and properties of (η2-C,X) chelate arylcarbene complexes [Fe(C5Me5)(L){η2-C(OMe)C6h4-o-X}][OTf] (L = CO, PMe3; x = OMe, Cl). Organometallics 1997; 16: 124–32. 10.1021/om960711z [DOI] [Google Scholar]
- 40. Wappel J, Grudzień K, Barbasiewicz Met al. Initiation efficacy of halo-chelated cis-dichloro-configured ruthenium-based second-generation benzylidene complexes in ring-opening metathesis polymerization. Monatsh Chem 2015; 146: 1153–60. 10.1007/s00706-015-1494-8 [DOI] [Google Scholar]
- 41. Grudzień K, Żukowska K, Malińska Met al. Mechanistic studies of Hoveyda–Grubbs metathesis catalysts bearing S-, Br-, I-, and N-coordinating naphthalene ligands. Chem Eur J 2014; 20: 2819–28. 10.1002/chem.201303826 [DOI] [PubMed] [Google Scholar]
- 42. Barbasiewicz M, Michalak M, Grela K. A new family of halogen-chelated Hoveyda–Grubbs-type metathesis catalysts. Chem Eur J 2012; 18: 14237–41. 10.1002/chem.201202817 [DOI] [PubMed] [Google Scholar]
- 43. Chen Z, Wannere CS, Corminboeuf Cet al. Nucleus-independent chemical shifts (NICS) as an aromaticity criterion. Chem Rev 2005; 105: 3842–88. 10.1021/cr030088+ [DOI] [PubMed] [Google Scholar]
- 44. Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made simple. Phys Rev Lett 1996; 77: 3865–8. 10.1103/PhysRevLett.77.3865 [DOI] [PubMed] [Google Scholar]
- 45. Weigend F, Ahlrichs R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: design and assessment of accuracy. Phys Chem Chem Phys 2005; 7: 3297–305. 10.1039/b508541a [DOI] [PubMed] [Google Scholar]
- 46. Pollak P, Weigend F. Segmented contracted error-consistent basis sets of double- and triple-ζ valence quality for one- and two-component relativistic all-electron calculations. J Chem Theory Comput 2017; 13: 3696–705. 10.1021/acs.jctc.7b00593 [DOI] [PubMed] [Google Scholar]
- 47. Foroutan-Nejad C, Vι´cha J, Ghosh A. Relativity or aromaticity? A first-principles perspective of chemical shifts in osmabenzene and osmapentalene derivatives. Phys Chem Chem Phys 2020; 22: 10863–9. 10.1039/D0CP01481H [DOI] [PubMed] [Google Scholar]
- 48. Mauksch M, Tsogoeva SB. Demonstration of “Möbius” aromaticity in planar metallacycles. Chem Eur J 2010; 16: 7843–51. 10.1002/chem.201000396 [DOI] [PubMed] [Google Scholar]
- 49. Craig DP, Paddock NL. A novel type of aromaticity. Nature 1958; 181: 1052–3. 10.1038/1811052a0 [DOI] [Google Scholar]
- 50. Heilbronner E. Hückel molecular orbitals of Möbius-type conformations of annulenes. Tetrahedron Lett 1964; 5: 1923–8. 10.1016/S0040-4039(01)89474-0 [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.

