An Approach to Carbide-Centered Cluster Complexes | Inorganic

pdf. ic8b03222_si_001.pdf (2.35 MB) ... Universitetsparken 5, DK. -. 2100 Copenhagen (Denmark). E. -. mail: [email protected],. [email protected]. [...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/IC

Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

An Approach to Carbide-Centered Cluster Complexes Anders Reinholdt,*,†,# Sean H. Majer,§,# Rikke M. Gelardi,† Samantha N. MacMillan,§ Anthony F. Hill,⊥ Ola F. Wendt,∥ Kyle M. Lancaster,*,§ and Jesper Bendix*,† †

Department of Chemistry, University of Copenhagen, Universitetsparken 5, DK-2100 Copenhagen, Denmark Department of Chemistry and Chemical Biology, Baker Laboratory, Cornell University, Ithaca, New York 14853, United States ⊥ Research School of Chemistry, Australian National University, Canberra, Australian Capital Territory 2601, Australia ∥ Centre for Analysis and Synthesis, Department of Chemistry, Lund University, P.O. Box 124, 22100 Lund, Sweden Inorg. Chem. Downloaded from pubs.acs.org by UNIV OF NEW ENGLAND on 03/27/19. For personal use only.

§

S Supporting Information *

ABSTRACT: We report the first examples of the carbide ligand in (Cy3P)2Cl2RuC (RuC) developing into a μ3 ligand toward metal centers. Conventionally, sterics exclude this coordination mode, but Fe2(CO)9 and Co2(CO)8 expel bridging CO ligands upon reaction with RuC to form trimetallic (Cy3P)2Cl2RuCFe2(CO)8 (RuCFe2) and (Cy3P)2Cl2Ru CCo2(CO)7 (RuCCo2) complexes. Thus, the proximity offered by metal− metal associations in bimetallic carbonyl complexes allows the formation of trinuclear carbide complexes as verified by NMR, Mössbauer, and X-ray spectroscopic techniques.



INTRODUCTION Composed of iron, sulfide, and a central six-coordinate carbide ligand, the active-site cofactor in nitrogenase is a remarkable bioinorganic construct,1 but due to its complexity, it has not been prepared by synthetic inorganic routes.2 In principle, molecular carbide complexes offer a desirable synthetic approach to carbide-centered clusters via repeated coordination of metal centers at their MC3 or MCM4 cores. However, branching at the interior of sterically encumbered molecules is demanding. In this respect, the closeness upheld in complexes with metal−metal associations offers reagents that are predisposed to overcome steric encumbrance and form polynuclear systems. Yet, molecules combining metal−metal and metal−ligand multiple bonding remain rare.5 Upon association with metal fragments, the carbide ligand in (Cy3P)2Cl2RuC (RuC, Cy = cyclohexyl) invariably assumes a μ2 coordination mode to produce (Cy3P)2Cl2RuC−[M] complexes.6 Similarly, reactions with nonmetal fragments almost exclusively generate two-coordinate carbon.7 A solitary example of three-coordinate RuC-derived carbon arises in reactions with MeO2CCCCO2Me, yielding the cyclopropenylidene complex, (Cy3P)2Cl2RuCC2(CO2Me)2.8 In their seminal computational study, Krapp and Frenking9 nevertheless suggested μ3 coordination to be energetically favored when the carbide ligand in RuC replaces one CO ligand in Fe2(CO)9. With PMe3 emulating the PCy3 ligands in RuC, the model complex leaves unanswered whether RuC is sufficiently compact to allow coordination of its carbide ligand to multiple metal centers. Moreover, the calculated energy for © XXXX American Chemical Society

ligand substitution in Fe2(CO)9 indicates that a RuC ligand is less strongly bonded than a CO ligand (by 6.4 kcal/mol), which in turn suggests relatively weak association in the μ3carbide complex. Yet, one conspicuous aspect of the chemistry of Fe2(CO)9,10 and by analogy the chemistry of Co2(CO)8,11 is the prevalence of bridged bimetallic analog complexes, i.e., Fe2(CO)8(μ-L) and Co2(CO)7(μ-L). The L ligands in these complexes span a wide range of the periodic table, including H, Pt, Au, Cd, B, Ga, In, Tl, C, Si, Ge, Sn, Pb, Sb, and Bi. Intrigued by the lack of experimental evidence to support μ3 coordination with RuC, we surmised that Fe2(CO)9 and Co2(CO)8 would represent attractive platforms for observing this bonding mode.



RESULTS AND DISCUSSION Given its π-accepting abilities,6b,c,9,12 we assumed RuC to be suited for coordination to low-valent first-row transition metals. Specifically, iron forms numerous homoleptic carbonyl complexes, which are both stable and readily available. Reaction of RuC with one equivalent of Fe2(CO)9 affords (Cy3P)2Cl2RuCFe2(CO)8 (RuCFe2, Scheme 1). From NMR spectroscopy, the appearance of a single carbide 13C resonance (438.1 ppm) as well as a single 31P resonance attests to clean adduct formation. The bridging carbide ligand couples with two equivalent PCy3 ligands (J = 5.5 Hz), indicating the integrity of the RuC unit upon metalation. The carbide 13C Received: November 17, 2018

A

DOI: 10.1021/acs.inorgchem.8b03222 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Co2(CO)8. Indeed, one equivalent of Co2(CO)8 reacts with RuC, producing (Cy3P)2Cl2RuCCo2(CO)7 (RuCCo2, Scheme 2) as the dominating carbide-bridged product. The

Scheme 1. Formation and Reactivity of RuCFe2

Scheme 2. Formation of RuCCo2 and (RuCCo)2

resonance falls very downfield for (Cy3P)2Cl2RuC−[M] complexes6a−d (326−446 ppm), indicating competition for πbackbonding between RuC and the carbonyl ligands. During early stages of conversion, 57Fe Mössbauer spectra only display resonances from RuCFe2 and Fe2(CO)9, corroborating the clean transformation of the diiron complex. The appearance of a single new Mössbauer resonance further indicates that the iron atoms from Fe2(CO)9 remain chemically equivalent upon incorporation into RuCFe2. The quadrupole splitting of RuCFe2 is large compared to that of Fe2(CO)9 (2.35(1) versus 0.40(1) mm s−1) in line with the decrease in symmetry around iron as RuC replaces a bridging CO ligand (C2v versus D3h). Conversely, the isomer shift of RuCFe2 is smaller than that of Fe2(CO)9 (−0.09(1) versus 0.18(1) mm s−1), which can be interpreted in terms of differences in backbonding, with RuC as a less potent π-backbonding ligand toward the Fe2(CO)8 fragment than a bridging carbonyl ligand or in terms of increased s−d mixing in the trimetallic complex. Attempts at isolating RuCFe2 by crystallization were hampered by decomposition, which also has prevented isolation of the related molybdenum complex, (Cy3P)2Cl2RuC−Mo(CO)5.6a Apparently, decomposition involves migration of iron-bound CO ligands onto Ru and rupture of the RuC bond, as demonstrated by the formation of the prototypical13 all-trans complex, RuCl2(CO)2(PCy3)2 (Figure 1). Ruthenium vinylidene complexes decompose in a

adduct exhibits a carbide 13C resonance at 423.0 ppm, but a minor carbide resonance at 426.7 ppm appears concomitantly. Changing the stoichiometry to one Co2(CO)8 and two equivalents of RuC favors the formation of the 426.7 ppm resonance, but remaining resonances from RuCCo2 and unconverted RuC suggest an equilibrium between the carbide complexes. Accordingly, RuCCo2 can be prepared selectively using excess Co2(CO)8. Given the ability of Co2(CO)8 to form adducts with phosphines, N-heterocyclic carbenes, or isonitriles capping both Co centers,15 the substoichiometric product is likely dimeric [(Cy3P)2Cl2RuC−Co(CO)3]2, (RuCCo)2. At room temperature, solutions of RuCFe2 and RuCCo2 decompose over the course of hours, hampering isolation and solid-state characterization. NMR and Mössbauer spectroscopies indicate selectivity and stoichiometry but carry limited structural information, and thus, [RuCM] and [RuCM2] complexes cannot readily be distinguished. IR spectroscopy does not reveal bridging carbonyl ligands for RuCFe2, suggesting a low energetic barrier16 for interconversion between terminal and bridging carbonyls, which parallels the IR spectroscopic behavior of analogous carbene-bridged complexes (Fe2(CO)8(μ-L), L = CH2, CF2).17 The CO ligands in RuCFe2 and RuCCo2 produce only two 13C NMR resonances, which do not display couplings to 13C-labeled carbide ligands, suggesting rapid interconversion at the NMR time scale as well. However, Ru and Fe K-edge extended X-ray absorption fine structure (EXAFS) experiments delineate the local structures around the metal centers (Figure 2). Fitting of alternative models (Supporting Information) to the RuCFe2 EXAFS data yields a C2v symmetric structural model. From the Ru K-edge, the four Ru−X separations (2.38(2) Å, X = Cl, P) closely match bond distances in (Cy3P)2Cl2RuC−[M] complexes,6a−d supporting the integrity of the RuC unit observed by NMR spectroscopy. By contrast, the large Ru−C separation (1.71(2) Å) suggests the transition from a triple bond toward a double bond (RuCM bonds: 1.622(7)−1.699(9) Å). Interestingly, the Ru−Fe scattering path involves two iron atoms at identical separation from Ru (3.61(2) Å), establishing the replacement of a bridging CO in Fe2(CO)9. Thereby, the experimental data accord with the structural prediction put forward by Krapp and Frenking.9 Metrics from Fe EXAFS complement the Ru data well, as directly illustrated by the Fe−Ru separation being 3.62(2) Å. Scattering paths from Fe to three independent C atoms and two independent O atoms allow differentiation between terminal and bridging CO ligands (Table 1) as well as the bridging carbide ligand (Fe−C = 2.14(2) Å). When the local structures around iron are compared, the Fe−CO distances are in close agreement between the EXAFS (RuCFe2) and X-ray crystallographically18 (Fe2(CO)9) derived structures, whereas the Fe−O separations are larger in RuCFe2

Figure 1. Molecular structure of RuCl2(CO)2(PCy3)2 with displacement ellipsoids at 50% probability. Hydrogen atoms and disordered parts are omitted. Bond distances (Å): Ru−P: 2.4547(5), Ru−Cl: 2.4096(8), Ru−C: 1.939(4), C−O: 1.118(5).

similar manner upon exposure to CO.14 An alternative route to RuCFe2 starts from RuC and Fe3(CO)12. This reaction is very slow but provides RuCFe2 as the only carbide-bridged product (13C NMR), illustrating the difficulty of installing RuC in higher carbonyl clusters as well as the lower reactivity of Fe3(CO)12 versus Fe2(CO)9. Furthermore, treatment of RuC with the mononuclear iron carbonyl complex, Fe(CO)5, fails to produce any RuCFe2. To probe the generality of μ3-bridged [RuCM2 ] complexes, we examined reactions between RuC and B

DOI: 10.1021/acs.inorgchem.8b03222 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 2. EXAFS of (a) Ru K-edge data for RuCFe2, (b) Fe K-edge data for RuCFe2, (c) Ru K-edge data for RuCCo2. F: Goodness of fit = [(Σin[ki3(EXAFSabs − EXAFScalc)i])2/n]1/2. Red lines: experimental data, dashed black lines: simulated fits.

and the Fe centers. This suggests that CO is a stronger πaccepting ligand than RuC. Though the EXAFS experiments only model interatomic separations, trigonometric relations produce Ru−C−Fe angles of 139−140° and an Fe−C−Fe angle of 80° when applied to separations within the [RuCFe2] core (Figure 3a). The planar μ3 carbide ligand (angle sum: 358−360°) displays angles closely matching those of bridging carbonyls in Fe2(CO)9. The Fe−Fe separation in RuCFe2 increases drastically compared to Fe2(CO)9, and while computational studies9 predict a modest elongation (0.023 Å), EXAFS reveals a remarkably large elongation (0.24 Å). Thus, the Fe−Fe linkage in RuCFe2 (2.76(2) Å) falls within

Table 1. Metrics (Å and °) for RuCFe2 (from EXAFS) and Fe2(CO)9 (from X-ray Crystallography)18 complex

RuCFe2

Fe2(CO)9

complex

RuCFe2

Fe2(CO)9

Fe−Fe Fe−Ca Fe−Oa

2.76(2) 1.81(2) 3.07(2)

2.523(1) 1.835(3) 2.960

Fe−C−Fe Fe−Cb Fe−Ob

80 1.99(2) 3.22(2)

77.6(1) 2.013(3) 3.006

a

Terminal CO. bBridging CO.

than in Fe2(CO)9 (by 0.11−0.16 Å). Upon going from Fe2(CO)9 to RuCFe2, the CO triple bonds elongate, indicating π-backbonding to become stronger between the CO ligands C

DOI: 10.1021/acs.inorgchem.8b03222 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Scheme 3. (a) Conceivable [RuCFe2] and [RuCFe] Complexes, (b) Schematic Indication of Bonding in RuCFe2, (c) HOMO of RuCFe2, (d) HOMO-4 of RuCFe2

Figure 3. (a) Geometry around the μ3 carbide ligand in RuCFe2. (b) Fe−Fe distances from the Cambridge Structural Database (CSD, v. 1.19), indicating intermetallic separations in RuCFe2 and Fe2(CO)9.

bonding interactions that are responsible for the stability of RuCFe2. Within Hoffmann’s fragment-based approach to molecular orbitals, the highest occupied molecular orbital (HOMO) of Fe2(CO)9 consists of antibonding combinations of iron 3dxz and 3dyz orbitals.23 This e set is stabilized by interacting with π* orbitals from bridging CO ligands. Similarly, RuC possesses energetically low-lying π* orbitals. Consequently, the μ3 binding mode of the carbide ligand in RuCFe2 reflects the stabilization of the Fe2(CO)8 fragment upon π-backdonation to RuC (Scheme 3b). Corroborating this bonding description, density functional theory (DFT)-derived frontier molecular orbitals (at the BP86 level of theory in conjunction with the ZORA-def2-SVP basis set using the ORCA program24) for RuCFe2 reveal a π-symmetric HOMO delocalized across the [RuCFe2] fragment (Scheme 3c). Moreover, bonding in RuCFe2 has large contributions from the HOMO-4 (Scheme 3d), which derives from a RuC π orbital. Qualitatively, the DFT-derived frontier molecular orbitals indicate a possible pathway for the initial steps of the decay of RuCFe2 to RuCl2(CO)2(PCy3)2. The HOMO of RuCFe2 has substantial Ru 4dxz character, while the lowest unoccupied molecular orbital (LUMO) has substantial Ru 4dz2 character. These features allude to an associative mechanism, where CO coordinates opposite of the carbide ligand in RuCFe2 (along z) while engaging in σ-donating and π-backbonding interactions with the ruthenium center, thus weakening the carbideruthenium multiple bond.

the longest decile of Fe−Fe distances in the CSD (Figure 3b). Theoretical studies have dismissed bonding descriptions with metal−metal bonds in Fe2(CO)9.19 Along these lines, Fe2(CO)8(μ-L) complexes exhibit a large range of intermetallic distances (μ-CH2: 2.507(1) Å, μ-CF2: 2.5816(6) Å, μ-[Ge0]: 2.7596(18) Å).17,20 In stoichiometric reactions between RuC and Co2(CO)8, the parallel formation of RuCCo2 and (RuCCo)2 frustrates exclusive generation of each complex. However, the clean formation of RuCCo2 with excess Co2(CO)8 allows EXAFS characterization at the Ru K-edge, but not at the Co K-edge. Qualitatively, RuCCo2 and RuCFe2 display similar Ru EXAFS features (Figure 2), suggesting similar molecular structures. Indeed, the Ru−X (X = Cl, P) and Ru−C separations are statistically indistinguishable between the Co and Fe complexes. More interestingly, the Ru−Co scattering path reveals two cobalt centers at identical separation from ruthenium (3.57(2) Å), slightly shorter than the Ru−Fe separation in RuCFe2. The separation is well beyond Ru−Co bond distances in triangular RuCo2 carbonyl clusters (by 0.85−0.98 Å),21 ruling out Ru−Co bonded triangular structures, which would arise by side-on association between the RuC and Co−Co units. Thereby, EXAFS demonstrates reactions of RuC with Fe2(CO)9 and Co2(CO)8 to provide isostructural trimetallic cores, [RuCFe2] and [RuCCo2], both displaying the unusual μ3 carbide ligand. Naively, the reaction between RuC and Fe2(CO)9 could produce di-iron complexes by replacement of terminal or bridging carbonyl ligands, while release of Fe(CO)5 would produce a monoiron complex akin to unstable22 (porphyrin)FeCFe(CO)4 (Scheme 3a). By evaluation of potential energy surfaces, Krapp and Frenking9 predicted μ3-bridging for the RuC-derived carbide ligand, and this has now been verified experimentally. A qualitative description of the frontier molecular orbitals in RuC and Fe2(CO)9 elucidates the



CONCLUSION The association between RuC and bimetallic iron and cobalt carbonyl complexes affords (Cy 3 P) 2 Cl 2 RuCFe 2 (CO) 8 (RuCFe2) and (Cy3P)2Cl2RuCCo2(CO)7 (RuCCo2). The reactions proceed with replacement of bridging CO ligands and with ultimate retention of the 3d metal−metal D

DOI: 10.1021/acs.inorgchem.8b03222 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry *(J.B.) E-mail: [email protected]. *(K.M.L.) E-mail: [email protected].

associations, while inaugurating the carbide ligand in RuC as a μ3 ligand. This coordination mode is in line with πbackdonation from the Fe2(CO)8 and Co2(CO)7 fragments. Given the prevalence of metal−metal bonding in low-valent early transition metal complexes, the μ3 bonding mode of the carbide ligand in RuC may be anticipated to arise in numerous other complexes, establishing a rational4s route to highnuclearity carbide clusters from simple precursors.



ORCID

Anders Reinholdt: 0000-0001-6637-8338 Samantha N. MacMillan: 0000-0001-6516-1823 Anthony F. Hill: 0000-0003-2167-0604 Ola F. Wendt: 0000-0003-2267-5781 Kyle M. Lancaster: 0000-0001-7296-128X Jesper Bendix: 0000-0003-1255-2868

EXPERIMENTAL SECTION

Author Contributions

Preparation of Compounds. The generation of RuCFe2, RuCCo2, and (RuCCo)2 was carried out under air. Benzene-d6 (99.6% D, Aldrich), dichloromethane-d2 (99.5% D, Aldrich), Co2(CO)8 (TCI), Fe2(CO)9 (Aldrich), and Fe3(CO)12 (Aldrich) were purchased from commercial suppliers and used as received. (Cy3P)2Cl2RuC (RuC) and (Cy3P)2Cl2Ru13C (Ru13C), were prepared according to the procedure by Johnson.3h To aid NMR spectroscopic characterization, Ru13C was used extensively (synthesized using CH3CO213CH13CH2, Sigma-Aldrich, 99% 13C). Caution. Fe2(CO)9 and Co2(CO)8 are toxic and release CO upon reaction with RuC. This is relevant to take into account when upscaling the reactions; the small quantities used below pose no significant risk. Co2(CO)8 is pyrophoric. Though Fe2(CO)9 is not pyrophoric itself, we have observed the formation of pyrophoric iron (black-brown solids) in some aged samples of Fe2(CO)9. (Cy3P)2Cl2RuCFe2(CO)8 (RuCFe2). Ru13C (4.9 mg, 6.6 μmol) and excess Fe2(CO)9 (5.1 mg, 14 μmol) were dissolved in 0.4 mL of CD2Cl2 and analyzed by NMR. 1H NMR, 500 MHz, CD2Cl2, δ (ppm): 3.02−2.54 (m, 6H), 2.37−2.01 (m, 12H), 1.99−1.43 (m, 30H), 1.42−0.94 (m, 18H). 13C NMR, 126 MHz, CD2Cl2, δ (ppm): 438.11 (t, J = 5.5 Hz), 213.31, 211.12, 33.44 (t, J = 8.7 Hz), 30.46, 28.32 (t), 27.03. 31P NMR, 202 MHz, CD2Cl2, δ (ppm): 26.00. (Cy3P)2Cl2RuCCo2(CO)7 (RuCCo2). Ru13C (4.7 mg, 6.3 μmol) and excess Co2(CO)8 (12.1 mg, 35.4 μmol) were dissolved in 0.4 mL of C6D6 and analyzed by NMR. 1H NMR, 500 MHz, C6D6, δ (ppm): 3.09−2.76 (m, 6H), 2.53−2.11 (m, 12H), 2.10−1.48 (m, 30H), 1.44−1.02 (m, 18H). 13C NMR, 126 MHz, C6D6, δ (ppm): 423.01, 201.22, 200.13, 33.27 (t), 30.65, 28.03, 26.75. 31P NMR, 202 MHz, C6D6, δ (ppm): 30.73. [(Cy3P)2Cl2RuC−Co(CO)3]2 ((RuCCo)2). Ru13C (9.2 mg, 12.3 μmol) and Co2(CO)8 (2.1 mg, 6.1 μmol) were dissolved in 0.4 mL of C6D6 and analyzed by NMR. 1H NMR, 500 MHz, C6D6, δ (ppm): 3.15−2.78 (m, 12H), 2.58−2.18 (m, 24H), 2.14−1.54 (m, 60H), 1.52−1.01 (m, 36H). 13C NMR, 126 MHz, C6D6, δ (ppm): 426.70, 201.24, 33.27 (t), 30.66, 28.10, 26.83. 31P NMR, 202 MHz, C6D6, δ (ppm): 29.50.



#

A.R. and S.H.M. contributed equally.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS A.R. thanks Interreg (Ref No. KU-036) for funding. K.M.L. thanks the National Science Foundation (CHE-1454455) for funding. S.H.M. is supported by the National Science Foundation Graduate Research Fellowship Program (DGE1650441). 57Fe Mössbauer data were obtained using an instrument acquired with partial financial support from the U.S. Department of Energy Office of Science (DESC0013997). XAS data were obtained at SSRL, which is supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Contract No. DE-AC02-76SF00515. The SSRL Structural Molecular Biology Program is supported by the Department of Energy’s Office of Biological and Environmental Research, and by NIH/HIGMS (including P41GM103393).



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b03222. Additional experimental details, spectral figures, graphs, and X-ray data for RuCl2(CO)2(PCy3)2 (PDF) Accession Codes

CCDC 1833487 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



REFERENCES

(1) (a) Lancaster, K. M.; Roemelt, M.; Ettenhuber, P.; Hu, Y.; Ribbe, M. W.; Neese, F.; Bergmann, U.; DeBeer, S. X-ray Emission Spectroscopy Evidences a Central Carbon in the Nitrogenase IronMolybdenum Cofactor. Science 2011, 334, 974−977. (b) Spatzal, T.; Aksoyoglu, M.; Zhang, L.; Andrade, S. L. A.; Schleicher, E.; Weber, S.; Rees, D. C.; Einsle, O. Evidence for Interstitial Carbon in Nitrogenase FeMo Cofactor. Science 2011, 334, 940−940. (c) Wiig, J. A.; Hu, Y.; Lee, C. C.; Ribbe, M. W. Radical SAM-Dependent Carbon Insertion into the Nitrogenase M-Cluster. Science 2012, 337, 1672−1675. (d) Lancaster, K. M.; Hu, Y.; Bergmann, U.; Ribbe, M. W.; DeBeer, S. X-ray Spectroscopic Observation of an Interstitial Carbide in NifENBound FeMoco Precursor. J. Am. Chem. Soc. 2013, 135, 610−612. (e) Wiig, J. A.; Lee, C. C.; Hu, Y.; Ribbe, M. W. Tracing the Interstitial Carbide of the Nitrogenase Cofactor during Substrate Turnover. J. Am. Chem. Soc. 2013, 135, 4982−4983. (f) Rees, J. A.; Bjornsson, R.; Schlesier, J.; Sippel, D.; Einsle, O.; DeBeer, S. The Fe− V Cofactor of Vanadium Nitrogenase Contains an Interstitial Carbon Atom. Angew. Chem., Int. Ed. 2015, 54, 13249−13252. (2) Č orić, I.; Holland, P. L. Insight into the Iron−Molybdenum Cofactor of Nitrogenase from Synthetic Iron Complexes with Sulfur, Carbon, and Hydride Ligands. J. Am. Chem. Soc. 2016, 138, 7200− 7211. (3) (a) Peters, J. C.; Odom, A. L.; Cummins, C. C. A terminal molybdenum carbide prepared by methylidyne deprotonation. Chem. Commun. 1997, 1995−1996. (b) Greco, J. B.; Peters, J. C.; Baker, T. A.; Davis, W. M.; Cummins, C. C.; Wu, G. Atomic Carbon as a Terminal Ligand: Studies of a Carbidomolybdenum Anion Featuring Solid-State 13C NMR Data and Proton-Transfer Self-Exchange Kinetics. J. Am. Chem. Soc. 2001, 123, 5003−5013. (c) Agapie, T.; Diaconescu, P. L.; Cummins, C. C. Methine (CH) Transfer via a Chlorine Atom Abstraction/Benzene-Elimination Strategy: Molybdenum Methylidyne Synthesis and Elaboration to a Phosphaisocyanide Complex. J. Am. Chem. Soc. 2002, 124, 2412−2413. (d) Enriquez, A.

AUTHOR INFORMATION

Corresponding Authors

*(A.R.) E-mail: [email protected]. E

DOI: 10.1021/acs.inorgchem.8b03222 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry E.; White, P. S.; Templeton, J. L. Reactions of an Amphoteric Terminal Tungsten Methylidyne Complex. J. Am. Chem. Soc. 2001, 123, 4992−5002. (e) Carlson, R. G.; Gile, M. A.; Heppert, J. A.; Mason, M. H.; Powell, D. R.; Velde, D. V.; Vilain, J. M. The Metathesis-Facilitated Synthesis of Terminal Ruthenium Carbide Complexes: A Unique Carbon Atom Transfer Reaction. J. Am. Chem. Soc. 2002, 124, 1580−1581. (f) Romero, P. E.; Piers, W. E.; McDonald, R. Rapidly Initiating Ruthenium Olefin-Metathesis Catalysts. Angew. Chem., Int. Ed. 2004, 43, 6161−6165. (g) van der Eide, E. F.; Romero, P. E.; Piers, W. E. Generation and Spectroscopic Characterization of Ruthenacyclobutane and Ruthenium Olefin Carbene Intermediates Relevant to Ring Closing Metathesis Catalysis. J. Am. Chem. Soc. 2008, 130, 4485−4491. (h) Caskey, S. R.; Stewart, M. H.; Kivela, J. E.; Sootsman, J. R.; Johnson, M. J. A.; Kampf, J. W. Two Generalizable Routes to Terminal Carbido Complexes. J. Am. Chem. Soc. 2005, 127, 16750−16751. (i) Stewart, M. H.; Johnson, M. J. A.; Kampf, J. W. Terminal Carbido Complexes of Osmium: Synthesis, Structure, and Reactivity Comparison to the Ruthenium Analogues. Organometallics 2007, 26, 5102−5110. (j) Buss, J. A.; Agapie, T. Four-electron deoxygenative reductive coupling of carbon monoxide at a single metal site. Nature 2016, 529, 72−75. (k) Buss, J. A.; Agapie, T. Mechanism of Molybdenum-Mediated Carbon Monoxide Deoxygenation and Coupling: Mono- and Dicarbyne Complexes Precede C−O Bond Cleavage and C−C Bond Formation. J. Am. Chem. Soc. 2016, 138, 16466−16477. (l) Morsing, T. J.; Reinholdt, A.; Sauer, S. P. A.; Bendix, J. Ligand Sphere Conversions in Terminal Carbide Complexes. Organometallics 2016, 35, 100−105. (4) (a) Mansuy, D.; Lecomte, J. P.; Chottard, J. C.; Bartoli, J. F. Formation of a complex with a carbide bridge between two iron atoms from the reaction of (tetraphenylporphyrin)iron(II) with carbon tetraiodide. Inorg. Chem. 1981, 20, 3119−3121. (b) Latesky, S. L.; Selegue, J. P. Preparation and structure of [(Me3CO)3WCRu(CO)2(Cp)], a heteronuclear, μ2-carbide complex. J. Am. Chem. Soc. 1987, 109, 4731−4733. (c) Rossi, G.; Goedken, V. L.; Ercolani, C. μ-Carbido-bridged iron phthalocyanine dimers: synthesis and characterization. J. Chem. Soc., Chem. Commun. 1988, 46−47. (d) Beck, W.; Knauer, W.; Robl, C. Synthesis and Structure of the Novel μ2-Carbido Complex [(TPP)FeCRe(CO)4 Re(CO)5]. Angew. Chem., Int. Ed. Engl. 1990, 29, 318−320. (e) Etienne, M.; White, P. S.; Templeton, J. L. An agostic μ-methyne molybdenumiron complex from protonation of a μ-carbide precursor, Tp′(CO)2MoCFe(CO)2Cp. J. Am. Chem. Soc. 1991, 113, 2324− 2325. (f) Miller, R. L.; Wolczanski, P. T.; Rheingold, A. L. Carbide Formation via Carbon Monoxide Dissociation Across a WW Bond. J. Am. Chem. Soc. 1993, 115, 10422−10423. (g) Caselli, A.; Solari, E.; Scopelliti, R.; Floriani, C. The Stepwise Four- and Six-Electron Reduction of Carbon Monoxide to Oxyalkylidyne, to Carbide and Oxide, Then to Carbide over an Nb−Oxo Surface Modeled by Calix[4]arene. J. Am. Chem. Soc. 2000, 122, 538−539. (h) Hong, S. H.; Day, M. W.; Grubbs, R. H. Decomposition of a Key Intermediate in Ruthenium-Catalyzed Olefin Metathesis Reactions. J. Am. Chem. Soc. 2004, 126, 7414−7415. (i) Cade, I. A.; Hill, A. F.; McQueen, C. M. A. Iridium−Molybdenum Carbido Complex via C−Se Activation of a Selenocarbonyl Ligand: (μ-Se2)[Ir2{CMo(CO)2(Tp*)}2(CO)2(PPh3)2] (Tp* = hydrotris(dimethylpyrazolyl)borate). Organometallics 2009, 28, 6639−6641. (j) Colebatch, A. L.; Cordiner, R. L.; Hill, A. F.; Nguyen, K. T. H. D.; Shang, R.; Willis, A. C. A Bis-Carbyne (Ethanediylidyne) Complex via the Catalytic Demercuration of a Mercury Bis(carbido) Complex. Organometallics 2009, 28, 4394−4399. (k) Takemoto, S.; Morita, H.; Karitani, K.; Fujiwara, H.; Matsuzaka, H. A Bimetallic Ru2Pt Complex Containing a Trigonal-Planar μ3-Carbido Ligand: Formation, Structure, and Reactivity Relevant to the Fischer−Tropsch Process. J. Am. Chem. Soc. 2009, 131, 18026−18027. (l) Hill, A. F.; Sharma, M.; Willis, A. C. Heterodinuclear Bridging Carbido and Phosphoniocarbyne Complexes. Organometallics 2012, 31, 2538−2542. (m) Borren, E. S.; Hill, A. F.; Shang, R.; Sharma, M.; Willis, A. C. A Golden Ring: Molecular Gold Carbido Complexes. J. Am. Chem. Soc. 2013, 135, 4942−4945. (n) Young, R. D.; Hill, A. F.; Cavigliasso, G. E.; Stranger, R. [(μ-

C){Re(CO)2(η-C5H5)}2]: A Surprisingly Simple Bimetallic Carbido Complex. Angew. Chem., Int. Ed. 2013, 52, 3699−3702. (o) Takemoto, S.; Ohata, J.; Umetani, K.; Yamaguchi, M.; Matsuzaka, H. A Diruthenium μ-Carbido Complex That Shows Singlet-Carbene-like Reactivity. J. Am. Chem. Soc. 2014, 136, 15889−15892. (p) Kalläne, S. I.; Braun, T.; Teltewskoi, M.; Braun, B.; Herrmann, R.; Laubenstein, R. Remarkable reactivity of a rhodium(I) boryl complex towards CO2 and CS2: isolation of a carbido complex. Chem. Commun. 2015, 51, 14613−14616. (q) Ahrens, T.; Schmiedecke, B.; Braun, T.; Herrmann, R.; Laubenstein, R. Activation of CS2 and COS at a Rhodium(I) Germyl Complex: Generation of CS and Carbido Complexes. Eur. J. Inorg. Chem. 2017, 2017, 713−722. (r) Barnett, H. J.; Burt, L. K.; Hill, A. F. Simple generation of a dirhodium μ-carbido complex via thiocarbonyl reduction. Dalton Trans. 2018, 47, 9570− 9574. (s) Takemoto, S.; Tsujita, M.; Matsuzaka, H. Diruthenium Carbido Complexes as N-Heterocyclic Carbene Like C-Donor Ligands to Group 11 Metals. Organometallics 2017, 36, 3686−3691. (t) Shoshani, M. M.; Johnson, S. A. Cooperative carbon-atom abstraction from alkenes in the core of a pentanuclear nickel cluster. Nat. Chem. 2017, 9, 1282. (u) Shoshani, M. M.; Semeniuchenko, V.; Johnson, S. A. Dismantling of Vinyl Ethers by Pentanuclear [(iPr3P)Ni]5H6: Facile Cooperative C−O, C−C and C−H Activation Pathways. Chem. - Eur. J. 2018, 24, 14282−14289. (v) Takemoto, S.; Matsuzaka, H. Recent advances in the chemistry of ruthenium carbido complexes. Coord. Chem. Rev. 2012, 256, 574−588. (5) (a) Pap, J. S.; DeBeer George, S.; Berry, J. F. Delocalized Metal− Metal and Metal−Ligand Multiple Bonding in a Linear Ru−RuN Unit: Elongation of a Traditionally Short RuN Bond. Angew. Chem., Int. Ed. 2008, 47, 10102−10105. (b) Kornecki, K. P.; Briones, J. F.; Boyarskikh, V.; Fullilove, F.; Autschbach, J.; Schrote, K. E.; Lancaster, K. M.; Davies, H. M. L.; Berry, J. F. Direct Spectroscopic Characterization of a Transitory Dirhodium Donor-Acceptor Carbene Complex. Science 2013, 342, 351−354. (c) Corcos, A. R.; Pap, J. S.; Yang, T.; Berry, J. F. A Synthetic Oxygen Atom Transfer Photocycle from a Diruthenium Oxyanion Complex. J. Am. Chem. Soc. 2016, 138, 10032−10040. (d) Das, A.; Reibenspies, J. H.; Chen, Y.-S.; Powers, D. C. Direct Characterization of a Reactive Lattice-Confined Ru2 Nitride by Photocrystallography. J. Am. Chem. Soc. 2017, 139, 2912− 2915. (6) (a) Hejl, A.; Trnka, T. M.; Day, M. W.; Grubbs, R. H. Terminal ruthenium carbido complexes as σ-donor ligands. Chem. Commun. 2002, 2524−2525. (b) Reinholdt, A.; Vibenholt, J. E.; Morsing, T. J.; Schau-Magnussen, M.; Reeler, N. E. A.; Bendix, J. Carbide complexes as π-acceptor ligands. Chem. Sci. 2015, 6, 5815−5823. (c) Reinholdt, A.; Herbst, K.; Bendix, J. Delivering carbide ligands to sulfide-rich clusters. Chem. Commun. 2016, 52, 2015−2018. (d) Reinholdt, A.; Bendix, J. Weakening of Carbide−Platinum Bonds as a Probe for Ligand Donor Strengths. Inorg. Chem. 2017, 56, 12492−12497. (e) Reinholdt, A.; Hill, A. F.; Bendix, J. Synthons for carbide complex chemistry. Chem. Commun. 2018, 54, 5708−5711. (f) Reinholdt, A.; Bendix, J.; Hill, A. F.; Manzano, R. A. Confluence of disparate carbido chemistries: [WRuAu2(μ-C)2Cl2(CO)2(PCy3)2(Tp*)]. Dalton Trans. 2018, 47, 14893−14896. (7) (a) Mutoh, Y.; Kozono, N.; Araki, M.; Tsuchida, N.; Takano, K.; Ishii, Y. Ruthenium Seleno- and Tellurocarbonyl Complexes: Selenium and Tellurium Atom Transfer to a Terminal Carbido Ligand. Organometallics 2010, 29, 519−522. (b) Reinholdt, A.; Vosch, T.; Bendix, J. Modification of σ-Donor Properties of Terminal Carbide Ligands Investigated Through Carbide−Iodine Adduct Formation. Angew. Chem., Int. Ed. 2016, 55, 12484−12487. (c) Suzuki, A.; Arai, T.; Ikenaga, K.; Mutoh, Y.; Tsuchida, N.; Saito, S.; Ishii, Y. A ruthenium tellurocarbonyl (CTe) complex with a cyclopentadienyl ligand: systematic studies of a series of chalcogenocarbonyl complexes [CpRuCl(CE)(H2IMes)] (E = O, S, Se, Te). Dalton Trans. 2017, 46, 44−48. (8) Caskey, S. R.; Stewart, M. H.; Johnson, M. J. A.; Kampf, J. W. Carbon−Carbon Bond Formation at a Neutral Terminal Carbido Ligand: Generation of Cyclopropenylidene and Vinylidene Complexes. Angew. Chem., Int. Ed. 2006, 45, 7422−7424. F

DOI: 10.1021/acs.inorgchem.8b03222 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

molecular structures of [Co2{μ-Ge(Me)Co(CO)4}(CO)7] and [Co4(μ4-GeMe)2{μ-Ge(Me)Co(CO)4}(CO)10], an edge-bridged square-bipyramidal Ge2Co4 species. J. Chem. Soc., Dalton Trans. 1991, 1201−1208. (i) Uhl, W.; Keimling, S. U.; Hiller, W.; Neumayer, M. Reaction of the Alkylindium(I) Compound In4[C(SiMe3)3]4 with Octacarbonyldicobalt: Bridging of the Co-Co Bond by one or two In-C(SiMe3)3 Groups. Chem. Ber. 1996, 129, 397−400. (j) Shimoi, M.; Ikubo, S.; Kawano, Y.; Katoh, K.; Ogino, H. Synthesis and Structure of a Dinuclear Cobalt Complex Bridged by Nonsubstituted Borylene−Trimethylphosphine. J. Am. Chem. Soc. 1998, 120, 4222−4223. (k) Berenbaum, A.; Jäkle, F.; Lough, A. J.; Manners, I. Reactions of the Si−H Functionalized Silicon-Bridged [1]Ferrocenophane Fe(η-C5H4)2SiMeH with Dicobalt Octacarbonyl: Ring-Opening Metalization and the Synthesis of a Novel Sila[1]ferrocenophane with a Co(CO)4 Substituent. Organometallics 2001, 20, 834−843. (l) Gallo, V.; Mastrorilli, P.; Nobile, C. F.; Braunstein, P.; Englert, U. Chelating versus bridging bonding modes of Nsubstituted bis(diphenylphosphanyl)amine ligands in Pt complexes and Co2Pt clusters. Dalton Trans. 2006, 2342−2349. (m) Schroeder, M.; Schnepf, A. Ge2Co6(CO)20: Ein Ge-Co-Cluster ausgehend von gelö stem GeBr. Z. Anorg. Allg. Chem. 2007, 633, 938−940. (n) McCrea-Hendrick, M. L.; Caputo, C. A.; Roberts, C. J.; Fettinger, J. C.; Tuononen, H. M.; Power, P. P. Reactions of Terphenyl-Substituted Digallene AriPr4GaGaAriPr4 (AriPr4 = C6H3-2,6(C6H3-2,6-iPr2)2) with Transition Metal Carbonyls and Theoretical Investigation of the Mechanism of Addition. Organometallics 2016, 35, 579−586. (12) Krapp, A.; Pandey, K. K.; Frenking, G. Transition Metal− Carbon Complexes. A Theoretical Study. J. Am. Chem. Soc. 2007, 129, 7596−7610. (13) (a) Cotton, F. A.; Meadows, J. H. Molecular and electronic structures of two 16-electron complexes of tungsten(II): WBr2(CO)2(C7H8) (C7H8 = norbornadiene) and WBr2(CO)2(PPh3)2. Inorg. Chem. 1984, 23, 4688−4693. (b) Robinson, P. D.; Hinckley, C. C.; Ikuo, A. Structure of dibromodicarbonylbis(triphenylphosphine)osmium(II). Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1988, 44, 1491−1492. (c) De Araujo, M. P.; Porcu, O. M.; Batista, A. A.; Oliva, G.; Souza, D.-H. F.; Bonfadini, M.; Nascimento, O. R. A Simple Route for Syntheses of Trihalide-Bridged Carbonyl Diruthenium(II,III) Complexes: Crystal and Molecular Structure of ttt-[RuIICl2(CO)2(PPh3)2],[(CO)(AsPh3)2RuII(μ-Cl3)RuIIICl2(AsPh3)] AND [(CO)(PPh3)2RuII(μBr3)RuIIIBr2(PPh3)], Spectroscopies, Electrochemistry and Properties. J. Coord. Chem. 2001, 54, 81−94. (14) Hill, A. F.; Hulkes, A. G.; White, A. J. P.; Williams, D. J. Selenolatovinylidene Complexes: Metal-Mediated Alkynyl Selenoether Rearrangements. Organometallics 2000, 19, 371−373. (15) (a) Casati, N.; Macchi, P.; Sironi, A. Staggered to Eclipsed Conformational Rearrangement of [Co2(CO)6(PPh3)2] in the Solid State: An X-ray Diffraction Study at High Pressure and Low Temperature. Angew. Chem., Int. Ed. 2005, 44, 7736−7739. (b) van Rensburg, H.; Tooze, R. P.; Foster, D. F.; Slawin, A. M. Z. The Synthesis and X-ray Structure of the First Cobalt Carbonyl−NHC Dimer. Implications for the Use of NHCs in Hydroformylation Catalysis. Inorg. Chem. 2004, 43, 2468−2470. (c) Du, J.; Wang, L.; Xie, M.; Deng, L. A Two-Coordinate Cobalt(II) Imido Complex with NHC Ligation: Synthesis, Structure, and Reactivity. Angew. Chem., Int. Ed. 2015, 54, 12640−12644. (d) Carpenter, A. E.; Wen, I.; Moore, C. E.; Rheingold, A. L.; Figueroa, J. S. [1,1-Co2(CO)6(CNArMes2)2]: A Structural Mimic of the Elusive D2d Isomer of [Co2(CO)8]. Chem. Eur. J. 2013, 19, 10452−10457. (16) Wu, Y.; Luo, Q.; Wang, C.; Li, Q.-s.; Xie, Y.; Bruce King, R. Binuclear methylene and difluoromethylene iron carbonyls: A density functional theory study. Inorg. Chim. Acta 2013, 394, 322−327. (17) (a) Sumner, C. E.; Riley, P. E.; Davis, R. E.; Pettit, R. Synthesis, crystal structure, and chemical reactivity of octacarbonyl-μ-methylenediiron. J. Am. Chem. Soc. 1980, 102, 1752−1754. (b) Petz, W.; Weller, F.; Barthel, A.; Mealli, C.; Reinhold, J. Preparation, Spectroscopic Properties, and Crystal Structures of Fe2(CO)6(μ-CO)(μ-CF2)2,

(9) Krapp, A.; Frenking, G. Carbon Complexes as Electronically and Sterically Tunable Analogues of Carbon Monoxide in Coordination Chemistry. J. Am. Chem. Soc. 2008, 130, 16646−16658. (10) (a) Chin, H. B.; Bau, R. Crystal structure of [(Ph3P)2N]+[HFe2(CO)8]−. Inorg. Chem. 1978, 17, 2314−2317. (b) Lagrone, C. B.; Whitmire, K. H.; Churchill, M. R.; Fettinger, J. C. Synthesis and characterization of an iron carbonyl cluster containing lead: crystal and molecular structure of [Et 4 N] 2 [Pb{Fe(CO)4}2{Fe2(CO)8}]. Inorg. Chem. 1986, 25, 2080−2085. (c) Cassidy, J. M.; Whitmire, K. H. Synthesis and structure of [PPN]2[Tl2Fe6(CO)24]: completion of a series of thallium-iron carbonyl clusters. Inorg. Chem. 1989, 28, 1432−1434. (d) Whitmire, K. H.; Shieh, M.; Cassidy, J. Synthesis, characterization, and reactivity of iron carbonyl clusters containing bismuth or antimony. Crystal structures of isomorphous [Et4N][BiFe3Cr(CO)17] and [Et4N][SbFe3Cr(CO)17] and the ring complex Bi2Fe2(CO)8Me2. Inorg. Chem. 1989, 28, 3164−3170. (e) Cassidy, J. M.; Whitmire, K. H. Structures of [(C2H5)4N]2[In2Fe6(CO)24] and [(C2H5)4N][(2,2′bipyridine)InFe2(CO)8]. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1990, 46, 1781−1785. (f) Rossell, O.; Seco, M.; Jones, P. G. Preparation and crystal structure of (NEt4)[Fe2(CO)8(μAuPPh3)]. Inorg. Chem. 1990, 29, 348−350. (g) Cassidy, J. M.; Whitmire, K. H.; Kook, A. M. Solution dynamics of tin and lead iron carbonyl compounds and the solid state structure of [Et4N]2[Sn{Fe2(CO)8}{Fe(CO)4}2]. J. Organomet. Chem. 1993, 456, 61−70. (h) Guzman-Jimenez, I. Y.; Whitmire, K. H. Bis(tetraethylammonium) tetra-μ-carbonyl-icosacarbonylhexairondithallate (2 Fe-Fe)(8 Fe-Tl)(2-). Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1998, 54, No. IUC9800053. (i) Albano, V. G.; Monari, M.; Demartin, F.; Macchi, P.; Femoni, C.; Iapalucci, M. C.; Longoni, G. Synthesis and chemical behavior of [MFe4(CO)16]n− (M = Au, Zn, Cd, Hg) clusters: X ray structure of [NMe 3 CH 2 Ph] 2 [Au{Fe2(CO)8}2]Cl and [PPh4]2[Cd{Fe2(CO)6(μ-CO)2}2]2CH3CN. Solid State Sci. 1999, 1, 597−606. (j) Femoni, C.; Iapalucci, M. C.; Longoni, G.; Tiozzo, C.; Wolowska, J.; Zacchini, S.; Zazzaroni, E. New Hybrid Semiconductor Materials Based on Viologen Salts of Bimetallic Fe−Pt and Fe−Au Carbonyl Clusters: First Structural Characterization of the Diradical π-Dimer of the Diethylviologen Monocation and EPR Evidence of its Triplet State. Chem. - Eur. J. 2007, 13, 6544−6554. (11) (a) Mills, O. S.; Robinson, G. Carbon compounds of the transition metals, VII; The structure of a γ-lactone bridged cobalt carbonyl (triclinic modification). Inorg. Chim. Acta 1967, 1, 61−64. (b) Ball, R. D.; Hall, D. The crystal structure of bis(2,4pentanedionato)(heptacarbonyldicobalt)tin(IV). J. Organomet. Chem. 1973, 56, 209−213. (c) Duffy, D. N.; Mackay, K. M.; Nicholson, B. K.; Robinson, W. T. Preparation and the crystal and molecular structure of [NEt4][Ge{Co2(CO)7}{Co2(CO)6[HgCo(CO)4]}]: an anion containing a mercury-bridged cobalt−cobalt bond. J. Chem. Soc., Dalton Trans. 1981, 381−384. (d) Gerlach, R. F.; Mackay, K. M.; Nicholson, B. K.; Robinson, W. T. Transition-metal carbonyl derivatives of the germanes. Part 13. Preparation, spectroscopic properties, and the crystal and molecular structure of bis[μ-carbonyl-bis(tricarbonylcobaltio)(Co−Co)]germanium (4Co− Ge), [Ge{Co2(CO)7}2]; a new type of group 4−tetracobalt species. J. Chem. Soc., Dalton Trans. 1981, 80−84. (e) Croft, R. A.; Duffy, D. N.; Nicholson, B. K. Anionic germanium polycobalt carbonyl clusters. Part 2. Preparation and structure of [NEt4][Ge{Co5(CO)16}], a cluster which contains five-co-ordinate germanium. J. Chem. Soc., Dalton Trans. 1982, 1023−1027. (f) Mackay, K. M.; Nicholson, B. K.; Sims, A. W.; Tan, C. C. Structure of bis[μ-carbonyl-bis(tricarbonylcobaltio)(Co-Co)]silicon(4 Co-Si). Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1987, 43, 633−635. (g) Adams, R. D.; Chen, G.; Wu, W.; Yin, J. Cluster synthesis. 28. New platinum-cobalt carbonyl cluster complexes and products obtained from their reactions with alkynes. Inorg. Chem. 1990, 29, 4208−4214. (h) Anema, S. G.; Lee, S. K.; Mackay, K. M.; Nicholson, B. K.; Service, M. A further study of the reaction of methylgermane with [Co2(CO)8], and some interconversions of the products. Crystal and G

DOI: 10.1021/acs.inorgchem.8b03222 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Fe2(CO)6(μ-CO)2(μ-CF2), and Fe2(CO)6(μ-CF2)(PPh3)2 − Theoretical Studies of Methylenic vs. Carbonyl Bridges in Diiron Complexes. Z. Anorg. Allg. Chem. 2001, 627, 1859−1869. (18) Cotton, F. A.; Troup, J. M. Accurate determination of a classic structure in the metal carbonyl field: nonacarbonyldi-iron. J. Chem. Soc., Dalton Trans. 1974, 800−802. (19) (a) Green, J. C.; Green, M. L. H.; Parkin, G. The occurrence and representation of three-centre two-electron bonds in covalent inorganic compounds. Chem. Commun. 2012, 48, 11481−11503. (b) Ding, S.; Hall, M. B. The Rich Structural Chemistry Displayed by the Carbon Monoxide as a Ligand to Metal Complexes. Struct. Bonding (Berlin, Ger.) 2016, 169, 199−248. (c) Pan, S.; Zhao, L.; Dias, H. V. R.; Frenking, G. Bonding in Binuclear Carbonyl Complexes M2(CO)9 (M = Fe, Ru, Os). Inorg. Chem. 2018, 57, 7780−7791. (20) Mandal, D.; Dhara, D.; Maiti, A.; Klemmer, L.; Huch, V.; Zimmer, M.; Rzepa, H. S.; Scheschkewitz, D.; Jana, A. Mono- and Dicoordinate Germanium(0) as a Four-Electron Donor. Chem. - Eur. J. 2018, 24, 2873−2878. (21) (a) Bernhardt, W.; Vahrenkamp, H. A “Cluster-Centered” Acetylene-Vinylidene Rearrangement. Angew. Chem., Int. Ed. Engl. 1984, 23, 141−142. (b) Roland, E.; Vahrenkamp, H. Zwei neue Metallcarbonyle: Darstellung und Struktur von RuCo2(CO)11 und Ru2Co2(CO)13. Chem. Ber. 1985, 118, 1133−1142. (c) Braunstein, P.; Graiff, C.; Massera, C.; Predieri, G.; Rosé, J.; Tiripicchio, A. Reactivity of the Heterometallic Clusters [HMCo3(CO)12] and [Et4N][MCo3(CO)12] (M = Fe, Ru) toward Phosphine Selenides, Including Selenium Transfer. Crystal Structures of [HRuCo3(CO)7(μ-CO)3(μdppy)], [MCo 2 (μ 3 -Se)(CO) 7 (μ-dppy)], and [RuCo 2 (μ 3 -Se)(CO)7(μ-dppm)] [dppy = Ph2(2-C5H4N)P, dppm = Ph2PCH2PPh2]. Inorg. Chem. 2002, 41, 1372−1382. (22) Knauer, W.; Beck, W. Carbid-verbrückte Komplexe [HB(pz) 3 (OC) 2 MoC-Pt(PPh 3 ) 2 Br], [(TPP)FeC-M(CO) 4 -M(CO)5] (M = Mn, Re), [(TPP)FeCCr(CO)5] und (TPP)FeCFe(CO)4] (pz = 3,5-dimethylpyrazol-1-yl; TPP = Tetraphenylporphyrinat) aus Halogeno-Carbin und -Carben-Komplexen. Z. Anorg. Allg. Chem. 2008, 634, 2241−2245. (23) Lauher, J. W.; Elian, M.; Summerville, R. H.; Hoffmann, R. Triple-decker sandwiches. J. Am. Chem. Soc. 1976, 98, 3219−3224. (24) Neese, F.; Becker, U.; Ganyushin, D.; Hansen, A.; Liakos, D. G.; Kollmar, C.; Kossmann, S.; Petrenko, T.; Reimann, C.; Riplinger, C.; Sivalingam, K.; Valeev, E.; Wezisla, B.; Wennmohs, F. ORCA: An ab initio, DFT, and Semiempirical Electronic Structure Package, version 4.0.0.2; University of Bonn: Bonn, Germany, 2010.

H

DOI: 10.1021/acs.inorgchem.8b03222 Inorg. Chem. XXXX, XXX, XXX−XXX