7 research outputs found

    Effect of Pyramidalization of the M<sub>2</sub>(SR)<sub>2</sub> Center: The Case of (C<sub>5</sub>H<sub>5</sub>)<sub>2</sub>Ni<sub>2</sub>(SR)<sub>2</sub>

    No full text
    The effects of simple thiolates vs chelating dithiolates on Mā€“M bonding, redox potentials, and synthetic outcomes have been probed experimentally and computationally. Nickelocene (Cp<sub>2</sub>Ni) has long been known to react with simple thiols to give diamagnetic Cp<sub>2</sub>Ni<sub>2</sub>(SR)<sub>2</sub> with planar Ni<sub>2</sub>S<sub>2</sub> cores and long Ni-ā€Æ-ā€Æ-Ni distances. Ethane- and propanedithiol (edtH<sub>2</sub> and pdtH<sub>2</sub>, respectively) instead give di-, tri-, and pentanickel complexes, with nonplanar Ni<sub>2</sub>S<sub>2</sub> cores. The 36e Cp<sub>2</sub>Ni<sub>2</sub>(pdt) (<b>1</b><sup><b>pdt</b></sup>) adopts a symmetrical butterfly Ni<sub>2</sub>S<sub>2</sub> structure. Variable-temperature NMR spectra indicate that <b>1</b><sup><b>pdt</b></sup> possesses a thermally accessible triplet state (Ī”<i>G</i> = 2.65(5) kcal/mol) in equilibrium with a diamagnetic ground state. DFT calculations indicate that the singletā€“triplet gap is highly sensitive to the nonplanarity of the Ni<sub>2</sub>S<sub>2</sub> core. The calculations further reveal that only the high-spin form of <b>1</b><sup><b>pdt</b></sup> features Niā€“Ni bonding, which is unprecedented. Cp<sub>2</sub>Ni<sub>3</sub>(pdt)<sub>2</sub> (<b>2</b><sup><b>pdt</b></sup>), which derives from <b>1</b><sup><b>pdt</b></sup>, crystallized as cis and trans isomers, both with a central NiĀ­(pdt)<sub>2</sub><sup>2ā€“</sup> unit that is S,S-chelated to two CpNi<sup>+</sup> centers. Reaction of Cp<sub>2</sub>Ni with 1,2-ethanedithiol (H<sub>2</sub>edt) and 1,2-benzenedithiol (bdtH<sub>2</sub>) exclusively gave the trinickel species <b>2</b><sup><b>edt</b></sup> and <b>2</b><sup><b>bdt</b></sup>, which are structurally analogous to <i>cis</i>-<b>2</b><sup><b>pdt</b></sup>. Solutions of <b>2</b><sup><b>edt</b></sup> are unstable, depositing crystals of Cp<sub>2</sub>Ni<sub>5</sub>(edt)<sub>4</sub> (<b>3</b><sup><b>edt</b></sup>). Cyclic voltammetric studies show that the Ni<sub>2</sub> species oxidize readily to give the mixed-valence cations [<b>1</b><sup><b>pdt</b></sup>]<sup>+</sup> and [<b>1</b><sup><b>edt</b></sup>]<sup>+</sup>. Crystallographic and EPR analyses indicate that these cations are delocalized mixed-valence NiĀ­(II)ā€“NiĀ­(III) species. Oxidation of Cp<sub>2</sub>Ni<sub>2</sub>(SEt)<sub>2</sub>, which features a planar Ni<sub>2</sub>S<sub>2</sub> core, afforded a mixed-valence cation, showing the pyramidal Ni<sub>2</sub>S<sub>2</sub> core observed in [<b>1</b><sup><b>pdt</b></sup>]<sup>0/+</sup> and [<b>1</b><sup><b>edt</b></sup>]<sup>0/+</sup>. Although not obtained from the Cp<sub>2</sub>Ni/H<sub>2</sub>edt reaction, the neutral complex <b>1</b><sup><b>edt</b></sup> was obtained by reduction of [<b>1</b><sup><b>edt</b></sup>]<sup>+</sup>. Variable-temperature NMR measurements and DFT calculations indicate that the triplet is further stabilized in this highly pyramidalized species

    DFT Dissection of the Reduction Step in H<sub>2</sub> Catalytic Production by [FeFe]-Hydrogenase-Inspired Models: Can the Bridging Hydride Become More Reactive Than the Terminal Isomer?

    No full text
    Density functional theory has been used to study diiron dithiolates [HFe<sub>2</sub>(xdt)Ā­(PR<sub>3</sub>)<sub><i>n</i></sub>(CO)<sub>5ā€“<i>n</i></sub>X] (<i>n</i> = 0, 2, 4; R = H, Me, Et; X = CH<sub>3</sub>S<sup>ā€“</sup>, PMe<sub>3</sub>, NHC = 1,3-dimethylimidazol-2-ylidene; xdt = adt, pdt; adt = azadithiolate; pdt = propanedithiolate). These species are related to the [FeFe]-hydrogenases catalyzing the 2H<sup>+</sup> + 2e<sup>ā€“</sup> ā†” H<sub>2</sub> reaction. Our study is focused on the reduction step following protonation of the Fe<sub>2</sub>(SR)<sub>2</sub> core. FeĀ­(H)Ā­s detected in solution are terminal (t-H) and bridging (Ī¼-H) hydrides. Although unstable versus Ī¼-Hs, synthetic t-Hs feature milder reduction potentials than Ī¼-Hs. Accordingly, attempts were previously made to hinder the isomerization of t-H to Ī¼-H. Herein, we present another strategy: in place of preventing isomerization, Ī¼-H could be made a stronger oxidant than t-H (<i>E</i>Ā°<sub>Ī¼ā€‘H</sub> > <i>E</i>Ā°<sub>tā€‘H</sub>). The nature and number of PR<sub>3</sub> unusually affect Ī”<i>E</i>Ā°<sub>tā€‘Hāˆ’Ī¼ā€‘H</sub>: 4PEt<sub>3</sub> models feature a Ī¼-H with a milder <i>E</i>Ā° than t-H, whereas the 4PMe<sub>3</sub> analogues behave oppositely. The correlation Ī”<i>E</i>Ā°<sub>tā€‘Hāˆ’Ī¼ā€‘H</sub> ā†” stereoelectronic features arises from the steric strain induced by bulky Et groups in 4PEt<sub>3</sub> derivatives. One-electron reduction alleviates intramolecular repulsions only in Ī¼-H species, which is reflected in the loss of bridging coordination. Conversely, in t-H, the strain is retained because a bridging CO holds together the Fe<sub>2</sub> core. That implies that <i>E</i>Ā°<sub>Ī¼ā€‘H</sub> > <i>E</i>Ā°<sub>tā€‘H</sub> in 4-PEt<sub>3</sub> species but not in 4PMe<sub>3</sub> analogues. Also determinant to observe <i>E</i>Ā°<sub>Ī¼ā€‘H</sub> > <i>E</i>Ā°<sub>tā€‘H</sub> is the presence of a Fe apical Ļƒ-donor because its replacement with a CO yields <i>E</i>Ā°<sub>Ī¼ā€‘H</sub> < <i>E</i>Ā°<sub>tā€‘H</sub> even in 4PEt<sub>3</sub> species. Variants with neutral NHC and PMe<sub>3</sub> in place of CH<sub>3</sub>S<sup>ā€“</sup> still feature <i>E</i>Ā°<sub>Ī¼ā€‘H</sub> > <i>E</i>Ā°<sub>tā€‘H</sub>. Replacing pdt with (Hadt)<sup>+</sup> lowers <i>E</i>Ā° but yields <i>E</i>Ā°<sub>Ī¼ā€‘H</sub> < <i>E</i>Ā°<sub>tā€‘H</sub>, indicating that Ī¼-H activation can occur to the detriment of the overpotential increase. In conclusion, our results indicate that the electron richness of the Fe<sub>2</sub> core influences Ī”<i>E</i>Ā°<sub>tā€‘Hāˆ’Ī¼ā€‘H</sub>, provided that (i) the R size of PR<sub>3</sub> must be greater than that of Me and (ii) an electron donor must be bound to Fe apically

    Electron-Rich, Diiron Bis(monothiolato) Carbonyls: Cā€“S Bond Homolysis in a Mixed Valence Diiron Dithiolate

    No full text
    The synthesis and redox properties are presented for the electron-rich bisĀ­(monothiolate)Ā­s Fe<sub>2</sub>(SR)<sub>2</sub>Ā­(CO)<sub>2</sub>Ā­(dppv)<sub>2</sub> for R = Me ([<b>1</b>]<sup>0</sup>), Ph ([<b>2</b>]<sup>0</sup>), CH<sub>2</sub>Ph ([<b>3</b>]<sup>0</sup>). Whereas related derivatives adopt <i>C</i><sub>2</sub>-symmetric Fe<sub>2</sub>(CO)<sub>2</sub>P<sub>4</sub> cores, [<b>1</b>]<sup>0</sup>ā€“[<b>3</b>]<sup>0</sup> have <i>C</i><sub>s</sub> symmetry resulting from the unsymmetrical steric properties of the axial vs equatorial R groups. Complexes [<b>1</b>]<sup>0</sup>ā€“[<b>3</b>]<sup>0</sup> undergo 1e<sup>ā€“</sup> oxidation upon treatment with ferrocenium salts to give the mixed valence cations [Fe<sub>2</sub>(SR)<sub>2</sub>Ā­(CO)<sub>2</sub>Ā­(dppv)<sub>2</sub>]<sup>+</sup>. As established crystallographically, [<b>3</b>]<sup>+</sup> adopts a rotated structure, characteristic of related mixed valence diiron complexes. Unlike [<b>1</b>]<sup>+</sup> and [<b>2</b>]<sup>+</sup> and many other [Fe<sub>2</sub>Ā­(SR)<sub>2</sub>L<sub>6</sub>]<sup>+</sup> derivatives, [<b>3</b>]<sup>+</sup> undergoes Cā€“S bond homolysis, affording the diferrous sulfido-thiolate [Fe<sub>2</sub>Ā­(SCH<sub>2</sub>Ph)Ā­(S)Ā­(CO)<sub>2</sub>Ā­(dppv)<sub>2</sub>]<sup>+</sup> ([<b>4</b>]<sup>+</sup>). According to X-ray crystallography, the first coordination spheres of [<b>3</b>]<sup>+</sup> and [<b>4</b>]<sup>+</sup> are similar, but the Feā€“sulfido bonds are short in [<b>4</b>]<sup>+</sup>. The conversion of [<b>3</b>]<sup>+</sup> to [<b>4</b>]<sup>+</sup> follows first-order kinetics, with <i>k</i> = 2.3 Ɨ 10<sup>ā€“6</sup> s<sup>ā€“1</sup> (30 Ā°C). When the conversion is conducted in THF, the organic products are toluene and dibenzyl. In the presence of TEMPO, the conversion of [<b>3</b>]<sup>+</sup> to [<b>4</b>]<sup>+</sup> is accelerated about 10Ɨ, the main organic product being TEMPO-CH<sub>2</sub>Ph. DFT calculations predict that the homolysis of a Cā€“S bond is exergonic for [Fe<sub>2</sub>Ā­(SCH<sub>2</sub>Ph)<sub>2</sub>Ā­(CO)<sub>2</sub>Ā­(PR<sub>3</sub>)<sub>4</sub>]<sup>+</sup> but endergonic for the neutral complex as well as less substituted cations. The unsaturated character of [<b>4</b>]<sup>+</sup> is indicated by its double carbonylation to give [Fe<sub>2</sub>Ā­(SCH<sub>2</sub>Ph)Ā­(S)Ā­(CO)<sub>4</sub>Ā­(dppv)<sub>2</sub>]<sup>+</sup> ([<b>5</b>]<sup>+</sup>), which adopts a bioctahedral structure

    Preparation and Protonation of Fe<sub>2</sub>(pdt)(CNR)<sub>6</sub>, Electron-Rich Analogues of Fe<sub>2</sub>(pdt)(CO)<sub>6</sub>

    No full text
    The complexes Fe<sub>2</sub>(pdt)Ā­(CNR)<sub>6</sub> (pdt<sup>2ā€“</sup> = CH<sub>2</sub>(CH<sub>2</sub>S<sup>ā€“</sup>)<sub>2</sub>) were prepared by thermal substitution of the hexacarbonyl complex with the isocyanides RNC for R = C<sub>6</sub>H<sub>4</sub>-4-OMe (<b>1</b>), C<sub>6</sub>H<sub>4</sub>-4-Cl (<b>2</b>), Me (<b>3</b>). These complexes represent electron-rich analogues of the parent Fe<sub>2</sub>(pdt)Ā­(CO)<sub>6</sub>. Unlike most substituted derivatives of Fe<sub>2</sub>(pdt)Ā­(CO)<sub>6</sub>, these isocyanide complexes are sterically unencumbered and have the same idealized symmetry as the parent hexacarbonyl derivatives. Like the hexacarbonyls, the stereodynamics of <b>1</b>ā€“<b>3</b> involve both turnstile rotation of the FeĀ­(CNR)<sub>3</sub> as well as the inversion of the chair conformation of the pdt ligand. Structural studies indicate that the basal isocyanide has nonlinear CNC bonds and short Feā€“C distances, indicating that they engage in stronger Feā€“C Ļ€-backbonding than the apical ligands. Cyclic voltammetry reveals that these new complexes are far more reducing than the hexacarbonyls, although the redox behavior is complex. Estimated reduction potentials are <i>E</i><sub>1/2</sub> ā‰ˆ āˆ’0.6 ([<b>2</b>]<sup>+/0</sup>), āˆ’0.7 ([<b>1</b>]<sup>+/0</sup>), and āˆ’1.25 ([<b>3</b>]<sup>+/0</sup>). According to DFT calculations, the rotated isomer of <b>3</b> is only 2.2 kcal/mol higher in energy than the crystallographically observed unrotated structure. The effects of rotated versus unrotated structure and of solvent coordination (THF, MeCN) on redox potentials were assessed computationally. These factors shift the redox couple by as much as 0.25 V, usually less. Compounds <b>1</b> and <b>2</b> protonate with strong acids to give the expected Ī¼-hydrides [H<b>1</b>]<sup>+</sup> and [H<b>2</b>]<sup>+</sup>. In contrast, <b>3</b> protonates with [HNEt<sub>3</sub>]Ā­BAr<sup>F</sup><sub>4</sub> (p<i>K</i><sub>a</sub><sup>MeCN</sup> = 18.7) to give the aminocarbyne [Fe<sub>2</sub>(pdt)Ā­(CNMe)<sub>5</sub>Ā­(Ī¼-CNĀ­(H)Ā­Me)]<sup>+</sup> ([<b>3</b>H]<sup>+</sup>). According to NMR measurements and DFT calculations, this species adopts an unsymmetrical, rotated structure. DFT calculations further indicate that the previously described carbyne complex [Fe<sub>2</sub>(SMe)<sub>2</sub>(CO)<sub>3</sub>Ā­(PMe<sub>3</sub>)<sub>2</sub>(CCF<sub>3</sub>)]<sup>+</sup> also adopts a rotated structure with a bridging carbyne ligand. Complex [<b>3</b>H]<sup>+</sup> reversibly adds MeNC to give [Fe<sub>2</sub>(pdt)Ā­(CNR)<sub>6</sub>(Ī¼-CNĀ­(H)Ā­Me)]<sup>+</sup> ([<b>3</b>HĀ­(CNMe)]<sup>+</sup>). Near room temperature, [<b>3</b>H]<sup>+</sup> isomerizes to the hydride [(Ī¼-H)Ā­Fe<sub>2</sub>Ā­(pdt)Ā­(CNMe)<sub>6</sub>]<sup>+</sup> ([H<b>3</b>]<sup>+</sup>) via a first-order pathway

    Preparation and Protonation of Fe<sub>2</sub>(pdt)(CNR)<sub>6</sub>, Electron-Rich Analogues of Fe<sub>2</sub>(pdt)(CO)<sub>6</sub>

    No full text
    The complexes Fe<sub>2</sub>(pdt)Ā­(CNR)<sub>6</sub> (pdt<sup>2ā€“</sup> = CH<sub>2</sub>(CH<sub>2</sub>S<sup>ā€“</sup>)<sub>2</sub>) were prepared by thermal substitution of the hexacarbonyl complex with the isocyanides RNC for R = C<sub>6</sub>H<sub>4</sub>-4-OMe (<b>1</b>), C<sub>6</sub>H<sub>4</sub>-4-Cl (<b>2</b>), Me (<b>3</b>). These complexes represent electron-rich analogues of the parent Fe<sub>2</sub>(pdt)Ā­(CO)<sub>6</sub>. Unlike most substituted derivatives of Fe<sub>2</sub>(pdt)Ā­(CO)<sub>6</sub>, these isocyanide complexes are sterically unencumbered and have the same idealized symmetry as the parent hexacarbonyl derivatives. Like the hexacarbonyls, the stereodynamics of <b>1</b>ā€“<b>3</b> involve both turnstile rotation of the FeĀ­(CNR)<sub>3</sub> as well as the inversion of the chair conformation of the pdt ligand. Structural studies indicate that the basal isocyanide has nonlinear CNC bonds and short Feā€“C distances, indicating that they engage in stronger Feā€“C Ļ€-backbonding than the apical ligands. Cyclic voltammetry reveals that these new complexes are far more reducing than the hexacarbonyls, although the redox behavior is complex. Estimated reduction potentials are <i>E</i><sub>1/2</sub> ā‰ˆ āˆ’0.6 ([<b>2</b>]<sup>+/0</sup>), āˆ’0.7 ([<b>1</b>]<sup>+/0</sup>), and āˆ’1.25 ([<b>3</b>]<sup>+/0</sup>). According to DFT calculations, the rotated isomer of <b>3</b> is only 2.2 kcal/mol higher in energy than the crystallographically observed unrotated structure. The effects of rotated versus unrotated structure and of solvent coordination (THF, MeCN) on redox potentials were assessed computationally. These factors shift the redox couple by as much as 0.25 V, usually less. Compounds <b>1</b> and <b>2</b> protonate with strong acids to give the expected Ī¼-hydrides [H<b>1</b>]<sup>+</sup> and [H<b>2</b>]<sup>+</sup>. In contrast, <b>3</b> protonates with [HNEt<sub>3</sub>]Ā­BAr<sup>F</sup><sub>4</sub> (p<i>K</i><sub>a</sub><sup>MeCN</sup> = 18.7) to give the aminocarbyne [Fe<sub>2</sub>(pdt)Ā­(CNMe)<sub>5</sub>Ā­(Ī¼-CNĀ­(H)Ā­Me)]<sup>+</sup> ([<b>3</b>H]<sup>+</sup>). According to NMR measurements and DFT calculations, this species adopts an unsymmetrical, rotated structure. DFT calculations further indicate that the previously described carbyne complex [Fe<sub>2</sub>(SMe)<sub>2</sub>(CO)<sub>3</sub>Ā­(PMe<sub>3</sub>)<sub>2</sub>(CCF<sub>3</sub>)]<sup>+</sup> also adopts a rotated structure with a bridging carbyne ligand. Complex [<b>3</b>H]<sup>+</sup> reversibly adds MeNC to give [Fe<sub>2</sub>(pdt)Ā­(CNR)<sub>6</sub>(Ī¼-CNĀ­(H)Ā­Me)]<sup>+</sup> ([<b>3</b>HĀ­(CNMe)]<sup>+</sup>). Near room temperature, [<b>3</b>H]<sup>+</sup> isomerizes to the hydride [(Ī¼-H)Ā­Fe<sub>2</sub>Ā­(pdt)Ā­(CNMe)<sub>6</sub>]<sup>+</sup> ([H<b>3</b>]<sup>+</sup>) via a first-order pathway

    Mechanistic Insight into Electrocatalytic H<sub>2</sub> Production by [Fe<sub>2</sub>(CN){Ī¼-CN(Me)<sub>2</sub>}(Ī¼-CO)(CO)(Cp)<sub>2</sub>]: Effects of Dithiolate Replacement in [FeFe] Hydrogenase Models

    No full text
    DFT has been used to investigate viable mechanisms of the hydrogen evolution reaction (HER) electrocatalyzed by [Fe<sub>2</sub>(CN)Ā­{Ī¼-CNĀ­(Me)<sub>2</sub>}Ā­(Ī¼-CO)Ā­(CO)Ā­(Cp)<sub>2</sub>] (<b>1</b>) in AcOH. Molecular details underlying the proposed ECEC electrochemical sequence have been studied, and the key functionalities of CN<sup>ā€“</sup> and amino-carbyne ligands have been elucidated. After the first reduction, CN<sup>ā€“</sup> works as a relay for the first proton from AcOH to the carbyne, with this ligand serving as the main electron acceptor for both reduction steps. After the second reduction, a second protonation occurs at CN<sup>ā€“</sup> that forms a FeĀ­(CNH) moiety: i.e., the acidic source for the H<sub>2</sub> generation. The hydride (formally 2e/H<sup>+</sup>), necessary to the heterocoupling with H<sup>+</sup> is thus provided by the Ī¼-CNĀ­(Me)<sub>2</sub> ligand and not by Fe centers, as occurs in typical L<sub>6</sub>Fe<sub>2</sub>S<sub>2</sub> derivatives modeling the hydrogenase active site. It is remarkable, in this regard, that CN<sup>ā€“</sup> plays a role more subtle than that previously expected (increasing electron density at Fe atoms). In addition, the role of AcOH in shuttling protons from CN<sup>ā€“</sup> to CNĀ­(Me)<sub>2</sub> is highlighted. The incompetence for the HER of the related species [Fe<sub>2</sub>{Ī¼-CNĀ­(Me)<sub>2</sub>}Ā­(Ī¼-CO)Ā­(CO)<sub>2</sub>(Cp)<sub>2</sub>]<sup>+</sup> (<b>2</b><sup><b>+</b></sup>) has been investigated and attributed to the loss of proton responsiveness caused by CN<sup>ā€“</sup> replacement with CO. In the context of hydrogenase mimicry, an implication of this study is that the dithiolate strap, normally present in all synthetic models, can be removed from the Fe<sub>2</sub> core without loss of HER, but the redox and acidā€“base processes underlying turnover switch from a metal-based to a ligand-based chemistry. The versatile nature of the carbyne, once incorporated in the Fe<sub>2</sub> scaffold, could be exploited to develop more active and robust catalysts for the HER

    Imine-Centered Reactions in Imino-Phosphine Complexes of Iron Carbonyls

    No full text
    Fundamental reactions of imino-phosphine ligands were elucidated through studies on Ph<sub>2</sub>PC<sub>6</sub>H<sub>4</sub>CHī—»NC<sub>6</sub>H<sub>4</sub>-4-Cl (PCHNAr<sup>Cl</sup>) complexes of iron(0), ironĀ­(I), and ironĀ­(II). The reaction of PCHNAr<sup>Cl</sup> with FeĀ­(bda)Ā­(CO)<sub>3</sub> gives FeĀ­(PCHNAr<sup>Cl</sup>)Ā­(CO)<sub>3</sub> (<b>1</b>), featuring an Ī·<sup>2</sup>-imine. DNMR studies, its optical properties, and DFT calculations suggest that <b>1</b> racemizes on the NMR time scale via an achiral N-bonded imine intermediate. The <i>N</i>-imine isomer is more stable in FeĀ­(PCHNAr<sup>OMe</sup>)Ā­(CO)<sub>3</sub> (<b>1</b><sup><b>OMe</b></sup>), which crystallized despite being the minor isomer in solution. Protonation of <b>1</b> by HBF<sub>4</sub>Ā·Et<sub>2</sub>O gave the iminium complex [<b>1</b>H]Ā­BF<sub>4</sub>. The related diphosphine complex FeĀ­(PCHNAr<sup>Cl</sup>)Ā­(PMe<sub>3</sub>)Ā­(CO)<sub>2</sub> (<b>2</b>), which features an Ī·<sup>2</sup>-imine, was shown to also undergo N protonation. Oxidation of <b>1</b> and <b>2</b> with FcBF<sub>4</sub> gave the FeĀ­(I) compounds [<b>1</b>]Ā­BF<sub>4</sub> and [<b>2</b>]Ā­BF<sub>4</sub>. The oxidation-induced change in hapticity of the imine from Ī·<sup>2</sup> in [<b>1</b>]<sup>0</sup> to Īŗ<sup>1</sup> in [<b>1</b>]<sup>+</sup> was verified crystallographically. Substitution of a CO ligand in <b>1</b> with PCHNAr<sup>Cl</sup> gave FeĀ­[P<sub>2</sub>(NAr<sup>Cl</sup>)<sub>2</sub>]Ā­(CO)<sub>2</sub> (<b>3</b>), which contains the tetradentate diamidodiphosphine ligand. This Cā€“C coupling is reversed by chemical oxidation of <b>3</b> with FcOTf. The oxidized product of [FeĀ­(PCHNAr<sup>Cl</sup>)<sub>2</sub>(CO)<sub>2</sub>]<sup>2+</sup> ([<b>4</b>]<sup>2+</sup>) was prepared independently by the reaction of [<b>1</b>]<sup>+</sup>, PCHNAr<sup>Cl</sup>, and Fc<sup>+</sup>. The Cā€“C scission is proposed to proceed concomitantly with the reduction of FeĀ­(II) via an intermediate related to [<b>2</b>]<sup>+</sup>
    corecore