44 research outputs found

    Synthesis of Diiron(I) Dithiolato Carbonyl Complexes

    No full text
    Virtually all organosulfur compounds react with Fe(0) carbonyls to give the title complexes. These reactions are reviewed in light of major advances over the past few decades, spurred by interest in Fe<sub>2</sub>(Ī¼-SR)<sub>2</sub>(CO)<sub><i>x</i></sub> centers at the active sites of the [FeFe]-hydrogenase enzymes. The most useful synthetic route to Fe<sub>2</sub>(Ī¼-SR)<sub>2</sub>(CO)<sub>6</sub> involves the reaction of thiols with Fe<sub>2</sub>(CO)<sub>9</sub> and Fe<sub>3</sub>(CO)<sub>12</sub>. Such reactions can proceed via mono-, di-, and triiron intermediates. The reactivity of Fe(0) carbonyls toward thiols is highly chemoselective, and the resulting dithiolato complexes are fairly rugged. Thus, many complexes tolerate further synthetic elaboration directed at the organic substituents. A second major route involves alkylation of Fe<sub>2</sub>(Ī¼-S<sub>2</sub>)Ā­(CO)<sub>6</sub>, Fe<sub>2</sub>(Ī¼-SH)<sub>2</sub>(CO)<sub>6</sub>, and Li<sub>2</sub>Fe<sub>2</sub>(Ī¼-S)<sub>2</sub>(CO)<sub>6</sub>. This approach is especially useful for azadithiolates Fe<sub>2</sub>[(Ī¼-SCH<sub>2</sub>)<sub>2</sub>NR]Ā­(CO)<sub>6</sub>. Elaborate complexes arise via addition of the Fe<i>SH</i> group to electrophilic alkenes, alkynes, and carbonyls. Although the first example of Fe<sub>2</sub>(Ī¼-SR)<sub>2</sub>(CO)<sub>6</sub> was prepared from ferrous reagents, ferrous compounds are infrequently used, although the FeĀ­(II)Ā­(SR)<sub>2</sub> + Fe(0) condensation reaction is promising. Almost invariably low-yielding, the reaction of Fe<sub>3</sub>(CO)<sub>12</sub>, S<sub>8</sub>, and a variety of unsaturated substrates results in Cā€“H activation, affording otherwise inaccessible derivatives. Thiones and related Cī—»S-containing reagents are highly reactive toward Fe(0), often giving complexes derived from substituted methanedithiolates and Cā€“H activation

    Hydrogen Activation by Biomimetic [NiFe]-Hydrogenase Model Containing Protected Cyanide Cofactors

    No full text
    Described are experiments demonstrating incorporation of cyanide cofactors and hydride substrate into [NiFe]-hydrogenase (H<sub>2</sub>ase) active site models. Complexes of the type (CO)<sub>2</sub>(CN)<sub>2</sub>FeĀ­(pdt)Ā­NiĀ­(dxpe) (dxpe = dppe, <b>1</b>; dxpe = dcpe, <b>2</b>) bind the Lewis acid BĀ­(C<sub>6</sub>F<sub>5</sub>)<sub>3</sub> (BAr<sup>F</sup><sub>3</sub>) to give the adducts (CO)<sub>2</sub>(CNBAr<sup>F</sup><sub>3</sub>)<sub>2</sub>Ā­FeĀ­(pdt)Ā­NiĀ­(dxpe), (<b>1</b>(BAr<sup>F</sup><sub>3</sub>)<sub>2</sub>, <b>2</b>(BAr<sup>F</sup><sub>3</sub>)<sub>2</sub>). Upon decarbonylation using amine oxides, these adducts react with H<sub>2</sub> to give hydrido derivatives [(CO)Ā­(CNBAr<sup>F</sup><sub>3</sub>)<sub>2</sub>Ā­FeĀ­(H)Ā­(pdt)Ā­NiĀ­(dxpe)]<sup>āˆ’</sup> (dxpe = dppe, [H<b>3</b>(BAr<sup>F</sup><sub>3</sub>)<sub>2</sub>]<sup>āˆ’</sup>; dxpe = dcpe, [H<b>4</b>(BAr<sup>F</sup><sub>3</sub>)<sub>2</sub>]<sup>āˆ’</sup>). Crystallographic analysis shows that Et<sub>4</sub>NĀ­[H<b>3</b>(BAr<sup>F</sup><sub>3</sub>)<sub>2</sub>] generally resembles the active site of the enzyme in the reduced, hydride-containing states (Niā€“C/R). The Feā€“HĀ·Ā·Ā·Ni center is unsymmetrical with <i>r</i><sub>Feā€“H</sub> = 1.51(3) ƅ and <i>r</i><sub>Niā€“H</sub> = 1.71(3) ƅ. Both crystallographic and <sup>19</sup>F NMR analyses show that the CNBAr<sup>F</sup><sub>3</sub><sup>ā€“</sup> ligands occupy basal and apical sites. Unlike cationic Niā€“Fe hydrides, [H<b>3</b>(BAr<sup>F</sup><sub>3</sub>)<sub>2</sub>]<sup>āˆ’</sup> and [H<b>4</b>(BAr<sup>F</sup><sub>3</sub>)<sub>2</sub>]<sup>āˆ’</sup> oxidize at mild potentials, near the Fc<sup>+/0</sup> couple. Electrochemical measurements indicate that in the presence of base, [H<b>3</b>(BAr<sup>F</sup><sub>3</sub>)<sub>2</sub>]<sup>āˆ’</sup> catalyzes the oxidation of H<sub>2</sub>. NMR evidence indicates dihydrogen bonding between these anionic hydrides and R<sub>3</sub>NH<sup>+</sup> salts, which is relevant to the mechanism of hydrogenogenesis. In the case of Et<sub>4</sub>NĀ­[H<b>3</b>(BAr<sup>F</sup><sub>3</sub>)<sub>2</sub>], strong acids such as HCl induce H<sub>2</sub> release to give the chloride Et<sub>4</sub>NĀ­[(CO)Ā­(CNBAr<sup>F</sup><sub>3</sub>)<sub>2</sub>Ā­FeĀ­(Cl)Ā­(pdt)Ā­NiĀ­(dppe)]

    Connecting [NiFe]- and [FeFe]-Hydrogenases: Mixed-Valence Nickelā€“Iron Dithiolates with Rotated Structures

    No full text
    New mixed-valence ironā€“nickel dithiolates are described that exhibit structures similar to those of mixed-valence diiron dithiolates. The interaction of tricarbonyl salt [(dppe)Ā­NiĀ­(pdt)Ā­FeĀ­(CO)<sub>3</sub>]Ā­BF<sub>4</sub> ([<b>1</b>]Ā­BF<sub>4</sub>, where dppe = Ph<sub>2</sub>PCH<sub>2</sub>CH<sub>2</sub>PPh<sub>2</sub> and pdt<sup>2ā€“</sup> = āˆ’SCH<sub>2</sub>CH<sub>2</sub>CH<sub>2</sub>Sāˆ’) with P-donor ligands (L) afforded the substituted derivatives [(dppe)Ā­NiĀ­(pdt)Ā­FeĀ­(CO)<sub>2</sub>L]Ā­BF<sub>4</sub> incorporating L = PHCy<sub>2</sub> ([<b>1a</b>]Ā­BF<sub>4</sub>), PPhĀ­(NEt<sub>2</sub>)<sub>2</sub> ([<b>1b</b>]Ā­BF<sub>4</sub>), PĀ­(NMe<sub>2</sub>)<sub>3</sub> ([<b>1c</b>]Ā­BF<sub>4</sub>), PĀ­(<i>i</i>-Pr)<sub>3</sub> ([<b>1d</b>]Ā­BF<sub>4</sub>), and PCy<sub>3</sub> ([<b>1e</b>]Ā­BF<sub>4</sub>). The related precursor [(dcpe)Ā­NiĀ­(pdt)Ā­FeĀ­(CO)<sub>3</sub>]Ā­BF<sub>4</sub> ([<b>2</b>]Ā­BF<sub>4</sub>, where dcpe = Cy<sub>2</sub>PCH<sub>2</sub>CH<sub>2</sub>PCy<sub>2</sub>) gave the more electron-rich family of compounds [(dcpe)Ā­NiĀ­(pdt)Ā­FeĀ­(CO)<sub>2</sub>L]Ā­BF<sub>4</sub> for L = PPh<sub>2</sub>(2-pyridyl) ([<b>2a</b>]Ā­BF<sub>4</sub>), PPh<sub>3</sub> ([<b>2b</b>]Ā­BF<sub>4</sub>), and PCy<sub>3</sub> ([<b>2c</b>]Ā­BF<sub>4</sub>). For bulky and strongly basic monophosphorus ligands, the salts feature distorted coordination geometries at iron: crystallographic analyses of [<b>1e</b>]Ā­BF<sub>4</sub> and [<b>2c</b>]Ā­BF<sub>4</sub> showed that they adopt ā€œrotatedā€ Fe<sup>I</sup> centers, in which PCy<sub>3</sub> occupies a basal site and one CO ligand partially bridges the Ni and Fe centers. Like the undistorted mixed-valence derivatives, members of the new class of complexes are described as Ni<sup>II</sup>Fe<sup>I</sup> (<i>S</i> = <sup>1</sup>/<sub>2</sub>) systems according to electron paramagnetic resonance spectroscopy, although with attenuated <sup>31</sup>P hyperfine interactions. Density functional theory calculations using the BP86, B3LYP, and PBE0 exchange-correlation functionals agree with the structural and spectroscopic data, suggesting that the spin for [<b>1e</b>]<sup>+</sup> is mostly localized in a Fe<sup>I</sup>-centered dĀ­(<i>z</i><sup>2</sup>) orbital, orthogonal to the Feā€“P bond. The PCy<sub>3</sub> complexes, rare examples of species featuring ā€œrotatedā€ Fe centers, both structurally and spectroscopically incorporate features from homobimetallic mixed-valence diiron dithiolates. Also, when the NiS<sub>2</sub>Fe core of the [NiFe]-hydrogenase active site is reproduced, the ā€œhybrid modelsā€ incorporate key features of the two major classes of hydrogenase. Furthermore, cyclic voltammetry experiments suggest that the highly basic phosphine ligands enable a second oxidation corresponding to the couple [(dxpe)Ā­NiĀ­(pdt)Ā­FeĀ­(CO)<sub>2</sub>L]<sup>+/2+</sup>. The resulting unsaturated 32e<sup>ā€“</sup> dications represent the closest approach to modeling the highly electrophilic Niā€“SI<sub>a</sub> state. In the case of L = PPh<sub>2</sub> (2-pyridyl), chelation of this ligand accompanies the second oxidation

    Cooperative Metalā€“Ligand Reactivity and Catalysis in Low-Spin Ferrous Alkoxides

    No full text
    This report describes examples of combined Fe- and O-centered reactivity of FeĀ­(P<sub>2</sub>O<sub>2</sub>)Ā­(CO)<sub>2</sub> (<b>1</b>), where P<sub>2</sub>O<sub>2</sub> is the diphosphinoglycolate (Ph<sub>2</sub>PC<sub>6</sub>H<sub>4</sub>CHO)<sub>2</sub><sup>2ā€“</sup>. This 18e low-spin ferrous dialkoxide undergoes substitution of CO to give the labile monosubstituted derivatives FeĀ­(P<sub>2</sub>O<sub>2</sub>)Ā­(CO)Ā­(L) (L = PMe<sub>3</sub>, pyridine, MeCN). Treatment of FeĀ­(P<sub>2</sub>O<sub>2</sub>)Ā­(CO)<sub>2</sub> with BrĆønsted acids results in stepwise O-protonation, affording rare examples of low-spin FeĀ­(II) complexes containing alcohol ligands. Substitution reactions with amides (RCĀ­(O)Ā­NH<sub>2</sub>) proceeds with binding of the carbonyl and formation of an intramolecular hydrogen bond between N<i>H</i> and the neighboring alkoxo ligand. This two-site binding was confirmed with crystallographic characterization of the thiourea-substituted derivative. FeĀ­(P<sub>2</sub>O<sub>2</sub>)Ā­(CO)<sub>2</sub> reacts with Ph<sub>2</sub>SiH<sub>2</sub> to give the O-silylated hydrido complex, which is inactive for hydrosilylation. The monocarbonyl derivatives FeĀ­(P<sub>2</sub>O<sub>2</sub>)Ā­(CO)Ā­(L) (L = NCMe, PMe<sub>3</sub>, acetamide) are precursors to catalysts for the hydrosilylation of benzaldehyde, acetophenone, and styrene

    Insights into the Hydrolytic Polymerization of Trimethoxymethylsilane. Crystal Structure of (MeO)<sub>2</sub>MeSiONa

    No full text
    The commercially practiced conversion of trimethoxymethylsilane (MTM) to [OSiĀ­(OMe)Ā­Me)]<sub><i>n</i></sub> polymers and resins is assumed to proceed via the silanol (MeO)<sub>2</sub>MeSiOH. Access to this crucial silanol is gained via the synthesis of (MeO)<sub>2</sub>MeSiONa, the first methoxysilanoate to be crystallographically characterized. Mild protonation of this silanoate gives (MeO)<sub>2</sub>MeSiOH, which is shown to condense with (MeO)<sub>2</sub>MeSiOH but not with MTM. Condensation generates reactive disiloxanols but does not produce symmetric disiloxanes. Parallel results were obtained with (MeO)<sub>2</sub>PhSiOH

    Phosphine-Iminopyridines as Platforms for Catalytic Hydrofunctionalization of Alkenes

    No full text
    A series of phosphine-diimine ligands were synthesized by the condensation of 2-(diphenylphosphino)Ā­aniline (PNH<sub>2</sub>) with a variety of formyl and ketopyridines. Condensation of PNH<sub>2</sub> with acetyl- and benzoylpyridine yielded the Ph<sub>2</sub>PĀ­(C<sub>6</sub>H<sub>4</sub>)Ā­Nī—»CĀ­(R)Ā­(C<sub>5</sub>H<sub>4</sub>N), respectively abbreviated PN<sup>Me</sup>py and PN<sup>Ph</sup>py. With ferrous halides, PN<sup>Ph</sup>py gave the complexes FeX<sub>2</sub>(PN<sup>Ph</sup>py) (X = Cl, Br). Condensation of pyridine carboxaldehyde and its 6-methyl derivatives with PNH<sub>2</sub> was achieved using a ferrous template, affording low-spin complexes [FeĀ­(PN<sup>H</sup>py<sup>R</sup>)<sub>2</sub>]<sup>2+</sup> (R = H, Me). Dicarbonyls FeĀ­(PN<sup>R</sup>py)Ā­(CO)<sub>2</sub> were produced by treating PN<sup>Me</sup>py with FeĀ­(benzylideneacetone)Ā­(CO)<sub>3</sub> and reduction of FeX<sub>2</sub>(PN<sup>Ph</sup>py) with NaBEt<sub>3</sub>H under a CO atmosphere. Cyclic voltammetric studies show that the [FeL<sub>3</sub>(CO)<sub>2</sub>]<sup>0/ā€“</sup> and [FeL<sub>3</sub>(CO)<sub>2</sub>]<sup>+/0</sup> couples are similar for a range of tridentate ligands, but the PN<sup>Ph</sup>py system uniquely sustains two one-electron reductions. Treatment of FeĀ­(PN<sup>Ph</sup>py)Ā­X<sub>2</sub> with NaBEt<sub>3</sub>H gave active catalysts for the hydroboration of 1-octene with pinacolborane. Similarly, these catalysts proved active for the addition of diphenylsilane, but not HSiMeĀ­(OSiMe<sub>3</sub>)<sub>2</sub>, to 1-octene and vinylsilanes. Evidence is presented that catalysis occurs via iron hydride complexes of intact PN<sup>Ph</sup>py

    Characterization of a Borane Ļƒ Complex of a Diiron Dithiolate: Model for an Elusive Dihydrogen Adduct

    No full text
    The azadithiolate complex Fe<sub>2</sub>[(SCH<sub>2</sub>)<sub>2</sub>NMe]Ā­(CO)<sub>6</sub> reacts with borane to give an initial 1:1 adduct, which spontaneously decarbonylates to give Fe<sub>2</sub>[(SCH<sub>2</sub>)<sub>2</sub>NMeBH<sub>3</sub>]Ā­(CO)<sub>5</sub>. Featuring a Feā€“Hā€“B three-center, two-electron interaction, the pentacarbonyl complex is a structural model for H<sub>2</sub> complexes invoked in the [FeFe]-hydrogenases. The pentacarbonyl compound is a ā€œĻƒ complexā€, where a Bā€“H Ļƒ bond serves as a ligand for iron. The structure of this Ļƒ complex was characterized by variable-temperature NMR spectroscopy and X-ray crystallography. Complementary to the 1:1 borane adduct is the quaternary ammonium complex [Fe<sub>2</sub>[(SCH<sub>2</sub>)<sub>2</sub>NMe<sub>2</sub>]Ā­(CO)<sub>6</sub>]<sup>+</sup>, which was also characterized. It represents a kinetically robust analogue of the N-protonated amine cofactor, as indicated by its mild reduction potential

    Borane-Protected Cyanides as Surrogates of Hā€‘Bonded Cyanides in [FeFe]-Hydrogenase Active Site Models

    No full text
    Triarylborane Lewis acids bind [Fe<sub>2</sub>(pdt)Ā­(CO)<sub>4</sub>(CN)<sub>2</sub>]<sup>2ā€“</sup> [<b>1</b>]<sup>2ā€“</sup> (pdt<sup>2ā€“</sup> = 1,3-propanedithiolate) and [Fe<sub>2</sub>(adt)Ā­(CO)<sub>4</sub>(CN)<sub>2</sub>]<sup>2ā€“</sup> [<b>3</b>]<sup>2ā€“</sup> (adt<sup>2ā€“</sup> = 1,3-azadithiolate, HNĀ­(CH<sub>2</sub>S<sup>ā€“</sup>)<sub>2</sub>) to give the 2:1 adducts [Fe<sub>2</sub>(xdt)Ā­(CO)<sub>4</sub>(CNBAr<sub>3</sub>)<sub>2</sub>]<sup>2ā€“</sup>. Attempts to prepare the 1:1 adducts [<b>1</b>(BAr<sub>3</sub>)]<sup>2ā€“</sup> (Ar = Ph, C<sub>6</sub>F<sub>5</sub>) were unsuccessful, but related 1:1 adducts were obtained using the bulky borane BĀ­(C<sub>6</sub>F<sub>4</sub>-<i>o</i>-C<sub>6</sub>F<sub>5</sub>)<sub>3</sub> (BAr<sup>F</sup>*<sub>3</sub>). By virtue of the N-protection by the borane, salts of [Fe<sub>2</sub>(pdt)Ā­(CO)<sub>4</sub>(CNBAr<sub>3</sub>)<sub>2</sub>]<sup>2ā€“</sup> sustain protonation to give hydrides that are stable (in contrast to [H<b>1</b>]<sup>āˆ’</sup>). The hydrides [H<b>1</b>(BAr<sub>3</sub>)<sub>2</sub>]<sup>āˆ’</sup> are 2.5ā€“5 p<i>K</i><sub>a</sub> units more acidic than the parent [H<b>1</b>]<sup>āˆ’</sup>. The adducts [<b>1</b>(BAr<sub>3</sub>)<sub>2</sub>]<sup>2ā€“</sup> oxidize quasi-reversibly around āˆ’0.3 V versus Fc<sup>0/+</sup> in contrast to ca. āˆ’0.8 V observed for the [<b>1</b>]<sup>2ā€“/ā€“</sup> couple. A simplified synthesis of [<b>1</b>]<sup>2ā€“</sup>, [<b>3</b>]<sup>2ā€“</sup>, and [Fe<sub>2</sub>(pdt)Ā­(CO)<sub>5</sub>(CN)]<sup>āˆ’</sup> ([<b>2</b>]<sup>āˆ’</sup>) was developed, entailing reaction of the diiron hexacarbonyl complexes with KCN in MeCN

    Unsensitized Photochemical Hydrogen Production Catalyzed by Diiron Hydrides

    No full text
    The diiron hydride [(Ī¼-H)Ā­Fe<sub>2</sub>(pdt)Ā­(CO)<sub>4</sub>(dppv)]<sup>+</sup> ([H<b>2</b>]<sup>+</sup>, dppv = <i>cis</i>-1,2-C<sub>2</sub>H<sub>2</sub>(PPh<sub>2</sub>)<sub>2</sub>) is shown to be an effective photocatalyst for the H<sub>2</sub> evolution reaction (HER). These experiments establish the role of hydrides in photocatalysis by biomimetic diiron complexes. Trends in redox potentials suggests that other unsymmetrically substituted diiron hydrides are promising catalysts. Unlike previous catalysts for photo-HER, [H<b>2</b>]<sup>+</sup> functions without sensitizers: irradiation of [H<b>2</b>]<sup>+</sup> in the presence of triflic acid (HOTf) efficiently affords H<sub>2</sub>. Instead of sacrificial electron donors, ferrocenes can be used as recyclable electron donors for the photocatalyzed HER, resulting in 4 turnovers

    <i>C</i><sub>2</sub>ā€‘Symmetric Iron(II) Diphosphineā€“Dialkoxide Dicarbonyl and Related Complexes

    No full text
    Reaction of FeĀ­(bda)Ā­(CO)<sub>3</sub> (bda = benzylideneacetone) and Ph<sub>2</sub>P-2-C<sub>6</sub>H<sub>4</sub>CHO (PCHO) affords the bisĀ­phosphine bisalkoxide complex FeĀ­[(Ph<sub>2</sub>PC<sub>6</sub>H<sub>4</sub>)<sub>2</sub>C<sub>2</sub>H<sub>2</sub>O<sub>2</sub>]Ā­(CO)<sub>2</sub> (<b>1</b>) arising from the head-to-head coupling of two formyl groups concomitant with oxidation of Fe(0) to FeĀ­(II). Crystallographic studies show that <b>1</b> features <i>cis</i> alkoxide ligands that are <i>trans</i> to CO; the two phosphine groups are mutually <i>trans</i> with a Pā€“Feā€“P angle of 167.44(4)Ā°. The pathway leading to <b>1</b> was examined, starting with the adduct FeĀ­(PCHO)Ā­(CO)<sub>4</sub> (<b>2</b>), which was obtained by addition of PCHO to Fe<sub>2</sub>(CO)<sub>9</sub>. Compound <b>2</b> decarbonylates to give tricarbonyl FeĀ­(Īŗ<sup>1</sup>,Ī·<sup>2</sup>-PCHO)Ā­(CO)<sub>3</sub> (<b>3</b>), which features a Ļ€-bonded aldehyde. Photolysis of <b>2</b> gives a mixture of <b>3</b> and isomeric hydride HFeĀ­(Īŗ<sup>2</sup>-PCO)Ā­(CO)<sub>3</sub>. Complex <b>3</b> reacts with an additional equivalent of PCHO to afford <b>1</b>, whereas treatment with PPh<sub>3</sub> afforded the substituted product FeĀ­(Īŗ<sup>1</sup>,Ī·<sup>2</sup>-PCHO)Ā­(PPh<sub>3</sub>)Ā­(CO)<sub>2</sub> (<b>4</b>). In <b>4</b>, the phosphine ligands are <i>trans</i> and the aldehyde is Ļ€-bonded. The geometry around Fe is pseudo trigonal bipyramidal. To gain insights into the mechanism and scope of the Cā€“C coupling reaction, complexes were prepared with the imine Ph<sub>2</sub>PC<sub>6</sub>H<sub>4</sub>CHī—»NC<sub>6</sub>H<sub>4</sub>Cl (abbreviated as PCHNAr), derived by condensation of 4-chloroaniline and PCHO. PCHNAr reacts with Fe<sub>2</sub>(CO)<sub>9</sub> and with FeĀ­(bda)Ā­(CO)<sub>3</sub> to afford the tetra- and tricarbonyl compounds FeĀ­(PCHNAr)Ā­(CO)<sub>4</sub> (<b>5</b>) and FeĀ­(PCHNAr)Ā­(CO)<sub>3</sub> (<b>6</b>), respectively. Treatment of <b>6</b> with PCHO gave the unsymmetrical Cā€“C coupling complex FeĀ­[(Ph<sub>2</sub>PC<sub>6</sub>H<sub>4</sub>)<sub>2</sub>CHĀ­(O)Ā­CHĀ­(NAr)]Ā­(CO)<sub>2</sub> (<b>7</b>). Compound <b>7</b> was also prepared by the reaction of <b>3</b> and PCHNAr. The solid-state structure of <b>7</b>, as established by X-ray crystallography, is similar to that of <b>1</b> but with an amido group in place of one alkoxide. The deuterium-labeled phosphine aldehyde PCDO was prepared by the reaction of <i>ortho</i>-lithiated phosphine Ph<sub>2</sub>PC<sub>6</sub>H<sub>4</sub>-2-Li with DMF-<i>d</i><sub>7</sub>. Reaction of <b>6</b> with PCDO gave <b>7</b>-<i>d</i><sub>1</sub> with no scrambling of the deuterium label. Attempted oxidation of <b>1</b> with FcBF<sub>4</sub> (Fc<sup>+</sup> = ferrocenium) gave the adduct FeĀ­[(Ph<sub>2</sub>PC<sub>6</sub>H<sub>4</sub>)<sub>2</sub>C<sub>2</sub>H<sub>2</sub>O<sub>2</sub>(BF<sub>3</sub>)<sub>2</sub>]Ā­(CO)<sub>2</sub> (<b>8</b>). The structures of <b>1</b> and <b>8</b> are almost identical. Compound <b>8</b> was independently synthesized by treating <b>1</b> with BF<sub>3</sub>OEt<sub>2</sub> via the intermediacy of the 1:1 adduct, which was detected spectroscopically. Qualitative tests showed that <b>1</b> also reversibly protonates with HOSO<sub>2</sub>CF<sub>3</sub> and binds TiCl<sub>4</sub>
    corecore