Ryan M.
Kirk
and
Anthony F.
Hill
*
Research School of Chemistry, Australian National University, Canberra, ACT, Australia. E-mail: a.hill@anu.edu.au
First published on 3rd July 2024
The σ-arsolido complex [Mo(AsC4Me4)(CO)3(η5-C5H5)] is alkylated at arsenic by MeOTf to afford the pentamethylarsole complex [Mo(MeAsC4Me4)(CO)3(η5-C5H5)](OTf) while iodomethane affords a mixture of [Me2AsC4Me4]I, [MoMe(CO)3(η5-C5H5)], [MoI(CO)3(η5-C5H5)] and the arsole complexes cisoid- and transoid-[MoI(MeAsC4Me4)(CO)2(η5-C5H5)] and transoid-[Mo{C(O)Me}(MeAsC4Me4)(CO)2(η5-C5H5)], The arsole ligand in [Mo(MeAsC4Me4)(CO)3(η5-C5H5)](OTf) is readily liberated by NaI in acetone to afford free MeAsC4Me4 and [MoI(CO)3(η5-C5H5)]. In a similar manner, the reaction of [Mo(AsC4Ph4)(CO)3(η5-C5H5)] with MeI affords MeAsC4Ph4 and [MoI(CO)3(η5-C5H5)], while [Mo{AsC4(SiMe3)-2-Me2-3,4}(CO)3(η5-C5H5)] with MeOTf affords [Mo{MeAsC4(SiMe3)-2-Me2-3,4}(CO)3(η5-C5H5)](OTf). The reaction of [Mo(AsC4Me4)(CO)3(η5-C5H5)] with activated alkynes (RCCR: R = CF3, CO2Me) does not proceed via [4 + 2] cyclo-addition but rather electrophilic attack at arsenic followed by metallacyclisation with incorporation of a carbonyl ligand in the spirocyclic complexes [Mo{As(C4Me4)CRCRCO}(CO)2(η5-C5H5)].
Fig. 1 Valence orbitals of interest for the complex [Mo(AsC4H4)(CO)3(η5-C5H5)].3c |
This has however yet to be experimentally explored beyond the simple addition of coordinatively unsaturated metal centres, e.g., ‘AuC6F5’, ‘Mn(CO)2(η5-C5H4Me)’ and ‘Fe(CO)2(η5-C5H5)+’ to [Mo(AsC4Me4)(CO)3(η5-C5H5)] (1a).2 Accordingly, we now report an investigation of the reactivity of a series of σ-arsolyl complexes of molybdenum, viz. [Mo(AsC4Me4)(CO)3(η5-C5H5)] (1a), [Mo(AsC4Ph4)(CO)3(η5-C5H5)] (1b) and [Mo{AsC4H(SiMe3)-2-Me2-3,4}(CO)3(η5-C5H5)] (1c) towards electrophilic alkylating agents and potentially dienophilic electrophilic alkynes.
Scheme 1 Synthesis of σ-arsolyl complexes via transmetallation.2c |
The complex 1a is readily alkylated at the heteroatom by methyl trifluoromethanesulfonate (MeOTf, OTf = SO3CF3) in Et2O solution at 0 °C. Bright yellow, X-ray quality crystals of the salt [Mo(MeAsC4Me4)(CO)3(η5-C5H5)]OTf [2a]OTf rapidly precipitated from solution and the product was isolated in near-quantitative yield (86%, Scheme 2). Care must be taken to avoid too great an excess of MeOTf which results in decomposition to oily black residues.
Spectroscopic data reflect the formal localisation of positive charge at the molybdenum centre with a shift of νCO absorptions to higher wavenumbers (CH2Cl2: 2057(vs), 1998(sh), 1965(vs) cm−1), and a shift of the 1H cyclopentadienyl resonance from 5.16 ppm (CD2Cl2) in 2a to higher frequency (5.54 ppm) in the cationic [2a]+. The As–CH3 group resonance in the 1H NMR spectrum (CD2Cl2: δH = 1.66) is to slightly lower frequency of the ring methyl groups (δH = 1.95, 1.82) which are essentially identical in chemical shift to those of the precursor 1a, being seemingly unaffected by alkylation of the arsenic. A shift to higher frequency is observed in the 13C{1H} NMR spectrum for the carbonyl ligands transoid (225.8 ppm) and cisoid (223.2 ppm) to the pentamethylarsole ligand (cf. 235.5 and 224.7 ppm for 1a in CD2Cl2) as the efficacy of Mo(d) → CO(π*) back-bonding is reduced; the same argument applies to the cyclopentadienyl signal (93.5 vs. 94.3 ppm). The arsolyl ring-carbon nuclei are similarly displaced to slightly higher frequency by As-methylation (145.8 and 134.5 ppm vs. 149.1 and 141.8 ppm for 1a), a feature which also manifests in binuclear complexes of the μ-AsC4Me4 ligand2b as the modest ring-current within the cyclic system is further diminished by appropriation of the arsenic lone pair. High-resolution mass spectrometry (ESI) reveals extensive fragmentation of the molecular ion, the primary mode of disintegration being loss of the arsenic-bound CH3 group in addition to sequential products of decarbonylation; single crystals suitable for X-ray diffractometry were collected from the reaction mixture and the molecular structure is shown in Fig. 2.
Fig. 2 The molecular structure of [2a]+ in a crystal of [2a]OTf (50% displacement ellipsoids, hydrogen atoms and triflate anion omitted). |
Crystallographic inspection of the reaction product confirms As-methylation. Reflecting the formal positive charge on the Mo(II) atom and the change in bonding description (X to L7) between the metal and pnictogen, a contraction of the Mo–As distance by almost 0.15 Å to 2.5844(4) Å is noted (cf. 2.7267(7) Å for 1a). A similar contraction (though of lesser magnitude) is observed upon coordination of Lewis acidic metal centres to 1a.2b Apparently, sequestration of the arsenic lone pair alleviates a repulsive interaction with the electronically saturated molybdenum centre, permitting closer approach of the two centres – the so-called transition metal gauche effect described by Gladysz.8 The arsenic atom adopts a deformed tetrahedral geometry with external bonding angles of 105–117° and an internal C–As–C ring-angle of ca. 89°. The pentamethylarsole ligand is swivelled about the Mo–As bond to lie in an unsymmetrical disposition about the molecule, similar to that found in 1b,22 at the expense of the intramolecular C–H⋯π contact found in the solid state structure of 1a. Interatomic distances within the arsole ring are reflective of decreased aromatic character with a slight compression of the CαCβ bonds (mean 1.327(5) Å) and elongation of the Cβ–Cβ bond (1.496(5) Å) compared to the neutral arsolyl complex (1.351(7) and 1.469(8) Å, respectively), though these variations lie at the precision limits of the data.
Alkylation of 1c was explored similarly: treatment with MeOTf in Et2O solution causes the product [2c]OTf to rapidly precipitate from solution in excellent yield (81%) as bright yellow crystals suitable for X-ray diffraction. Spectroscopic features are comparable to those of the pentamethylarsole complex [2a]+ and accordingly, only salient features are noted. The solution IR spectrum reveals the usual trio of carbonyl vibrations (CH2Cl2: 2065(vs), 1992(sh), 1971(vs) cm−1) shifted to higher wavenumbers from 1c (2007(vs), 1938(sh), 1922(vs) cm−1), and the resonance for the olefinic hydrogen α to the arsenic atom appears at 6.97 ppm in CD2Cl2 solution. The cyclopentadienyl ligand resonances appear at 5.82 and 92.4 ppm in the 1H and 13C{1H} NMR spectra, respectively, and the carbonyl ligands at 235.7 (transoid), 225.6 and 223.6 ppm (diastereotopic cisoid) in the latter. The four chemically inequivalent arsole ring-carbon nuclei are found at 158.4, 157.1, 149.7 and 146.7 ppm with the resonance at δC = 149.7 corresponding to the methylidyne carbon as indicated by both HSQC NMR measurements, and by nOe enhancement. The molecular structure was determined by X-ray diffractometry and is shown in Fig. 3.
The molecular structure of [2c]+ in [2c]OTf is largely similar to that of [2a]+. The arsole ligand lies “sideways” against the Mo(CO)3(η5-C5H5) centre with the SiMe3 group occupying the region of space between the cisoid carbonyl ligands. The As→Mo dative bond length of 2.6016(6) Å is slightly longer than that in [2a]+ (2.5844(4) Å), though it is noted that the Mo–As covalent bond in 1c is also ca. 0.02 Å longer than in [2a]+ no doubt due to the intrinsic proximal steric influence of the SiMe3 residue. Dimensions of the arsole ring follow the same pattern as for [2a]+ with a compression of the inequivalent CαCβ′ bonds (1.326(6), 1.342(6) Å) and elongation of the Cβ–Cβ bond (1.489(6) Å) upon methylation of the arsenic atom, compared to neutral 1c.
The tetraphenylarsolyl complex 1b undergoes alkylation with reluctance. Treatment with MeOTf in toluene solution gives no immediate precipitation of a salt, and only upon standing for several days at ambient temperature do a very small quantity of yellow crystals form which were regrettably unsuitable for X-ray diffractometry; accompanying these crystals were a larger quantity of oily brown residues. Dissolution of the isolated crystals precipitated more of these insipid residues and MeAsC4Ph4 (3, Fig. 5) was the only identifiable species. The same outcome was obtained in CH2Cl2 solution which resulted in 3 being the only isolable product. The observation that tetraphenylarsoles are generally hesitant to quarternisation suggests a very poorly nucleophilic arsenic centre, in line with observations to follow involving methyl iodide as the alkylating agent. Geometrical minimisation at the ωB97X-D/6-31G*/LANL2Dζ level of density functional theory2c returns rather similar natural charges (+0.678 a.u. for 1a; +0.677 a.u. for 1b) and negligible atom-condensed Fukui functions (f+ = 0.005 for 1a; 0.002 for 1b). The latter are somewhat misleading because for 1a the arsenic lone pair lies only 0.1 eV below the HOMO which is associated with the butadiene component of the arsolyl group. For 1b, the energy gap is larger (0.4 eV) and the HOMO is delocalised out to the 2,5-phenyl substituents. This suggests that the low nucleophilicity of 1b is primarily due to the steric bulk surrounding the arsenic (Fig. 4) rather than any significant electronic impact. The tetraphenylarsolyl complex 1b is, however, cleaved by an excess of CH3I in solution overnight to provide [MoI(CO)3(η5-C5H5)] (1H NMR (C6D6): δ 4.44 ppm. IR (CH2Cl2): 2043(vs), 1961(vs) cm−1) and 1-methyl-2,3,4,5-tetraphenylarsole (3) as the only products by NMR spectroscopy. Repeating the reaction in CH2Cl2 on a preparative scale allowed isolation of the dark red molybdenum iodide complex (83%) and the light yellow MeAsC4Ph4 arsole (3, 70%, Fig. 5) in high yields.
Fig. 4 Corey–Pauling–Kolton models for (a) [Mo(AsC4Me4)(CO)3(η5-C5H5)] and (b) [Mo(AsC4Ph4)(CO)3(η5-C5H5)]. |
The reaction almost certainly proceeds via initial methylation of the arsenic atom (which may or may not follow a radical pathway, vide infra) followed by rapid and irreversible nucleophilic substitution of the arsole ligand from the [Mo(MeAsC4Ph4)(CO)3(η5-C5H5)]+ [2b]+ complex by the iodide anion. This process is presumably dissociative in nature since the Mo(II) centre is both electronically-saturated and seven-coordinate, and the displacement of the arsole ligand from the cationic complex must be at least as rapid as its formation since at no point during the reaction were any other cyclopentadienyl signals observed in the 1H NMR spectrum (Scheme 3).
The more nucleophilic 1a also reacts with excess CH3I in C6D6 solution, though the reaction was considerably more complex than anticipated (Scheme 4). Multiple products are generated, including large colourless crystals of [Me2AsC4Me4]I [4]I which separate in 62% yield from solution. This salt is the first example of an authentic cationic arsole (henceforth arsolylium), although there are examples of annulated arsoles undergoing quaternisation.9 These are excluded from the present discussion given the ring system is not based on a localised butadiene moiety, one of the major identifying features of arsoles. The salt [4]I was structurally characterised and the molecular geometry of the cation is shown in Fig. 6 in addition to the unit cell packing which indicates weak hydrogen bonding (≈3.0 Å) between the iodide and protons on four methyl groups.
The arsolylium salt [4]I was isolated from the reaction mixture in good yield and was fully characterised. The internal symmetry of the cation is reflected in the 1H NMR spectrum (CDCl3) which finds only three equally integrating signals at 2.58, 2.32 and 2.00 ppm of which the latter is assigned to the arsenic-bound methyl groups by comparison with the As–methyl resonance in [2a]OTf. In the 13C{1H} NMR spectrum the arsolyl ring-carbon nuclei manifest at 150.3 and 123.5 ppm (α and β to the heteroatom, respectively) with three methyl environments at 15.0, 14.2 and 9.2 ppm. The latter is identified as the arsenic-bound methyl carbon atoms on the basis of HSQC and HMBC NMR experiments. High-resolution mass spectrometry reveals the molecular ion in high intensity (m/z ≈ 213.0628). The formation of [4]I was confirmed in a separate experiment to arise from simple alkylation of MeAsC4Me4 by methyl iodide. Methyl triflate will also alkylate pentamethylarsole in Et2O at 0 °C, though the [4]OTf salt formed degraded over several weeks in the solid state (under argon) to oily brown residues. Conversely, the iodide salt retained its integrity for >2 years. Other methylating agents such as [Me3O]BF4 were not tested. Methyltetraphenylarsole MeAsC4Ph4 (3) did not undergo such a reaction, recalling the resistance of 1b to alkylation. A reasonable course of events involves initial alkylation of the arsenic in [Mo(AsC4Me4)(CO)3(η5-C5H5)] to form a salt [2a]I, from which dissociation of the arsole is followed by irreversible coordination of the iodide counter anion.
Hexamethylarsolylium iodide [4]I crystallises (monoclinic P21/c) with four separated cation–anion pairs contained in the unit cell. Incidentally, the iodide anions have distorted bisphenoidal coordination geometry (τ4 = 0.86, τ′4 = 0.57) in the solid-state since only four hydrogen bonds operate (per iodide) with approximate valence angles of ca. 165° (pseudo-trans axial) and 82° (pseudo-equatorial). The [4]+ cation is approximately though not crystallographically laterally-symmetric (C2v), and the quaternary arsenic atom exists in a distorted tetrahedral environment with external (local) bonding angles of 109–118°, though the internal Cα–As–Cα vertex is regular within this work at 88°. Compared to other RAsC4Me4 arsoles which have been structurally characterised,10 interatomic distances within the [Me2AsC4Me4]+ cation reflect a true 2,4-diene localisation of the π-electron density about the ring with a contraction of the As–Cα (mean 1.903(3) Å) and CαCβ bonds (mean 1.327(4) Å), and an elongation of the Cβ–Cβ′ bond (1.500(3) Å). That said, exploration of the molecular orbital manifold for the model cation [Me2AsC4H4]+ (DFT:ωB97X-D/6-31G*, Fig. 7) does suggest a modest contribution from arsenic to the π-system (HOMO−1, HOMO−4). It is also noteworthy that rather than the arsenic, as is typical for conventional arsonium cations, Cα and Cα′ would be expected to be the preferred sites for frontier orbital-controlled nucleophilic attack considering the topologies of the LUMO, as well as the atom-condensed Fukui functions (f−)11 for arsenic (f− = 0.024) and Cα/α′ (f− = 0.184). This is despite the arsenic carrying a substantial natural positive charge (+1.63 a.u.). Such processes would be expected to yield dimethyl(butadienyl)arsines (Me2AsCHCHCHCH–Nu), however this propensity would be somewhat curtailed in the real cases of arsolyliums bearing Cα-substituents, as most do.
Fig. 7 Frontier orbitals of interest for the hypothetical dimethylarsolylium cation [Me2AsC4H4]+ (DFT:ωB97X-D/6-31G*/gas phase). |
As an aside, the colourless arsolylium [4]I did appear to react with LiNiPr2 in Et2O solution at −78 °C to provide a cherry-red solution, presumed to contain an arsenic ylide H2CAsMeC4Me4 (by analogy with H2CAsPh312) which, however, decomposed to poorly-soluble light brown materials upon warming. The addition of labile [W(THF)(CO)5] or [Au(C6F5)(THT)] (THT = tetrahydrothiophene) to this red solution failed to afford isolable ylide complexes upon workup, and no trapping experiments with organic electrophiles (e.g., ketones) were attempted due to limited quantities of the arsolylium on-hand.
The remainder of the C6D6 reaction solution contained the expected [MoI(CO)3(η5-C5H5)] as the major cyclopentadienyl-containing species. Interestingly, [MoMe(CO)3(η5-C5H5)] is also observed, which speaks to the possible occurrence of radical reaction pathways, since the addition of NaI to [2a]OTf gave [MoI(CO)3(η5-C5H5)] (1H NMR (acetone-d6): δH = 5.88 ppm cf. an authentic sample) and MeAsC4Me4exclusively; the presence of a fleeting pentavalent As(V) species i.e., [Mo{AsMe(I)C4Me4}(CO)3(η5-C5H5)] which then dismutates to [MoI(CO)3(η5-C5H5)] and [MoMe(CO)3(η5-C5H5)], is therefore discounted. Pentavalent (λ5) arsoles13 are a subject to which we will return in detail in a subsequent paper. Repeating the reaction in n-hexane solution followed by storage at −30 °C provides an abundance of crystalline specimens comprising no fewer than five molybdenum-containing compounds which were identified by X-ray diffraction. In addition to [4]I (Fig. 6), they are: (a) [MoI(CO)3(η5-C5H5)] (Fig. 8), which despite being a septuagenarian compound14 had not been previously structurally characterised; (b) [MoMe(CO)3(η5-C5H5)], a similarly venerated organometallic,14 (c) cisoid-[MoI(MeAsC4Me4)(CO)2(η5-C5H5)] (cisoid-5, Fig. 8a), arising from carbonyl rather than arsole substitution, and (d) its isomer transoid-[MoI(MeAsC4Me4)(CO)2(η5-C5H5)] (transoid-5, Fig. 9b) as a co-crystallite with (e) the novel acyl complex transoid-[Mo{C(O)Me}(MeAsC4Me4)(CO)2(η5-C5H5)] (6, Fig. 10).
Fig. 8 Molecular structure of [MoI(CO)3(η5-C5H5)] in a crystal (50% displacement ellipsoids, most hydrogen atoms omitted). |
Despite being first reported in 1956 by Wilkinson,14 and having been utilised as a synthon in a variety of settings in the interregnum,15 the crystal structure of red [MoI(CO)3(η5-C5H5)] would appear to have been neglected. It transpires however that the structure had in fact been previously determined, but incorrectly identified as the diamagnetic (sic., d3) compound [Mo(Te)(CO)3(η5-C5H5)] (CCDC 216750†). This was purported to arise from the reaction of [Mo2(CO)6(η5-C5H5)2] with diphenylditelluride and [nBu4N]I,16a but is clearly [MoI(CO)3(η5-C5H5)]. Understandably, molecular features are rather unexciting given the simplistic nature of the complex and only salient crystallographic aspects are noted here. A seven-coordinate molybdenum(II) centre is found which approximates a distorted square-based pyramid, where the three carbonyls and the iodide ligand comprise the basal plane and the centroid of the cyclopentadienyl ring occupies the apical site. The Mo–I bond length of 2.8419(6) Å is typical among similar complexes,16 and the Mo–C bond transoid to the iodide (1.992(6) Å) is marginally shorter than the pair cisoid (mean 2.035(7) Å) due to the π-donor character of the halide. The unit cell (monoclinic P21/n) contains four symmetry-related molecules which display two intermolecular H⋯I interactions (per molecule) of ca. 3.1 Å in length that likely assist with relative orientations during lattice formation. The yellow organometallic complex [MoMe(CO)3(η5-C5H5)] first prepared by Wilkinson14a was structurally characterised much later by Valente and co-workers14b who sought to utilise it for the catalytic epoxidation of cyclic alkenes in the presence of tert-butylperoxide. The Mo–C(sp3) distance was 2.326(3) Å and no unusual intermolecular interactions were found.
Dark red cisoid-[MoI(MeAsC4Me4)(CO)2(η5-C5H5)] (cisoid-5, Fig. 9a) crystallises in triclinic P and the unit cell contains four molecules comprising two pairs of enantiomers, which differ in the relative position of the arsole ligand to the iodide (either to the “left” or “right”) when viewed along the molybdenum-cyclopentadienyl centroid axis. The pair of molecules within the asymmetric unit are crystallographically unique though interatomic distances and angles are statistically indistinguishable between the two entities. The molybdenum atom adopts the usual seven-coordinate geometry which approximates a square-based pyramid. The As→Mo dative bond measures 2.5640(6) (molecule A) and 2.5717(6) (B) Å and the Mo–I covalent bonds are unremarkable at 2.8524(4) (A) and 2.8442(4) (B) Å in length. The pentamethylarsole ligand lies approximately orthogonal to the vertical plane bisecting the Mo(CO)3(η5-C5H5) moiety. Distances between arsole ring atoms are typical of σ-dative coordination complexes for this ligand (which are measurably different from [4]I above), with mean As–Cα, CαCβ and Cβ–Cβ′ bond lengths of ca. 1.93, 1.35 and 1.50 Å, respectively.
The third set of red crystals contained two compounds in an unequal occupancy ratio which are apparently (or approximately) isosteric and superimposed within the lattice. They are transoid-[MoI(MeAsC4Me4)(CO)2(η5-C5H5)] (transoid-5, minor component, 11% occupancy) and transoid-[Mo{C(O)Me}(MeAsC4Me4)(CO)2(η5-C5H5)] (transoid-6, major component, 89% occupancy). Refinement (including occupancy) led to a satisfactory structural model solution (R1 = 0.0398), albeit with some eccentric displacement ellipsoids though no restraints/constraints to correct these were applied. The iodide complex is again presumed to arise from carbonyl rather than arsole substitution, and the simplest explanation for the occurrence of the acyl complex would superficially seem to involve migratory insertion of [MoMe(CO)3(η5-C5H5)] (vide infra) trapped by free pentamethylarsole.
The two compounds co-crystallise in monoclinic P21/n and in both species the pentamethylarsole ligand approximately straddles the (non-crystallographic) vertical symmetry plane of the Mo(CO)2(η5-C5H5) centre with the As–methyl substituent projected between the pair of cisoid carbonyl ligands. Since the two complexes are superimposed within the crystal lattice, the positions of the Mo(MeAsC4Me4)(CO)2(η5-C5H5) atoms are modelled as a weighted average of both and all interatomic dimensions therein are thus identical viz. the mean As→Mo dative bond for both complexes in this particular crystal is 2.5226(6) Å. All crystallographic uniqueness is observed in the region of the overlaid acyl and iodide substituents, with Mo–C(sp2), CO and Mo–I distances of 2.230(9), 1.215(9) and 2.898(4) Å, respectively. The former two are typical17 whereas the latter is elongated compared the other Mo–I bond lengths observed previously (ca. 2.4–2.5 Å) since the presence of the arsenic atom in a transoid position is presumably a repulsive influence. It is noted that a transoid configuration appears to be the preferred condition for Group 6 acyl complexes of the form [M{C(O)R}L(CO)2(η5-C5H5)]18 with which [Mo{C(O)Me}(MeAsC4Me4)(CO)2(η5-C5H5)] also complies. For [MoIL(CO)2(η5-C5H5)] it is reasonable to expect that, in general, ligands of superior π-acidity would favour the coordination site transoid to the iodide though pentamethylarsole is not expected to be particularly outstanding in this respect, possibly explaining why both cisoid and transoid isomers have crystallised. No general trend emerges from a survey of the literature and any regio preferences appear to be highly contextual e.g., steric demands.19 In many cases the cisoid and transoid isomers interconvert depending on inter alia temperature and solvent polarity15f,20 and in others it is not possible to unambiguously distinguish isomers from the spectroscopic data reported.
Acyl complexes of the mid-transition metals are well studied: their deliberate method of synthesis entails treatment of a metal carbonyl anion with an acyl halide or by migratory insertion of an alkyl complex,21 often in the presence of a Lewis base18a,b,20,22 although some Lewis acids have also been found to promote this process.23 For acyl complexes prepared from the corresponding alkyl–carbonyl in the presence of a neutral two-electron ligand, the extraneous ligand is usually of reasonable basicity (e.g., isocyanides, phosphines or phosphites) and weaker nucleophiles (amines, nitriles, arsines, stibines etc.) either provide the product in low yield24 or afford products of decarbonylation, especially for Group 6 centres. It is therefore intriguing that an acyl complex possessing an arsole ligand was observed, since (a) arsines appear less inclined to affect migratory insertion reactions upon exposure to [MoR(CO)3(η5-C5H5)],25 (b) the reaction was carried out at −30 °C whereas thermal impetus (generally ≫25 °C) is often required, and (c) the reaction was carried out in n-hexane medium, which is considered to disfavour the migration process, being unable to effectively stabilise the coordinatively unsaturated and dipolar transition state.22a Attempts were made to deliberately synthesise [Mo{C(O)Me}(MeAsC4Me4)(CO)2(η5-C5H5)] (6) by heating [MoMe(CO)3(η5-C5H5)] to 80 °C in CD3CN with one equivalent of MeAsC4Me4. The metal alkyl signals were unaffected and only gradual decomposition of the arsole was observed; a similar outcome was obtained with PhAsC4Me4, an arsole of superior thermal stability, wherein the spectrum was totally unchanged after heating. It is also noted that these arsoles did not proceed to carbonyl substitution products either (i.e., cisoid- or transoid-[MoMe(RAsC4Me4)(CO)2(η5-C5H5)], R = Me, Ph), speaking to their poor nucleophilicity. The presence of 6 among the crystalline products therefore remains inexplicable since the reaction conditions would be expected to disfavour formation of such a compound (vide supra); like the occurrence of [MoMe(CO)3(η5-C5H5)] in the mixture, it is possible that radical processes may be operative though not via any photochemical route since this reaction was performed in the dark.
The nucleophilic arsolyl 1a reacts rapidly with a slight excess of DMAD in CH2Cl2 solution at ambient temperature and complete consumption of 1a occurs within only a few minutes of mixing as confirmed by IR spectroscopy. Purification via column chromatography on Florisil® provided an air-stable, reddish-brown solid as the only isolable compound in good yield (85%). High-resolution mass spectrometry confirmed the formation of a 1:1 1a/DMAD cycloadduct (m/z ≈ 572.9797) though both the 1H NMR (C6D6) and IR data were not consistent with a Cs-symmetric [4 + 2] cycloaddition 9-arsanorbornadien-9-yl product of the form [Mo{AsC6Me4(CO2Me)2}(CO)3(η5-C5H5)] (Scheme 5).
Neither the arsolyl nor alkyne moieties eventuate in symmetrical environments: a pair of methyl ester (3.45, 3.23 ppm, relative integration 3:3) and trio of methyl (1.88, 1.59 and 1.54 ppm, relative integration 3:3:6) signals in the 1H NMR spectrum, in addition to three equal intensity carbonyl resonances (δC = 263.2, 245.5 and 234.7) indicate the molecule has no element of symmetry. The 9-arsanobornadienyl formulation would allow a time-averaged molecular plane of symmetry on the 13C NMR timescale through rotation around the Mo–As bond and may therefore be discarded. Furthermore, the solution IR spectrum (CH2Cl2) exhibits a νCO intensity profile that is not suggestive of a Cs-symmetric Mo(CO)3(η5-C5H5) centre. Instead, only two terminal carbonyl modes are observed at 1962(vs) and 1889(s) cm−1 in addition to a lower-energy absorption at 1732(s) cm−1 indicative of a ketonic carbonyl. A pair of medium-intensity methyl ester carbonyl absorptions are found at 1605 and 1580 cm−1.
The compound crystallised from Et2O at −20 °C and X-ray diffraction revealed its identity as a novel spirocyclic 3-(arsolyl)propenoyl chelate complex [Mo{C(O)CRCRAsC4Me4}(CO)2(η5-C5H5)] (R = CO2Me, 7a). The Z-propenoyl linkage arises from insertion of a DMAD molecule into the region of space between the nucleophilic arsenic atom and an electrophilic cisoid carbonyl ligand, forming a new five-membered molybda-arsacyclopentenone (Scheme 5).
The molecular structure is shown in Fig. 11, and a proposed mechanism of formation is given in Scheme 4. Consistent with the nucleophilicity of the arsenic centre discussed above, the mechanism commences with DMAD acting as an electrophile towards arsenic, rather than as a dienophile to the arsolyl butadiene component. It should be noted that a not dissimilar process has been suggested to occur in the reaction of triphenylarsine with DMAD, for which a somewhat questionable and yet to be substantiated, λ5-arsenole Ph3AsC4(CO2Me)4 has been proposed to arise.29,30
The cycloadduct crystallises in triclinic P and a pair of symmetry-related enantiomeric molecules (Fig. 11) are found in the unit cell. The net result of the reaction with DMAD is the formation of a new Mo–C covalent bond and conversion of the Mo–As bond to a As→Mo dative interaction. Thus, the molecular structure of the spirocyclic product features a pair of vertex-sharing pentagons united at the deformed tetrahedral arsenic atom – their angle of intersection (defined by the separate mean planes of each five-membered ring) is nearly perpendicular at 87°. This indicative geometric parameter comes with some imprecision as only the arsole ring is approximately planar (envelope angle ca. 7°) whereas the molybda-arsacyclopentenone is puckered at the arsenic and acyl carbon atoms by ca. 20°. Reflecting the fact that all interatomic bonds within the ring are unique, the dimensions of the pentagonal metallacyclic system are expectedly distorted from the ideal such that the molybdenum atom is skewed toward the acyl carbon by over 0.25 Å (Mo–C 2.244(2) vs. As→Mo 2.5110(2) Å). The latter compares to the marginally longer As→Mo dative interaction of 2.5844(4) Å found in [2a]+ (Fig. 2). Ring-atom distances within the arsolylpropenoyl chelate are typical of the individual bonding partners, comprising regular localised CO, C–C, CC and C–As covalent bonds. Reflecting the appreciable adjustment to bonding patterns surrounding the Mo(II) centre, the 13C{1H} NMR spectrum (C6D6) of 7a reveals two resonances for the diastereotopic terminal carbonyl ligands, to low frequency of 1a (263.2, 245.5 ppm) with the acyl carbonyl resonance at 234.7 ppm. The chemically unique methyl ester carbon signals appear at 167.1 and 159.9 ppm to slightly higher frequency of the propenoyl olefin carbon atoms at 161.0 (acyl-bound) and 149.7 (arsenic-bound) ppm. The arsole ring atoms are all inequivalent, manifest as four distinct signals in the range 146.5–132.9 ppm while the remaining cyclopentadienyl and methyl resonances are unremarkable.
Given the formation of 7a from 1a and DMAD, the generality of the reaction with respect to alkyne functionalisation was explored. Combining 1a with hexafluorobut-2-yne (HFB, F3CCCCF3) in CH2Cl2 solution at −78 °C followed by warming to ambient temperature was accompanied by a colour change from orange to a deep red, and purification by column chromatography on Florosil® provided the second example of a 3-(arsolyl)propenoyl species, viz. [Mo{C(O)CRCRAsC4Me4}(CO)2(η5-C5H5)] (R = CF37b), in high yield (90%). Alkyne insertion is confirmed by comparison of IR, NMR and MS data with those for 7a in addition to the 19F{1H} NMR spectrum (C6D6) which confirmed the two chemically unique CF3 groups (δF = –56.7, −59.1 ppm, 5JFF = 10.2 Hz). Single crystals of 7b were grown by evaporation of a Et2O solution at 0 °C and the molecular structure, determined by X-ray diffraction, is shown in Fig. 12.
The complex crystallises in centrosymmetric monoclinic P21/n and the centric unit cell therefore contains both enantiomers, though only one is shown in Fig. 12. The gross molecular configuration of 7b is analogous to that of 7a (Fig. 10) and most salient features of the latter are applicable to the former. Thus, the formal [3 + 2] addition of one HFB molecule to the region between the arsenic atom and a cisoid carbonyl provides a new five-membered ring encompassing the arsolylpropenoyl chelate ligand. The respective As→Mo and Mo–C bonds measure 2.5161(5) and 2.216(3) Å, respectively, and the mean planes of the vertex sharing five-membered rings intersect at an angle more acute than in the previously discussed DMAD [2 + 3] cycloadduct, approximately 78°, though a similar puckering of the molybda-arsacyclopentenone ring (ca. 17°) is measured.
The bright yellow/orange colours of 7a and 7b may be traced to primarily MLCT transitions from metal-based orbitals (HOMO−0,1,2) to unoccupied orbitals located primarily on the metallacycle (LUMO), the arsole (LUMO+1) and carbonyl ligands (LUMO+2) on the basis of computational analysis (TD-DFT:ωB97X-D/6-31G*/LANL2Dζ/gas-phase) of 7b (see ESI†).
In contrast to the facile replacement of pentamethylarsole by iodide from [2a]+, the arsolylpropenoyl complexes were surprisingly resistant toward displacement of the arsenic donor: no reaction whatsoever was observed upon treatment of 7a with an excess of PMe3, P(OMe)3 or tBuNC at elevated temperatures (C6D6, 80 °C). The NMR spectrum of 7b was unchanged after being sealed under an atmosphere of carbon monoxide for one week, and the addition of [NO]PF6 to a CD3CN solution of the same led to complete decomposition with gas evolution. No reaction was observed between 1a and electron rich alkynes such as 2-butyne and diphenylacetylene, nor between [2a]+ or [MoAu(μ-AsC4Me4)(C6F5)(CO)2(η5-C5H5)] (from 1a and [Au(C6F5)(THT)]2b) and DMAD. The addition of other typically dienophilic substrates such as ortho-benzyne (via the Kobayashi protocol31), tetracyanoethylene and N-phenylmaleimide did not afford similar spirocyclic cycloadducts (by mass spectrometry) and instead yielded chromatographically immobile compounds, except for the former which afforded a very small quantity of [Mo2(CO)6(η5-C5H5)2]. The course of these latter reactions was not reliably deduced from spectroscopic data.
In light of the above reactions between 1a and DMAD or HFB, it is unsurprising that other transition metal arsenidos undergo analogous ring forming process with electron-poor alkynes: Davidson et al., have treated the cacodyl complexes [M(AsMe2)(CO)n(η5-C5H5)] (M = Fe, n = 2; M = W, n = 3) with the same alkynes to produce the metallacycles [M{C(O)CRCRAsMe2}(CO)n−1(η5-C5H5)]32 and in this regard the behaviour of the AsC4Me4 ligand resembles that of a conventional arsenido ligand despite the nucleophilic diene function in combination with the demonstrated dienophilic proclivities of DMAD and HFB. Notably, a large excess of the alkyne was employed yet the arsole butadiene remained intact, even when the arsenic lone pair of electrons is already sequestered in As:→M bonding to reinforce the diene (cf. aromatic elementole) character. Mathey has noted that the disubstituted phospholyl complexes [W(PC4R2R′2)(CO)3(η5-C5H5)] (R = H, R′ = Me; R = Ph, R′ = H) are also reluctant to undergo [4 + 2] cycloaddition reactions with dienophiles (despite the same phosphole rings entering into cycloaddition reactions when coordinated to W(CO)533), instead yielding phospholylpropenoyl complexes.34 Cullen has reported an arsinopropenoyl complex of rhenium(I) obtained by photolysis of [Re2(CO)10] with the bis(arsino)butene Z-Me2AsC(CF3)C(CF3)AsMe2, though the low yields discourage mechanistic conjecture.35
Crystallographic data have been deposited at the CCDC under accession numbers 2145381–2145383, 2149526, 2145364, 2145367, 2145351 and 2145459 and can be obtained from https://www.ccdc.cam.ac.uk.
Footnote |
† Electronic supplementary information (ESI) available: Spectroscopic, computational and crystallographic data. CCDC 2145351, 2145364, 2145367, 2145381–2145383, 2145459 and 2149526. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4dt01371a |
This journal is © The Royal Society of Chemistry 2024 |