Fig 2 - uploaded by John C Huffman
Content may be subject to copyright.
31 P M 1 H N NMR (162 MHz) spectrum of RuHCl[C(Me)(OCH CHOCH )](PPr i ) , showing the inequivalence of the phosphorus nuclei 2 caused 2 by the 3 2 epoxide carbon asymmetry. Starred peaks are impurities and the arrows indicate the outer lines of the AB pattern. 

31 P M 1 H N NMR (162 MHz) spectrum of RuHCl[C(Me)(OCH CHOCH )](PPr i ) , showing the inequivalence of the phosphorus nuclei 2 caused 2 by the 3 2 epoxide carbon asymmetry. Starred peaks are impurities and the arrows indicate the outer lines of the AB pattern. 

Source publication
Article
Full-text available
Dehydrohalogenation of RuH2Cl2L2 (L = PPr3i) gives (RuHClL2)(2), shown to be a halide-bridged dimer by X-ray crystallography; the fluoride analog is also a dimer. (RuHClL2)(2) reacts with N-2, pyridine and C2H4 (L') to give RuHCIL'L-2, but with vinyl ether and vinyl amides, H2C=CH(E) [E = OR, NRC(O)R'] such olefin binding is followed by isomerizati...

Context in source publication

Context 1
... this is the ground state structure of this molecule, reaction of RuHClL with excess C D in C H for 1 h at 20 ¡C shows ( 2 H NMR) 2 deuteration of 2 the 4 H on 6 6 Ru and also the methyl groups of coordinated PPr i . This is indicative of 3 reversible insertion of C D into the Ru È H bond and it also 2 4 indicates that the 16-electron oleÐn hydride form ( B ) is more stable than Ru(C H )ClL . Deuteration of the Pr i methyls is 2 5 2 accounted for by reversion to RuDCl[P(CH(CH ) ) ] and 3 2 3 2 then CH /RuD scrambling, as was independently established 3 for this species deuterated by two independent methods. Inser- tion of ethylene into Ru È H is also evidenced by the formation of ethane, detected by 1 H NMR. Its formation most likely occurs by alkane elimination from unstable Ru(C H )ClL , which has oxidatively added an Pr i methyl group (C 2 È H). 5 This 2 C È H activation also o†ers an additional explanation for deuteration of the Pr i methyl with C D . The metal-containing 2 4 products after CH CH elimination could not be identiÐed. 3 3 There is no isomerization of ethylene into the carbene ligand CH(CH ). 3 RuHClL also reacts within 30 min with the oleÐns 1- 2 hexene and styrene, but only 10% adduct is formed, with unreacted RuHClL comprising the bulk of the resulting 2 mixture at 25 ¡C. Both these adducts are identiÐed by a new hydride signal in 1 H NMR ( [ 23.7 ppm, m, for 1-hexene and [ 22.1 ppm, apparent triplet, for styrene) and corresponding 31 P M 1 H N NMR AB patterns centered at 37.2 ppm ( 2 J \ 287 Hz) for 1-hexene and at 85.9 ppm ( 2 J \ 34 Hz) for P v styrene. P P v P The spectroscopic similarity of the ethylene, 1-hexene, and ethyl vinyl ether adducts and large di†erence of the styrene adduct suggests that styrene may coordinate di†erently (perhaps as g 2 : g 1 h 2 -vinyl : arene or g 2 h 4 -arene). Longer reaction times yield no carbenes, but only complex mixtures of products. RuHClL rapidly e†ects a formal 1,2-hydrogen migration 2 [eqn. (7)] of vinyl ethers into the coordinated carbene. The reaction occurs for a variety of groups R, including those with SiMe , ether, alcohol, tertiary amino, Ñuoro, and epoxide functionality. 3 All products show diastereotopic Pr i methyl groups, consistent with the presence of three di†erent substituents, H, Cl and carbene, on Ru. While this reveals nothing about the square pyramidal vs . trigonal bipyramidal geometry around Ru, the hydride chemical shifts (Table 3), rather far upÐeld ( [ 21 ppm), could be interpreted in terms of the hydride being approximately trans to an empty site ( i . e ., square pyramidal). The hydride chemical shift is thus also perhaps the most generally sensitive indicator of whether there is a ligand (from functionality in the substituent R) coordinated trans to hydride. The similarity of the hydride chemi- cal shift for R Et (Table 3) to those for all components with functionalized alkyl groups in eqn. (7) suggests that none of the latter O, N or F donors coordinates to Ru. 4 Perhaps this is caused by the difficulty of forming six-membered rings. For the case of F, this conclusion is reinforced by a 19 F chemical shift that lies within 1 ppm of that of the free oleÐn, and by the absence of coupling to F in the 31 P NMR spectrum. Even at [ 80 ¡C, 19 F NMR spectra reveal that F remains uncoordi- nated. When O is from an epoxide ring, the presence of a chiral carbon b to the vinyloxy oxygen necessitates phosphine inequivalence. From the magnitude of their 31 P M 1 H N NMR (Fig. 2) chemical shift di†erence [ * d \ only 0.09 ppm (15 Hz), 2 J \ 220 Hz], it is safe to assume that the chiral center is P v P far from the phosphorus atoms, indicating no epoxide binding to ruthenium. The case of dihydrofuran [eqn. (8)] is interesting as it shows that a cyclic internal vinyl ether is also easily isomerized. 5 The hydride chemical shift is consistent with no donation to Ru ...

Citations

... State of the Art Computations on Ruthenium Carbenes: The aim of a computational study has to be the exploration of the whole mechanism. Especially in the past 12 monthspresumably encouraged by the now available experimen¬ tal dataseveral computational studies on ruthenium carbene complexes have appeared [19,59,71,73,78,79,98,144,145,[159][160][161][162][163][164][165][166][167]. While most of the studies consider only a few species of the catalytic cycle focusing either on the ruthenium carbene formation process [159][160][161][162][163][164][165] of the catalytic cycle [19,71,73,98] there are only few studies which treat the complete mechanism or/and alternative reaction pathways [78,79,144,145,166,167]. ...
... Especially in the past 12 monthspresumably encouraged by the now available experimen¬ tal dataseveral computational studies on ruthenium carbene complexes have appeared [19,59,71,73,78,79,98,144,145,[159][160][161][162][163][164][165][166][167]. While most of the studies consider only a few species of the catalytic cycle focusing either on the ruthenium carbene formation process [159][160][161][162][163][164][165] of the catalytic cycle [19,71,73,98] there are only few studies which treat the complete mechanism or/and alternative reaction pathways [78,79,144,145,166,167]. An overview of possible reaction pathways is given in Scheme 3.1. ...
... The same activation mechanism as suggested for 59 may also elucidate the transformation of Ru(II) precursors like RuCl2(PPh3)4, RuCl2(PPh3)3, RuHCl(PPh3)3, RuCl2(py)2(PPli3)2 in the presence of norbornene into active olefin metathesis cat-alysts where so far the participation of traces of di oxygen [312] or C-H activation and a-hydrogen migration [159][160][161][162][163][164][165] ...
Article
Ruthenium(II) photocatalysis has emerged as one of the most advanced tools amongst modern synthetic chemistry whereas its catalytic mode is generally limited to single electron transfer and triplet energy transfer...
Article
A series of mononuclear σ-phenyl ruthenium complexes Ru(CO)Cl(C6H4-R)(PⁱPr3)2 (R = OMe, CH3, H, F, CF3) were synthesized and analyzed with respect to their electrochemical and spectroscopic properties. To these ends, cyclic voltammetry, IR, and UV/Vis/NIR spectroelectrochemistry as well as EPR spectroscopy on their one-electron oxidized radical cations were employed. Experimental work is complimented by quantum chemical calculations. Our studies reveal that the σ-phenyl ligand strongly contributes to the HOMO and actively participates in the redox processes. Despite comparatively smaller ligand contributions, the redox potentials, the position of the CO stretch as well as the oxidation induced CO band shifts are more sensitive toward the σ-Hammett parameter of the 4-substituent than for related styryl complexes with the same Ru(CO)Cl(PⁱPr3)2 metal coligand platform. The comparatively high spin density/positive charge at the 4-position of the phenyl ligand leads to oxidatively induced dehydrodimerization of the radical cation of the parent phenyl complex Ru(CO)Cl(C6H5)(PⁱPr3)2 (1) to the biphenylene-bridged dinuclear complex [{Ru(CO)Cl(PⁱPr3)2}2(μ-C6H4–C6H4-4,4′)]ⁿ⁺ (6ⁿ⁺). The latter was identified in spectroelectrochemical experiments and authenticated by independent preparation of neutral 6 and monitoring of its spectroelectrochemical behavior.
Article
Dimeric [RuHClL2](2) (L = PPr3i), a source of the 14-electron fragment RuHClL2, reacts at reflux in THF to cleanly form equimolar L2HClRu=C(CH2)(3)O and RuHCl(H-2)L-2, The products of stoichiometric geminal dehydrogenation of the alpha -C of THF. The same products are produced slowly at 25 degreesC by reaction of the transient RuHClL2. Double dehydrogenation of the sp(3) alpha -C of THF is also effected at 25 degreesC by if two Os(H)(3)ClL2 if two H ligands are removed with Bu-t(H)C=CH2.
Article
Mar Gómez-Gallego studied chemistry at the Universidad Complutense de Madrid (UCM) where she obtained her Ph.D. in 1987. She continued her scientific education with a Fleming Postdoctoral Fellowship with Professor W. M. Horspool and she returned to Madrid where she was appointed Professor Ayudante in 1990 and then Professor Titular in 1992. Her current research interests are focused on organometallic chemistry as well as the development of new iron-chelating agents and the study of their environmental impact.
Article
Aspects of the chemistry of zero-valent ruthenium complexes having a cyclic diene ligand, and their use for homo- and cross-dimerizations are reviewed. The naphthalene ligand in [Ru(η6-naphthalene)(η4-1,5-COD)] is readily removed to generate a formal 12e species, “Ru(η4-1,5-COD)”, having 6e coordination sites. These 6e coordination sites can be divided into 4e and 2e sites, and a conjugated compound and substituted alkene therefore selectively coordinate to them to satisfy the 18e rule. The zero-valent ruthenium center is highly Lewis basic and the oxidative coupling reaction readily occurs to form a ruthenacycle. The ruthenium(0) complexes catalyze homo- and cross-dimerizations by the oxidative coupling mechanism and some of the solid evidence in support of the mechanism is summarized. The coordination of substrates is prochiral-selective and when a chiral cyclic diene is incorporated in the catalyst, the first enantioselective cross-dimerizations are achieved.
Article
The anionic dihydride complex [Cp2TaH2](-) was synthesized as a well-defined molecular species by deprotonation of Cp2TaH3 while different solubilizing agents, such as [2.2.2]cryptand and 18-crown-6, were applied to encapsulate the alkali-metal counterion. The ion pairs were characterized by multiple spectroscopic methods as well as X-ray crystallography, revealing varying degrees of interaction between the hydride ligands of the anion and the respective countercation in solution and in the solid state. The [Cp2TaH2](-) complex anion shows slow exchange of the hydride ligands when kept under a D2 atmosphere, but a very fast reaction is observed when [Cp2TaH2](-) is reacted with CO2, from which Cp2TaH(CO) is obtained as the tantalum-containing reaction product, along with inorganic salts. Furthermore, [Cp2TaH2](-) can act as a synthon in heterobimetallic coupling reactions with transition-metal halide complexes. Thus, the heterobimetallic complexes Cp2Ta(μ-H)2Rh(dippp) and Cp2Ta(μ-H)2Ru(H)(CO)(P(i)Pr3)2 were synthesized and characterized by various spectroscopies and via single-crystal X-ray diffraction. The new hydride bridged tantalum-rhodium heterobimetallic complex is cleaved under a CO atmosphere to yield mononuclear species and slowly exchanges protons and hydride ligands when exposed to D2 gas.
Article
Synthesis, spectroscopic, and X-ray structural characterization of Ru2HnCl4-nL4 (n = 2, 3) and Ru2H2F2L4 (L = PiPr3) are reported. The structure of Ru2HCl3L4 is also reported. These are dinuclear species containing two five-coordinate, approximately square-pyramidal metal atoms. Halides, not hydrides, preferentially occupy bridging sites, and the RuXL2 terminal moiety shows limited fluxionality, but hydrides do not migrate between metals. The limited steric protection provided by PiPr3 is evident from the dimerization observed and from the fact that all these structures have rather small ∠P−Ru−P (∼105°). Also reported are RuHXL2 species with X = acetylacetonate, phenoxide, O3SCH3, and O3SCF3. Several examples of coordinated olefin to complexed carbene conversions are used to test the influence of anion X on reactivity.
Article
Reactions of the phenyliridium(III) complex [Cp*IrCl(Ph)(PMe3)] with internal alkynes (RCCR, R = Ph or Me) in the presence of NaBAr4F give rise to the formation of the (o-yinylaryl) iridium complexes [Cp*Ir{o-C6H4C(R)=CHR}(PMe3)][BAr4F] (2a, R = Ph; 2b, R = Me) via the alkyne insertion and the subsequent vinyl-to-aryl 1,4-miration of the Ir(III) center, while a similar reaction of PhC CMe affords the pi-allyl complex [Cp*Ir{eta(3)-CH(Ph)C(Ph)CH2}-(PMe3)][BAr4F] (4) in high yield. The latter is the formal 1,3-Ir migration product formed from the vinyliridium complex [Cp*Ir{C(Ph)=C(Me)Ph}(PMe3)][BAr4F] (3c). Deuterium labeling experiments have revealed that this pi-allyl complex is produced by two distinct mechanisms, (1) C=C bond rotation in 3c, followed by direct 1,3-Ir migration, and (2) successive vinyl-to-aryl and aryl-to-allyl 1,4-Ir migrations by way of the o-vinylaryl complex [Cp*Ir{o-C6H4C(Me)=CHPh}(PMe3)][BAr4F] (2c). This reaction provides the first experimental evidence of the direct 1,3-metal migration accompanied by the C-H bond activation. In addition, detailed analysis of the reaction system has demonstrated that a cyclometalated carbene complex 5, which is formed from 2c or 3c through the intramolecular C-H oxidative addition and the subsequent 1,3-hydrido migration, is involved in the isomerization process.