首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 359 毫秒
1.
Carbonylation of the anionic iridium(III) methyl complex, [MeIr(CO)2I3] (1) is an important step in the new iridium-based process for acetic acid manufacture. A model study of the migratory insertion reactions of 1 with P-donor ligands is reported. Complex 1 reacts with phosphites to give neutral acetyl complexes, [Ir(COMe)(CO)I2L2] (L = P(OPh)3 (2), P(OMe)3 (3)). Complex 2 has been isolated and fully characterised from the reaction of Ph4As[MeIr(CO)2I3] with AgBF4 and P(OPh)3; comparison of spectroscopic properties suggests an analogous formulation for 3. IR and 31P NMR spectroscopy indicate initial formation of unstable isomers of 2 which isomerise to the thermodynamic product with trans phosphite ligands. Kinetic measurements for the reactions of 1 with phosphites in CH2Cl2 show first order dependence on [1], only when the reactions are carried out in the presence of excess iodide. The rates exhibit a saturation dependence on [L] and are inhibited by iodide. The reactions are accelerated by addition of alcohols (e.g. 18× enhancement for L = P (OMe)3 in 1:3 MeOH-CH2Cl2). A reaction mechanism is proposed which involves substitution of an iodide ligand by phosphite, prior to migratory CO insertion. The observed rate constants fit well to a rate law derived from this mechanism. Analysis of the kinetic data shows that k1, the rate constant for iodide dissociation, is independent of L, but is increased by a factor of 18 on adding 25% MeOH to CH2Cl2. Activation parameters for the k1 step are ΔH = 71 (±3) kJ mol, ΔS = −81 (±9) J mol−1 K−1 in CH2Cl2 and ΔH = 60(±4) kJ mol−1, ΔS = −93(± 12) J mol−1 K−1 in 1:3 MeOH-CH2Cl2. Solvent assistance of the iodide dissociation step gives the observed rate enhancement in protic solvents. The mechanism is similar to that proposed for the carbonylation of 1.  相似文献   

2.
Metathesis of [(η33−C10H16)Ru(Cl) (μ−Cl)]2 (1) with [R3P) (Cl)M(μ-Cl)]2 (M = Pd, Pt), [Me2NCH2C6H4Pd(μ-Cl)]2 and [(OC)2Rh(μ-Cl)]2 affords the heterobimetallic chloro bridged complexes (η33-C10H16) (Cl)Ru(μ-Cl)2M(PR3)(Cl) (M = Pd, Pt), (η33-C10H16) (Cl)Ru(μ-Cl)2PdC6H4CH2NMe2 and (η33-C10H16) (Cl)Ru(μ-Cl)2Rh(CO)2, respectively. Complex 1 reacts with [Cp*M(Cl) (μ-Cl)]2 (M = Rh, Ir), [p-cymene Ru(Cl) (μ-Cl]2 and [(Cy3P)Cu(μ-Cl)]2 to give an equilibrium of the heterobimetallic complexes and of educts. The structures of (η33-C10H16)Ru(μ-Cl)2Pd(PR3) (Cl) (R = Et, Bu) and of one diastereoisomer of (η33-C10H16)Ru(μ-Cl)2IrCp*(Cl) were determined by X-ray diffraction.  相似文献   

3.
The phosphinoalkenes Ph2P(CH2)nCH=CH2 (n= 1, 2, 3) and phosphinoalkynes Ph2P(CH2)n C≡CR (R = H, N = 2, 3; R = CH3, N = 1) have been prepared and reacted with the dirhodium complex (η−C5H5)2Rh2(μ−CO) (μ−η2−CF3C2CF3). Six new complexes of the type (ν−C5H5)2(Rh2(CO) (μ−η11−CF3C2CF3)L, where L is a P-coordinated phosphinoalkene, or phosphinoalkyne have been isolated and fully characterized; the carbonyl and phosphine ligands are predominantly trans on the Rh---Rh bond, but there is spectroscopic evidence that a small amount of the cis-isomer is formed also. Treatment of the dirhodium-phosphinoalkene complexes with (η−CH3C5H4)Mn(CO)2thf resulted in coordination of the manganese to the alkene function. The Rh2---Mn complex [(η−C5H5)2Rh2(CO) (μ−η11−CF3C2CF3) {Ph2P(CH2)3CH=CH2} (η−CH3C5H4)Mn(CO)2] was fully characterized. Simi treatment of the dirhodium-phosphinoalkyne complexes with Co2(CO)8 resulted in the coordination of Co2(CO)6 to the alkyne function. The Rh2---Co2 complex [(η−C5H5)2Rh2(CO) (μ−η11−CF3C2CF3) {Ph2PCH2C≡CCH3}Co2(CO)2], C37H25Co2F6O7PRh2, was fully characteriz spectroscopically, and the molecular structure of this complex was determined by a single crystal X-ray diffraction study. It is triclinic, space group (Ci1, No. 2) with a = 18.454(6), B = 11.418(3), C = 10.124(3) Å, = 112.16(2), β = 102.34(3), γ = 91.62(3)°, Z = 2. Conventional R on |F| was 0.052 fo observed (I > 3σ(I)) reflections. The Rh2 and Co2 parts of the molecule are distinct, the carbonyl and phosphine are mutually trans on the Rh---Rh bond, and the orientations of the alkynes are parallel for Rh2 and perpendicular for Co2. Attempts to induce Rh2Co2 cluster formation were unsuccessful.  相似文献   

4.
1H NMR line broadening is found to be an effective complimentary method to chemical trapping for determining the rates and activation parameters for organo-metal bond homolysis events that produce freely diffusing radicals. Application of this method is illustrated by measurement of bond homolysis activation parameters for a series of organo-cobalt porphyrin complexes ((TPP)Co-C(CH3)2CN (ΔH = 19.5±0.9 kcal mol−1, ΔS = 12±3 cal°K−1 mol−1), (TMP)Co-C(CH3)2CN (ΔH = 20±1 kcal mol−1S = 13±2 cal°K−1 mol−1), (TAP)Co-C(CH3)2CO2CH3H = 18.2±0.5 kcal mol−1, ΔS = 12±2 cal °K−1 mol−1), (TAP)Co-CH(CH3)C6H5H = 22.5±0.5, ΔS = 17±2 cal °K−1 mol−1)). The line broadening method is particularly useful in determining activation parameters for dissociation of weakly bonded organometallics where the rate of homolysis can exceed the range measurable by conventional chemical trapping methods.  相似文献   

5.
The first η2-olefinic monocarbon metallacarbone closo-2-(Ph3P)-1-N,2-[μ-(η2-CH2CH=Ch2)]-1-N-(σ-CH2CH=CH2)-2,1- RhCB10H10 has been prepared by the reaction of the dimeric anion {[Ph3PRhB10H10CNH2]2-μ-H}[PPN]+ with allyl bromide and characterized by a combination of spectroscopic methods and a single-crystal X-ray diffraction study. The variable temperature 1H and 13C NMR studies revealed the fluxional behavior of the η2-olefinic complex in CD2Cl2 solution which is associated with the allyl side-chain exchange process.  相似文献   

6.
Single crystal X-ray diffraction studies of trans-[(Ph3P)2Pd(Ph)X] (X = F (1), Cl (2), Br (3), and I (4) were carried out. The four structures split in two isostructural and isomorphous groups, namely orthorhombic for 1 and 2 (space group Pbca, Z = 8) and triclinic for 3 and 4 (space group P-1, Z = 2). According to the Pd---C bond length, the trans influence of X within these pairs follows the trend Cl>F and 1>Br. However, the trans influence of Cl is slightly stronger than that of Br. Both structural and 13C NMR studies revealed that electron-donating effects of (Ph3P)2PdX increase along the series X=I− for the Pd centre in [(Ph3P)2Pd(Ph)] were studied by 31P NMR in rigorously anhydrous CH2Cl2 solutions, and equilibrium constants and ΔG values were obtained for all possible combinations. The sequence F > Cl > Br > I is characteristic of halide preference for the Pd complexes. Dissolving 1 and PPN Cl in dry CH2Cl2 resulted in the release of ‘naked’ F which fluorinated the solvent smoothly to give a mixture of CH2ClF and CH2F2 in high yield. When chloroform was used instead of CH2Cl2, dichlorocarbene was generated slowly, forming the corresponding cyclopropane in the presence of styrene. All observations were rationalized successfully in terms of the filled/filled effect and push/pull interactions.  相似文献   

7.
Rotational barriers about the M-S bonds of 16-electron bent metallocene monothiolates (η5-C5H5)2Zr(Cl) (SR) (R = −CH3, −CH2CH3, −CH(CH3)2, −C(CH3)3) (1a–d) have been measured by dynamic 1H NMR methods: 32, 33, 35 and 26 kJ mol−1, respectively. The ground-state orientation about the Zr-S bonds of 1 that maximizes Spπ → Mdπ bonding (Cl-Zr-S-R ≈ 90°) also maximizes CpR steric interaction, whereas the rotational transition-state orientation (Cl-Zr-S-R ≈ 0°) is one that minimizes Spπ → Mdπ bonding and maximizes ClR steric interaction. Deviation from a ground-state orientation that is ideal for Spπ → Mdπ bonding might be expected as the size of the R group and CpR steric interaction increases. Thus, the aberrant trend for the R = −C(CH3)3 derivative could be attributed to a ground-state steric effect where the sterically demanding −C(CH3)3 group forces an unfavorable (misdirected) orientation for Mdπ-Spπ bonding, but a favorable orientation with respect to CpR and ClR steric interactions. However, the solid-state structures of (η5-C5H5)2Zr(SR)2 (R = −CH3, −CH2CH3, −CH(CH3)2, −C(CH3)3) (2a–d) exhibit regular variation of their metric parameters as evidenced by their Zr-S-C bond angles of 108, 109, 113, and 124° and S-Zr-S′ bond angles of 97, 99, 100 and 106°, respectively. Neither the S′-Zr-S-R torsion angles nor the dihedral angles that describe the relationship between the S/Zr/S′ and Cp(centroid)/Zr/Cp′ (centroid) planes (both indicators of the relative orientation of the Zr dπ acceptor orbital and the thiolate S pπ donor orbital) reflect the steric demand of the R group. Thus, the size of the R group imposes a measured effect on the geometry of 2 and the tert-butyl group is not extraordinary. Although the enthalpic and entropic effects could not be deconvoluted for rotation about the Zr-S bond of 1 in the present study, literature precedents suggest that both enthalpic and entropic effects may play a role in determining the irregular trend that is observed.  相似文献   

8.
The reaction of RuCl3(H2O), with C5Me4CF3J in refluxing EtOH gives [Ru25-C5Me1CF2)2 (μ-Cl2] (20 in 44% yield. Dimer 2 antiferromagnetic (−2J=200 cm1). The crystal structures of 2 (rhombohedral system, R3 space group, Z=9, R=0.0589) and [Rh25-C5Me4CF3(2Cl2(μ-Cl)2] (3) (rhombohedral system. space group, Z = 9, R = 0.0641) were solved; both complexes have dimeric structures with a trans arrangement of the η5-C5Me4CF4 rings. Comparison of the geometry of 2 and 3 with those of the corresponding η5-C5Me5 complexes shows that lowering the ring symmetry causes significant distortion of the M2(μ-Cl)2 moiety. The analysis of the MCl3 fragment conformations in 2 and 3 and in the η5-C5ME5 analogues shows that they are correlated with the M---M distances. The Cl atoms are displaced by Br on reaction of 2 with KBr in MeOH to give the diamagnetic dimer [Ru25-C5Me4CF3)2Br2 (μ-Br2] (4). Complex 2 reacts with O2 in CH2Cl2 solution at ambient temperature to form a mixture of isomeric η6-fulvene dimers [Ru26-C5Me3CF3 = CH2)2Cl2(μ-Cl)2] (5). Reactions of 5 with CO and allyl chloride give Ru(η5-C5Me3CF3CH2Cl)(CO)2Cl (6) and Ru(η5-C5Me3CF3CF3CH2Cl)(η3-C3H5)Cl2 (7) respectively.  相似文献   

9.
The observation of homolytic S---CH3 bond cleavage in (Ph2P(o-C6H4)SCH3)2Ni0 under photochemical conditions has prompted further investigation of nickel(0) complexes and their stability. Tetradentate P2S′2 donor ligands (S′ = thioether type S donor) with aromatic rings incorporated into the P to S links, Ph2P(o-C6H4)S(CH2)3S(o-C6H4)PPh2 (arom-PSSP), or the S to S links, Ph2P(CH2)2SCH2(o-C6H4)CH2S(CH2)2PPh2 (PS-xy-SP), have been used to form four-coordinate, square planar nickel(II) complexes, [(arom-PSSP)Ni](BF4)2 (2) and [(PS-xy-SP)Ni](BF4)2 (3). The bidentate and tetradentate ligands, Ph2P(o-C6H4)SCH2CH3 (arom-PSEt) and Ph2P(CH2)2S(CH2)3S(CH2)2PPh2 (PSSP), give similar complexes, [(arom-PSEt)2Ni](BF4)2 (1) and [(PSSP)Ni](BF4)2 (4), respectively. Cyclic voltammograms of the Ni11 complexes in CH3CN show two reversible redox events assigned to and . The one-electron reduction product produced by stoichiometric amounts of Cp2Co can be characterized by EPR. At 100 K rhombic signals show hyperfine coupling to two phosphorus atoms. Complete bulk chemical reduction of complexes 1, 2, 3 and 4 with Na/Hg amalgam provided the corresponding nickel(0) complexes 1R, 2R, 3R and 4R which were isolated as red solutions or solids characterized by magnetic resonance properties and reaction products. Photolysis of these nickel(0) complexes leads to S-dealkylation to produce alkyl radicals and dithiolate nickel(II) complexes. Complex 3 crystallized in the monoclinic space group P2t/c with a=20.740(5), B=9.879(3), C=17.801(4) åA, ß=92.59(2)°, V=3644(2) Å3 and Z=4; complex 4: P21/c with A=13.815(4), B=13.815(4), C=15.457(5) åA, V=3365.4(14) Å3 and Z=4.  相似文献   

10.
Reaction of RuCl(η5-C5H5(pTol-DAB) with AgOTf (OTf = CF3SO3) in CH2Cl2 or THF and subsequent addition of L′ (L′ = ethene (a), dimethyl fumarate (b), fumaronitrile (c) or CO (d) led to the ionic complexes [Ru(η5-C5H5)(pTol-DAB)(L′)][OTf] 2a, 2b and 2d and [Ru(η5-C5H5)(pTol-DAB)(fumarontrile-N)][OTf] 5c. With the use of resonance Raman spectroscopy, the intense absorption bands of the complexes have been assigned to MLCT transitions to the iPr-DAB ligand. The X-ray structure determination of [Ru(η5-C5H5)(pTol-DAB)(η2-ethene)][CF3SO3] (2a) has been carried out. Crystal data for 2a: monoclinic, space group P21/n with A = 10.840(1), b = 16.639(1), C = 14.463(2) Å, β = 109.6(1)°, V = 2465.6(5) Å3, Z = 4. Complex 2a has a piano stool structure, with the Cp ring η5-bonded, the pTol-DAB ligand σN, σN′ bonded (Ru-N distances 2.052(4) and 2.055(4) Å), and the ethene η2-bonded to the ruthenium center (Ru-C distances 2.217(9) and 2.206(8) Å). The C = C bond of the ethene is almost coplanar with the plane of the Cp ring, and the angle between the plane of the Cp ring and the double of the ethene is 1.8(0.2)°. The reaction of [RuCl(η5-C5H5)(PPh)3 with AgOTf and ligands L′ = a and d led to [Ru(η5-C5H5)(PPh3)2(L′)]OTf] (3a) and (3d), respectively. By variable temperature NMR spectroscopy the rottional barrier of ethene (a), dimethyl fumarate (b and fumaronitrile (c) in complexes [Ru(η5-C5H5)(L2)(η2-alkene][OTf] with L2 = iPr-DAB (a, 1b, 1c), pTol-DAB (2a, 2b) and L = PPh3 (3a) was determined. For 1a, 1b and 2b the barrier is 41.5±0.5, 62±1 and 59±1 kJ mol−1, respectively. The intermediate exchange could not be reached for 1c, and the ΔG# was estimated to be at least 61 kJ mol. For 2a and 3a the slow exchange could not be reached. The rotational barrier for 2a was estimated to be 40 kJ mol. The rotational barier for methyl propiolate (HC≡CC(O)OCH3) (k) in complex [Ru(η5-C5H5)(iPr-DAB) η2-HC≡CC(O)OCH3)][OTf] (1k) is 45.3±0.2 kJ mol−1. The collected data show that the barrier of rotational of the alkene in complexes 1a, 2a, 1b, 2b and 1c does not correlate with the strength of the metal-alkene interaction in the ground state.  相似文献   

11.
The enthalpies of reaction of HMo(CO)3C5R5 (R = H, CH3) with diphenyldisulfide producing PhSMo(CO)3C5R5 and PhSH have been measured in toluene and THF solution (R = H, ΔH= −8.5 ± 0.5 kcal mol−1 (tol), −10.8 ± 0.7 kcal mol−1 (THF); R = CH3, ΔH = −11.3±0.3 kcal mol−1 (tol), −13.2±0.7 kcal mol−1 (THF)). These data are used to estimate the Mo---SPh bond strength to be on the order of 38–41 kcal mol−1 for these complexes. The increased exothermicity of oxidative addition of disulfide in THF versus toluene is attributed to hydrogen bonding between thiophenol produced in the reaction and THF. This was confirmed by measurement of the heat of solution of thiophenol in toluene and THF. Differential scanning calorimetry as well as high temperature calorimetry have been performed on the dimerization and subsequent decarbonylation reactions of PhSMo(CO)3Cp yielding [PhSMo(CO)2Cp]2 and [PhSMo(CO)Cp]2. The enthalpies of reaction of PhSMo(CO)3Cp and [PhSMo(CO)2Cp]2 with PPh3, PPh2Me and P(OMe)3 have also been measured. The disproportionation reaction: 2[PhSMo(CO)2Cp]2 → 2PhSMo(CO)3Cp + [PhSMP(CO)Cp]2 is reported and its enthalpy has also been measured. These data allow determination of the enthalpy of formation of the metal-sulfur clusters [PhSMo(CO)nC5H5]2, N = 1,2.  相似文献   

12.
The reactions of [(H5C6)3P]2ReH6 with (CH3CN)3Cr(CO)3, (diglyme)Mo(CO)3 or (C3H7CN)3W(CO)3 led to the formation of [(H5C6)3P]2ReH6M(CO)3 (M = Cr, Mo, W) complexes. These have been characterized by IR and NMR spectroscopies, as well as elemental analyses. A single crystal X-ray diffraction study has also been carried out for the M = Cr complex as a K(18-crown-6)+ salt. The complex crystallizes as a THF monosolvate in the monoclinic space group P21/n with a = 22.323(6), B = 9.523(2), C = 27.502(5) Å, β = 104.98(2)0 and V = 5648 Å3 for Z = 4. The Re---Cr separation is 2.5745(12) Å, and the two phosphine ligands are oriented unsymmetrically. Although the hydride ligands were not found, the presence of three bridging hydrides and a dodecahedral coordination geometry about rhenium could be inferred. Low temperature 1H and 31P NMR spectroscopic studies did not reveal the low symmetry of the solid state structure.  相似文献   

13.
Kinetic results are reported for intramolecular PPh3 substitution reactions of Mo(CO)21-L)(PPh3)2(SO2) to form Mo(CO)22-L)(PPh3)(SO2) (L = DMPE = (Me)2PC2H4P(Me)2 and dppe=Ph2PC2H4PPh2) in THF solvent, and for intermolecular SO2 substitutions in Mo(CO)32-L)(η2-SO2) (L = 2,2′-bipyridine, dppe) with phosphorus ligands in CH2Cl2 solvent. Activation parameters for intramolecular PPh3 substitution reactions: ΔH values are 12.3 kcal/mol for dmpe and 16.7 kcal/mol for dppe; ΔS values are −30.3 cal/mol K for dmpe and −16.4 cal/mol K for dppe. These results are consistent with an intramolecular associative mechanism. Substitutions of SO2 in MO(CO)32-L)(η2-SO2) complexes proceed by both dissociative and associative mechanisms. The facile associative pathways for the reactions are discussed in terms of the ability of SO2 to accept a pair of electrons from the metal, with its bonding transformations of η2-SO2 to η1-pyramidal SO2, maintaining a stable 18-e count for the complex in its reaction transition state. The structure of Mo(CO)2(dmpe)(PPh3)(SO2) was determined crystallographically: P21/c, A=9.311(1), B = 16.344(2), C = 18.830(2) Å, ß=91.04(1)°, V=2865.1(7) Å3, Z=4, R(F)=3.49%.  相似文献   

14.
The kinetics of the displacement reactions of the bromide ligands of trans-[FeBr2(depe)2] (depe = Et2PCH2CH2PEt2) by the organonitrile NCCH2C6H4OMe-4, in tetrahydrofuran (either in the absence or in the presence of added Br), to give the corresponding mono- and dinitrile complexes trans-[FeBr(NCCH2C6H4OMe-4)(depe)2]+ and trans-[Fe(NCCH2C6H4OMe-4)2(depe)2]2+, have been investigated by stopped-flow spectrophotometry. The substitution reaction occurs by a mechanism involving rate-limiting dissociation of bromo ligands to form the unsaturated intermediates [FeBr(depe)2]+ (k1 = 1.52 ± 0.02 s−1) and [Fe(NCR)(depe)2]2+ (k3 = 0.063 ± 0.008 s−1) which add the nitrile ligand to form those nitrile complexes. The competition between the nitrile and Br for such metal centres has also been investigated and a stronger inhibiting effect of added Br is observed for the substitution of the second bromo ligand relative to the first one. The kinetic data are rationalized in terms of π-electronic effects of these unsaturated metal centres and of the bromide and nitrile ligands.  相似文献   

15.
The reactions of various proton donors (phenol, hexafluoro-2-propanol, perfluoro-2-methyl-2-propanol, monochloroacetic acid, and tetrafluoroboric acid) with the rhenium (I) hydride complex [(triphos)Re(CO)2H] (1) have been studied in dichloromethane solution by in situ IR and NMR spectroscopy. The proton donors from [(triphos)Re(CO)2H…HOR] adducts exhibiting rather strong H…H interactions. The enthalpy variations associated with the formation of the H-bonds (−ΔH = 4.4–6.0 kcal mol−1) have been determined by IR spectroscopy, while the H…H distance in the adduct [(triphos)Re(CO)2H…HOC(CF3)3] (1.83 Å) has been calculated by NMR spectroscopy through the determination of the T1min relaxation time of the Re---H proton. It has been shown that the [(triphos)Re(CO)2H…HOR] adducts are in equilibrium with the dihydrogen complex [(triphos)Re(CO)22-H2)]+, which is thermodynamically more stable than any H-bond adduct.  相似文献   

16.
The solution of [RhCl(PPh3)3] in acidic 1-ethyl-3-methylimidazolium chloroaluminate(III) ionic liquid (AlCl3 molar fraction, xAlCl3=0.67) was investigated by 1H and 31P{1H} NMR. One triphenyl phosphine is lost from the complex and is protonated in the acidic media, and cis-[Rh(PPh3)2ClX], (2), where X is probably [AlCl4], is formed. On, standing, 2 is converted to trans-[Rh(H)(PPh3)2X], (3). The reaction of 2 and H2 also produces trans-[Rh(H)(PPh3)2X], (3). 1H and 31P{1H} NMR support the suggestion that a weak ligand such as [AlCl4], present in solution may interact with the metal centre. When [RhCl(PPh3)3] is dissolved in CH2Cl2/AlCl3/HCl for comparison, two exchanging isomers of what is probably [RhH{(μ-Cl)2AlCl2}{(μ-Cl)AlCl3}(PPh3)2], (6) and (7), are formed.  相似文献   

17.
Reactions of [Rh(COD)Cl]2 with the ligand RN(PX2)2 (1: R = C6H5; X = OC6H5) give mono- or disubstituted complexes of the type [Rh2(COD)Cl22−C6H5N(P(OC6H5)2)2}] or [RhCl{ν2−C6H5 N(P(OC6H5)2)2 }]2 depending on the reaction conditions. Reaction of 1 with [Rh(CO)2Cl]2 gives the symmetric binuclear complex, [Rh(CO)Cl{μ−C6H5N(P(OC6H5)2)2} 2, whereas the same reaction with 2 (R = CH3; X = OC6H5) leads to the formation of an asymmetric complex of the type [Rh(CO)(μ−CO)Cl{μ−CH3N(P(OC6H5)2)2}2 containing both terminal and bridging CO groups. Interestingly the reaction of 3 (R = C6H5, X = OC6H4Br−p with either [Rh(COD)Cl]2 or [Rh(CO)2Cl]2 leads only to the formation of the chlorine bridged binuclear complex, [RhCl{ν2−C6H5N(P(OC6H4Br−p)2)2}]2. The structural elucidation of the complexes was carried out by elemental analyses, IR and 31P NMR spectroscopic data.  相似文献   

18.
Rapid reactions occur between [OsVI(tpy)(Cl)2(N)]X (X = PF6, Cl, tpy = 2,2′:6′,2″-terpyridine) and aryl or alkyl phosphi nes (PPh3, PPh2Me, PPhMe2, PMe3 and PEt3) in CH2Cl2 or CH3CN to give [OsIV(tpy)(Cl)2(NPPh3)]+ and its analogs. The reaction between trans-[OsVI(tpy)(Cl)2(N)]+ and PPh3 in CH3CN occurs with a 1:1 stoichiometry and a rate law first order in both PPh3 and OsVI with k(CH3CN, 25°C) = 1.36 ± 0.08 × 104 M s−1. The products are best formulated as paramagnetic d4 phosphoraniminato complexes of OsIV based on a room temperature magnetic moment of 1.8 μB for trans-[OsIV(tpy)(Cl)2(NPPh3)](PF6), contact shifted 1H NMR spectra and UV-Vis and near-IR spectra. In the crystal structures of trans-[OsIV(tpy)(Cl)2( NPPh3)](PF6)·CH3CN (monoclinic, P21/n with a = 13.384(5) Å, b = 15.222(7) Å, c = 17.717(6) Å, β = 103.10(3)°, V = 3516(2) Å3, Z = 4, Rw = 3.40, Rw = 3.50) and cis-[OsIV(tpy)(Cl)2(NPPh2Me)]-(PF6)·CH3CN (monoclinic, P21/c, with a = 10.6348(2) Å, b = 15.146(9) ÅA, c = 20.876(6) Å, β = 97.47(1)°, V = 3334(2) Å3, Z = 4, R = 4.00, Rw = 4.90), the long Os-N(P) bond lengths (2.093(5) and 2.061(6) Å), acute Os-N-P angles (132.4(3) and 132.2(4)°), and absence of a significant structural trans effect rule out significant Os-N multiple bonding. From cyclic voltammetric measurements, chemically reversible OsV/IV and OsIV/III couples occur for trans-[OsIV(tpy)(Cl)2(NPPh3)](PF6) in CH3CN at +0.92 V (OsV/IV) and −0.27 V (OsIV/III) versus SSCE. Chemical or electrochemical reduction of trans-[OsIV(tpy)(Cl)2(NPPh3)](PF6) gives isolable trans-OsIII(tpy)(Cl)2(NPPh3). One-electron oxidation to OsV followed by intermolecular disproportionation and PPh3 group transfer gives [OsVI(tpy)Cl2(N)]+, [OSIII(tpy)(Cl)2(CH3CN)]+ and [Ph3=N=PPh3]+ (PPN+). trans-[OsIV(tpy)(Cl)2(NPPh3)](PF6) undergoes reaction with a second phosphine under reflux to give PPN+ derivatives and OsII(tpy)(Cl)2(CH3CN) in CH3CN or OsII(tpy)(Cl)2(PR3) in CH2Cl2. This demonstrates that the OsVI nitrido complex can undergo a net four-electron change by a combination of atom and group transfers.  相似文献   

19.
A series of cationic nickel complexes [(η3-methally)Ni(PP(O))]SbF6 (1–4) [PP(O) = Ph2P(CH2)P(O)Ph2 (dppmO) (1), Ph2P(CH2)2P(O)Ph2 (dppeO) (2), Ph2P(CH2)3P(O)Ph2 (dpppO) (3), pTol2P(CH2)P(O)pTol2 (dtolpmO) (4)] has been synthesized in good yields by treatment of [(η3-methally)NiBr]2 with biphosphine monoxides and AgSbF6. The ligands are coordinated in a bidentate way. Starting from [(η3-all)PdI]2 the cationic complexes [(η3-all)PP(O))]Y (8–14). [PP(O) = dppmO, dppeO, dpppO, dtolpmO;Y = BF4, SbF6, CF3SO3, pTolSO3] were synthesized in good yields. The coordination mode of the ligand is dependent on the backbone and the anion, revealing a monodentate coordination with dppmO for stronger coordinating anions. The intermediates [(η3-all)Pd(I)(PP(O)-κ1-P)] (5–7) [PP(O) = dppmO (5), dppeO (6), dtolpmO (7)] were isolated and characterized. Neutral methyl complexes [(Cl)(Me)Pd(PP(O))] (15–18). [PP(O) = dppmO (15), dppeO (16), dpppO (17), dtolpmO (18)] can easily be obtained in high yields starting from [(cod)PdCl2]. For dppmO two different routes are presented. The structure of [(Me)(Cl)Pd{;Ph2P(CH2-P(O)Ph22-P,O};] · CH2Cl2 (15) with the chlorine atom trans to phosphorus was determined by X-ray diffraction.  相似文献   

20.
The reaction of meso-tetrakis (4-dimethoxyphenyl) porphinatomanganese(II), MnTPOMeP, with TCNE (TCNE = tetracyanoethylene) leads to the formation of [MnTPOMeP]+ [TCNE] and [MnTPOMeP]+[OC(CN)C(CN)2]. The single-crystal X-ray structures of the latter as well as [Cu(bipy)2Cl]+ [OC(CN)C(CN)2] were determined. The former has a disordered [OC(CN)C(CN)2] bridging via C and O between a pair of MnIII sites, whereas the latter has an isolated [OC(CN)C(CN)2] unbound to CuII. The IR characterization for μ2-C,O bound [OC(CN)C(CN)2] is at 2219m and 2196s (νCN) cm−1 and at 1558s (νCO) cm−1 while for unbound [OC(CN)C(CN)2] it is at 2210m, 2203m, 2181m (νCN) cm−1 and at 1583s (νCO) cm−1.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号