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 15 毫秒
1.
Bis(ferrocenyl)-substituted allenylidene complexes, [(CO)5MCCCFc2] (1a-c, Fc = (C5H4)Fe(C5H5), M = Cr (a), Mo (b), W (c)) were obtained by sequential reaction of Fc2CO with Me3Si-CCH, KF/MeOH, n-BuLi, and [(CO)5M(THF)]. For the synthesis of related mono(ferrocenyl)allenylidene chromium complexes, [(CO)5CrCCC(Fc)R] (R = Ph, NMe2), three different routes were developed: (a) reaction of the deprotonated propargylic alcohol HCCC(Fc)(Ph)OH with [(CO)5Cr(THF)] followed by desoxygenation with Cl2CO, (b) Lewis acid induced alcohol elimination from alkenyl(alkoxy)carbene complexes, [(CO)5CrC(OR)CHC(NMe2)Fc], and (c) replacement of OMe in [(CO)5CrCCC(OMe)NMe2] by Fc. Complex 1a was also formed when the mono(ferrocenyl)allenylidene complex [(CO)5CrCCC(Fc)NMe2] was treated first with Li[Fc] and the resulting adduct then with SiO2. The replacement route (c) was also applied to the synthesis of an allenylidene complex (7a) with a CC spacer in between the ferrocenyl unit and Cγ of the allenylidene ligand, [(CO)5CrCCC(NMe2)-CCFc]. The related complex containing a CHCH spacer (9a) was prepared by condensation of [(CO)5CrCCC(Me)NMe2] with formylferrocene in the presence of NEt3. The bis(ferrocenyl)-substituted allenylidene complexes 1a-c added HNMe2 across the Cα-Cβ bond to give alkenyl(dimethylamino)carbene complexes and reacted with diethylaminopropyne by regioselective insertion of the CC bond into the Cβ-Cγ bond to afford alkenyl(diethylamino)allenylidene complexes, [(CO)5MCCC(NEt2)CMeCFc2]. The structures of 5a, 7a, and 9a were established by X-ray diffraction studies.  相似文献   

2.
Oxidative addition of 1-bromo-1H-indene to [Mo(CO)3(NCMe)3] and [W(CO)3(NCEt)3] is a suitable method for preparation of the indenyl compounds [IndMo(CO)3Br] and [IndW(CO)3Br], respectively. These products were fully characterised using spectroscopic methods. Structure of [IndW(CO)3Br] was determined by single crystal X-ray diffraction analysis.  相似文献   

3.
《Inorganica chimica acta》2002,327(1):169-178
New complexes [MI(CO)2(dppe){S2P(OEt)2}] (M=W, 1a; M=Mo, 1b), [MI(CO)2(dppm){S2P(OEt)2}] (M=W, 2a; M=Mo, 2b) and [W(CO)(dppe){S2P(OEt)2}2][O2dppe] (3a), were synthesised from [MI2(CO)3(NCMe)2] (M=Mo, W), after treatment with ammonium diethyldithiophosphate and phosphine under different conditions. The structure of the tungsten complexes was determined by single crystal X-ray diffraction. During the synthesis of 3a, oxidation of the phosphine took place and a molecule of oxidised phosphine occupies channels in the crystal. DFT/B3LYP calculations on models of 1a and 2a showed the capped octahedron structure, observed in most dicarbonyl complexes of this family, to be preferred by 1.4 and 2.6 kcal mol−1 for the dppm and the dppe complexes, respectively. Strong steric repulsions can reverse this trend, as happens with the rigid dppm ligand. Complex 1a adopts a pentagonal bipyramidal geometry, which is often found in related monocarbonyl complexes.  相似文献   

4.
The complex [Et4N][W(CO)5OMe] (1) has been prepared from the reaction of the photochemically generated W(CO)5THF adduct and [Et4N][OH] in methanol. Complex 1 was shown to undergo rapid CO dissociation in THF to quantitatively provide the dimeric dianion, [W(CO)4OMe]22−. The resulting THF insoluble salt [Et4N]2[W(CO)4OMe]2 (2) has been structurally characterized by X-ray crystallography, with the doubly bridging methoxide ligands being in an anti configuration. Complex 2 was found to subsequently react with excess methoxide ligand in a THF slurry to afford the face-sharing octahedron complex [Et4N]3[W2(CO)6(OMe)3] (3) which contains three doubly bridging methoxide groups. In the absence of excess methoxide ligand complex 2 cleanly yields the tetrameric complex [Et4N]4[W(CO)3OMe]4 (4) which has been structurally characterized as a cubane-like arrangement with triply bridging μ3-methoxide groups and W(CO)3 units. Although complex 3 was not characterized in the solid state, the closely related glycolate derivative [Et4N]3[W2(CO)6(OCH2CH2OH)3] (5) was synthesized and its structure determined by X-ray crystallography. The trianions of complex 5 are linked in the crystal lattice by strong intermolecular hydrogen bonds. Crystal data for 2: space group P21/n, a = 7.696(2), b = 22.019(4), c = 9.714(2) Å, β = 92.22(3)°, Z = 4, R = 6.43%. Crystal data for 4: space group Fddd, a = 12.433(9), b = 24.01(2), c = 39.29(3) Å, Z = 8, R = 8.13%. Crystal data for 5: space group P212121, a = 11.43(2), b = 12.91(1), c = 29.85(6) Å, Z = 8, R = 8.29%. Finally, the rate of CO ligand dissociation in the closely related aryloxide derivatives [Et4N][W(CO)5OR] (R = C6H5 and 3,5-F2C6H3) were measured to be 2.15 × 10−2 and 1.31 × 10−3 s−1, respectively, in THF solution at 5°C. Hence, the value of the rate constant of 2.15 × 10−2 s−1 establishes a lower limit for the first-order rate constant for CO loss in the W(CO)5OMe anion, since the methoxide ligand is a better π-donating group than phenoxide.  相似文献   

5.
A variety of Group 6 mono bipyridine (bpy) complexes were prepared, and substitution reactions of [(bpy)(MeIm)M(CO)2(NO)]PF6 complexes (MeIm = 1-methylimidazole, M = W or Mo) were investigated. Nitrosylation of complexes having the general formula (bpy)(L)M(CO)3 (L = a variable ligand) gave cationic complexes of the form [(bpy)(L)M(CO)2(NO)]PF6. The structure of [(bpy)(MeIm)W(CO)2(NO)]PF6 was confirmed by single-crystal X-ray diffractometry. [(bpy)(MeIm)M(CO)2(NO)]PF6 complexes undergo facile substitutions with mono-, tri- and tetra-dentate ligands, yielding di- or mono-carbonyl mononitrosyl complexes. The structures of [(bpy)(PMe3)2W(CO)(NO)]PF6 and [(dien)(PMe3)W(CO)(NO)]PF6 (dien = diethylenetriamine) were determined by X-ray diffraction.  相似文献   

6.
An improved synthetic procedure for pentabenzylcyclopentadiene Bz5C5H was developed. Six new organomolybdenum and organotungsten halides η5-Bz5C5M(CO)3X(M = Mo, W; X = Cl, Br, I) were syntesized through the reaction of η5-Bz5C5M(CO)3Li (derived from Bz5C5H, n-BuLi and M(CO)6) with PCl3, PBr3 or I2 and characterized by elemental analysis, IR and 1H NMR spectroscopy. The structure of η5-Bz5C5Mo(CO)3I was determined by single-crystal X-ray diffraction techniques. It crystallized in the monoclinic space groupp P2/c with cell parameters a = 13.294(4), B = 15.147(4), C = 19.027(3) Å, β = 108.32(2)°, V = 3637(2) Å3, Z = 4 and Dx = 1.50 g cm−3. The final R value was 0.035 for 4564 observed reflections.  相似文献   

7.
The reactions of [Mo(CO)6] towards a 2,6-di(imino)pyridine L1 and related ligands were studied. The reaction with L1 afforded two new complexes, [Mo(CO)4L1] (1) and [Mo(CO)4L2] (2), where L2 is the 2-amino-6-iminopyridine ligand arising from the hydrogenation of one imine function of L1; similar reaction with a 2-acetyl-6-iminopyridine ligand L3 afforded [Mo(CO)4L3] (3). Compounds 1, 2 and 3 have been fully characterised by IR, 1H NMR and X-ray crystallography; they present a metal ion in a pseudo-octahedral environment, the three organic ligands acting with bidentate N2 coordination modes. One of the imine functions in 1, the amine function in 2, and the ketone function in 3 are uncoordinated.  相似文献   

8.
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   

9.
The reaction of the metal complexes MO2Cl2(mebipy) (M = Mo, W) with two equivalents thiophenol by the exact same procedure leads to two different products for molybdenum [Mo2O4(SPh)2(mebipy)2] and tungsten [WO2(SPh)2(mebipy)]. To understand why this is the case the redox potentials of the starting materials were measured showing that the redox potential for thiophenol is lower than the redox potentials (MV ↔ MVI) for both of the metal precursors. A reduction of the metal and oxidation of the sulfur should be possible for both reactions but occurs only for the molybdenum compound. Theoretical calculations show that different metal-sulfur bond strengths are as well and equally responsible for the differing reaction behaviour as are the redox potentials.  相似文献   

10.
Four new tungsten-oxo(VI) complexes have been synthesized, characterized spectroscopically and their molecular structure established by X-ray diffraction analysis. These bear identical environment as previously reported molybdenum-oxo(VI) complexes, which allowed direct comparison of their spectroscopic properties. Their capability as oxygen atom transfer agents was found to be significantly lower than their molybdenum analogs.  相似文献   

11.
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   

12.
Adducts with MoO42− tetrahedra coordinated to Cr(III) or Co(III) complexes have been synthesized and studied by IR and high resolution 95Mo NMR spectroscopy. The 95Mo chemical shifts of the adducts with cobalt(III) lie in the range −33.2 to + 49.4 ppm. This may be compared with an overall known chemical shift range in excess of 7000 ppm and implies a similarity in the molybdenum environment in all cases. For adducts with chelated cobalt(III) complexes several rather broad 95Mo singnals are obtained with linewidths up to 260 Hz.  相似文献   

13.
Red-black [TpiPr∗MoVO]2(μ-O)(μ-MoVIO4) (1, TpiPr∗ = hydrobis(3-isopropylpyrazolyl)(5-isopropylpyrazolyl)borate) has been isolated as a by-product in the synthesis of NEt4[TpiPrMo(CO)3] (TpiPr = hydrotris(3-isopropylpyrazolyl)borate) and characterized by spectroscopic and X-ray crystallographic techniques. The trinuclear, mixed-valence complex contains two distorted octahedral anti-TpiPr∗MoVO centers bridged by bent oxo (Mo-O-Mo av. 158.7°) and tetrahedral κO,κO′-molybdate ligands. The complex contains a six-membered, non-planar Mo3(μ-O)3 core and two 1,2-borotropically-shifted TpiPr∗ ligands (with the shifted pyrazolyl trans to MoV=O). Aerial decomposition of solid NEt4[TpiPrMo(CO)3] produces sky-blue, diamagnetic TpiPrMoO(iPrpz)(iPrpzH) (2, iPrpz- = 3-isopropylpyrazolate, iPrpzH = 3-isopropyl-2H-pyrazole). Molecules of 2 feature a tridentate fac-TpiPr ligand and mutually cis terminal oxo (MoO = 1.665(2) Å) and monodentate iPrpz and iPrpzH ligands. The latter are formed by B-N bond cleavage of TpiPr. The complex can also be synthesized by reacting NEt4[TpiPrMo(CO)3] with excess 3-isopropylpyrazole and dioxygen at 100 °C. Cleavage of the B-N bond(s) of TpiPr was also observed in the formation of TpiPrMoO(SPh)(iPrpzH) (3) as a by-product in the synthesis of TpiPrMoO2(SPh). In the monohydrate, 3 exhibits a distorted octahedral geometry defined by a tridentate fac-TpiPr ligand and mutually cis terminal oxo (MoO = 1.676(3) Å) and monodentate SPh and iPrpzH ligands. The pyrazole β-NH group is observed to participate in a hydrogen-bond to the lattice water molecule. The complex can be synthesized in high yield by reducing TpiPrMoO2(SPh) by HSPh or PPh3 in the presence of excess 3-isopropylpyrazole.  相似文献   

14.
Rh(I) and Ir(I) complexes of the type [Rh(cod)(η2-TMPP)]1+ (1) and M(cod)(η2-TMPP-O) (M = Rh (2), Ir (3); cod = cyclooctadiene; TMPP = tris(2,4,6-trimethoxyphenyl)phosphine; TMPP-O = mono-demethylated form of TMPP) have been isolated from reactions of [M(cod)Cl]2 with M′BF4 (M′ = Ag+, K+, Na+) followed by addition of the tertiary phosphine ligand. This chemistry is dependent on the identity of the metal, as both the cationic phosphine complex and the neutral phosphino-phenoxide compound are stable for Rh(I), whereas only the latter is stable for Ir(I). The three complexes have been characterized by IR and NMR (1H and 31P) spectroscopies as well as by cyclic voltammetry. The 1H NMR spectrum of [Rh(cod)(η2-TMPP)]1+ (1) is in accord with the formula and reveals that the TMPP phenyl rings are undergoing rapid exchange between coordinated and non-coordinated modes; the corresponding spectra of 2 and 3 support free rotation about the P---C bonds of the unbound phenyl rings with no fluxionality of the bound demethylated ring. The 31P{1H} NMR spectrum of the neutral species 2 exhibits a significant upfield shift with respect to the analogous cationic compound 1. This shielding is the result of improved electron donation to the metal from a phenoxide group as compared to an ether substituent. In situ addition of CO to the reaction between TMPP and [Rh(cod)Cl]2 or [Ir(cod)Cl]2 in the presence of M′BF4 results in the isolation of the monocarbonyl species [Rh(TMPP)(η2-TMPP)(CO)][BF4] (5) and the stable dicarbonyl compound [Ir(TMPP)2(CO)2][BF4] (4), respectively. Single crystal X-ray data for . The geometry of 4 is square planar, with essentially ideal angles for the mutually trans disposed phosphine and carbonyl ligands, as found in earlier studies for the analogous Rh dicarbonyl compound. The 1H NMR spectrum of 4 supports the assignment of magnetically equivalent phosphorus nuclei in solution. The results of this study indicate that cyclooctadiene is a particularly strong ligand for monovalent late transition metals ligated by TMPP, to the extent that it is inert with respect to substitution in the absence of π-acceptor ligands such as carbon monoxide.  相似文献   

15.
Reaction of [MoVI(TpMe,Me)(O)2Cl] with a variety of pyridine-based ligands [pyridine (py), 4,4′-bipyridine (bpy), 4-phenylpyridine (phpy) and 1,2′-bis(4-pyridyl)ethene (bpe)] in toluene in the presence of Ph3P affords the mononuclear oxo-Mo(IV) complexes [Mo(TpMe,Me)(O)Cl(L)] (L=py, phpy or monodentate bpy; abbreviated as Mo(py), Mo(phpy) and Mo(bpy), respectively) and the dinuclear complexes [{Mo(TpMe,Me)(O)Cl}2(μ-L)] (L=bpy, bpe; abbreviated as Mo2(bpy), Mo2(bpe), respectively). The complex Mo2(bpy), together with the by-product [{Mo(TpMe,Me)(O)Cl}2(μ-O)], have been crystallographically characterised. Electrochemical studies on the oxo-Mo(IV) complexes reveal the presence of reversible Mo(IV)/Mo(V) couples at around −0.3 V versus ferrocene/ferrocenium in every case. For the dinuclear complexes Mo2(bpy) and Mo2(bpe) these redox processes are coincident, indicating that they are largely metal-centred and not significantly delocalised across the bridging ligand. In contrast, Mo2(bpe) alone shows two reversible reductions, separated by 320 mV; these could be described as ligand-centred reductions of the bpe bridge, or as Mo(IV)/Mo(III) couples which—because of their separation—are substantially delocalised onto the bridging ligand. UV-Vis spectroelectrochemical studies using an OTTLE cell at 243 K revealed that oxidation of the complexes results in spectral changes (collapse of the Mo(IV) d-d transitions, loss in intensity of the Mo→pyridine MLCT transition) consistent with the formation of a Mo(V) state following metal-centred oxidation, but that one-electron reduction of Mo2(bpe) results in appearance of numerous intense transitions more characteristic of a ligand radical following ligand-centred reduction.  相似文献   

16.
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   

17.
Several novel dimers of the composition [M2Cl4(trans-dppen)2] (M=Ni (1), Pd (2), Pt (3)) containing trans-1,2-bis(diphenylphosphino)ethene (trans-dppen) have been prepared and characterized by X-ray diffraction methods, NMR spectroscopy (195Pt{1H}, 31P{1H}), elemental analyses, and melting points. The intramolecular [2+2] photocycloaddition of the two diphosphine-bridges in 3 produces [Pt2Cl4(dppcb)] (4), where dppcb is the new tetradentate phosphine cis,trans,cis-1,2,3,4-tetrakis(diphenylphosphino)cyclobutane. Neither 1 nor the free diphosphine trans-dppen shows this reaction. In the case of 2 the photocycloaddition is slower than in 3. This difference can be explained by the shorter distance between the two aliphatic double bonds in 3 than in 2, but also different transition probabilities within ground and excited states of the used metals could be involved. Furthermore, variable-temperature 31P{1H} NMR spectroscopy of 2 or 3 reveals a negative activation entropy of 2 for the [2+2] photocycloaddition, but a positive of 3. The removal of chloride from 4 by precipitating AgCl with AgBF4, and subsequent treatment with 2,2′-bipyridine (bipy) or 1,10-phenanthroline (phen) leads to [Pt2(dppcb)(bipy)2](BF4)4 (5) and [Pt2(dppcb)(phen)2](BF4)4 (6), respectively. In an analogous reaction of 4 with PMe2Ph or PMePh2, [Pt2(dppcb)(PMe2Ph)4](BF4)4 (7) and [Pt2(dppcb)(PMePh2)4](BF4)4 (8) are formed. Complexes 1–8 show square–planar coordinations, where the compounds 4–8 have also been characterized by the above mentioned methods together with fast atom bombardment mass spectrometry (7, 8). The crystal structure of 4 reveals two conformations, which arise from an energetic competition between the sterical demands of dppcb and an ideal square–planar environment of Pt(II). The free tetraphosphine dppcb can be obtained easily from 4 by treatment with NaCN. It has been characterized fully by the above methods including 13C{1H} and 1H NMR spectroscopy. The X-ray structure analysis shows the pure MMMP-enantiomer in the solid crystal, which is therefore optically active. This chirality is induced by a conformation of dppcb, where all four PPh2 groups are non-equivalent. Variable-temperature 31P{1H} NMR spectroscopy of dppcb confirms this explanation, since the single signal at room temperature is split into two doublets at 183 K. The goal of this article is to demonstrate the facile production of a new tetradentate phosphine from a diphosphine precursor via Pt(II) used as a template.  相似文献   

18.
[Pt(COD)Cl2] (1) reacts with PPh2(C6H4COOH) (2a,b,c), PPh2(C6H4COONa) (2d), PPh(C6H4COOH)2 (4b,c) and P(C6H4COOH)3 (6b,c) with formation of the corresponding complexes [Pt(L)2Cl2] (3a,b,c,d, 5b,c, 7b,c). Halide abstraction from 3a by Ag+ promotes coordination of the ortho-carboxylate function to platinum, yielding [ -2)}{PPh2(C6H4COOH-2)}Cl] (bd8) and [ovbar|{PPh2(C6H4COO-2)}2] (bd9). Reaction of 1 with CO and 2a or 2b gives [Pt(CO)(L)Cl2] (10a,b), wherea 1 and 2,3-bis(diphenylphosphino) maleic anhydride yields (bd12) and [Pt{Ph2PC(COOH)=C(COOMe)-PPh2}Cl2] (13). The 1H, 13C and 31P NMR spectra are reported and discussed. The X-ray structural analysis of 3b showed the compound to be monoclinic, space group P21/n, Z=4, with a=1038.5(3), B=1792.6(4), C=2311.5(4) pm, β=91.6(2)° and Dcalc=1.353 g cm−3. The structure was solved from 4832 observed reflections with F0 > 4 σ(F0) and refined to a final R value of 0.0743. The Pt atom is surrounded by two Cl and two P atoms in a square planar arrangement.  相似文献   

19.
The cis effects of phosphine, arsine and stibine ligands have been evaluated by measuring the IR stretching frequency in dichloromethane of the carbonyl ligand in a series of Rh(I) Vaska-type complexes, trans-[RhCl(CO)(L)2]. These data were correlated with those obtained by Tolman for the electronic trans influences in the [Ni(L)(CO)3] complexes. The electronic contribution, χFc, of ferrocenyl was determined as 0.8 from these plots by evaluating PPh2Fc as ligand. In order to accommodate arsine and stibine ligands an additional correction term, to compensate for differences in the donor atom, was added to Tolman’s equation for calculation of the Tolman electronic parameter of phosphine ligands. In the resulting equation: ν(CONi)=2056.1+∑i=13χi+CL values for CL of CP=0, CAs=−1.5 and CSb=−3.1 are suggested for phosphine, arsine and stibine ligands, respectively. The crystal and molecular structures of trans-[RhCl(CO)(PPh2Fc)2] · 2C6H6, trans-[RhCl(CO){P(NMe2)3}2] and trans-[RhCl(CO)(AsPh3)2] are reported. The Tolman cone angles for PPh2Fc and P(NMe2)3 were determined as 169° and 166°, while the effective cone angles for PPh2Fc, P(NMe2)3 and AsPh3 were determined as 171°, 168° and 147°, respectively.  相似文献   

20.
Ruthenium phosphine complexes with a CO ligand [Ru(tpy)(PR3)(CO)Cl]+ (tpy = 2,2′:6′,2″-terpyridine, R = Ph or p-tolyl), were prepared by introduction of CO gas to the corresponding dichloro complexes at room temperature. New carbonyl complexes were characterized by various methods including structural analyses. They were shown to release CO following the addition of several N-donors to form the corresponding substituted complexes. The kinetic data and structural results observed in this study indicated that the CO release reactions proceeded in an interchange mechanism. The molecular structures of [Ru(tpy)(PPh3)(CO)Cl]PF6, [Ru(tpy)(P(p-tolyl)3)(CO)Cl]PF6 and [Ru(tpy)(PPh3)(CH3CN)Cl]PF6 were determined by X-ray crystallography.  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