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 46 毫秒
1.
Photolysis of M2(CO)4(μ-S-t-Bu)2, where M = Rh or Ir, in Nujol matrices at ca. 90 K results in simple CO loss to form a tricarbonyl intermediate analogous to that observed for Rh2(CO)4(μ-Cl)2. Photolysis of the anions, [M(CO)2Cl2]1−, where M = Rh or Ir, in inert ionic matrices at ca. 90 K, results in CO-loss to form an intermediate analogous to that formed by Rh(CO)2(i-Pr2HN)Cl. Finally, photolysis of trans-Ir(CO)(PMe3)2Cl in a Nujol matrix at ca. 90 K gives rise to a new species whose carbonyl band is shifted slightly down in energy as has been observed for trans-Rh(CO)(PMe3)2Cl. In all cases the iridium compounds behave similarly to the rhodium species although the photon energy for iridium photochemistry is typically above that of the rhodium compounds.  相似文献   

2.
The reactions of trans-[(PPh3)2M(CO)Cl] (M = Rh and Ir) with benzildiimine (H2BDI = 2) derived from benzil-bis(trimethylsilyl)diimine (Si2BDI) (1) in a 1:2 and 1:1 molar ratio afforded the cationic bis-benzildiiminato complexes [Rh(PPh3)2(HBDI)2]Cl (3) and the mono-benzildiimine complex [Ir(PPh3)2(CO)(H2BDI)]Cl (4), respectively. Both complexes are fully characterized using IR, FAB-MS, NMR spectroscopy and elemental analysis. The single crystal X-ray structure analysis reveals a distorted octahedral coordination geometry for the Rh(III) in 3 and a highly distorted square pyramidal geometry for Ir(I) in 4. In addition, the solid-state structure of Si2BDI is reported here for the first time showing the substituents highly twisted because of steric reasons.  相似文献   

3.
《Inorganica chimica acta》2005,358(2):303-309
The reactions of two equivalents of the ligands POT or POZ with one equivalent of the rhodium complex [Rh(μ-Cl)(CO)2]2 afford the complexes [(POT)Rh(CO)Cl] (1) and [(POZ)Rh(CO)Cl] (2), respectively. The crystal structures of both complexes have been determined showing the rhodium centers to be into slightly distorted square planar environments. Preliminary screening of the catalytic systems POT/Rh and POZ/Rh in the asymmetric hydroformylation of styrene has been carried out.  相似文献   

4.
The crystal structures of the four-coordinate trans-[Rh(Cl)(CO)(SbPh3)2] (1) and the five-coordinate trans-[Rh(Cl)(CO)(SbPh3)3] (2) are reported, as well as the unexpected oxidative addition product, trans-[Rh(I)2(CH3)(CO)(SbPh3)2] (3), obtained from the reaction of 2 with CH3I. The formation constants of the five-coordinate complex were determined in dichloromethane, benzene, diethyl ether, acetone and ethyl acetate as 163±8, 363±10, 744±34, 1043±95 and 1261±96 M−1, respectively. While coordinating solvents facilitate the formation of the five-coordinate complex, the four-coordinate complex could be obtained from diethyl ether due to the favorable low crystallization energy. The tendency of stibine ligands to form five-coordinate rhodium(I) complexes is attributed mainly to electron deficient metal centers in these systems, with smaller contributions by the steric effects. The average effective cone angle for the SbPh3 ligand in the three crystallographic studies was determined as 139° with individual values ranging from 133 to 145°.  相似文献   

5.
Coordination compounds of chelating 8-methylthioquinoline (MTQ) with the complex fragments ReI(CO)3Cl, [RuII(bpy)2]2+, [RhIII(C5Me5)Cl]+, [IrIII(C5Me5)Cl]+, and PtIVMe4 were synthesized and structurally characterized. Whereas the ruthenium(II) complex displays the strongest preference of bonding to N versus S, the compound (MTQ)PtMe4 shows the most balanced metal-donor bonding within the chelate ring due to a relatively short bond to S (2.319 Å) versus N (2.150 Å). The complex fac-(MTQ)Re(CO)3Cl exhibits a particularly long metal-sulfur bond at 2.472 Å. Cyclic voltammetry of [(MTQ)Ru(bpy)2](PF6)2 reveals one reversible oxidation to RuIII and three closely spaced reduction waves for the coordinated ligands. In comparison with the imine/thioether chelate ligand 1-methyl-2-(methylthiomethyl)-1H-benzimidazole (mmb) the MTQ ligand with its more rigid chelate setting N(sp2)-C(sp2)-C(sp2)-S forms generally shorter M-S bonds and displays stronger π acceptor behaviour.  相似文献   

6.
[Ir(CO)Cl(PCy3)2] was obtained by the slow attack of 1,2-dichloromethane on (Bu4N)[Ir2(μ-Dcbp)(CO)2(PCy3)2] (Dcbp = 3,5-dicarboxylatepyrazole). The Ir-CO, Ir-Cl and Ir-P bond distances are 1.778(10), 2.374(3) and 2.3486(8) Å, respectively. The Ir-P bond distances for a number of different Vaska complexes indicate the shortest bond distances for phoshines containing electron withdrawing groups. An excellent correlation between DFT (OLYP/ZORA/TZP) and experimental structures is obtained as reflected by the RMSD values (H excluded) of between 0.083 and 0.268 Å for the different complexes studied. The calculated Ir-P bond distances and ν(CO) stretching frequencies closely follow the trends obtained from the experimental results.  相似文献   

7.
Treatment of [Rh(β-diketonato)(cod)] with CO resulted in better yields of [Rh(FcCOCHCOR)(CO)2] than by treating [Rh(Cl)(CO)2]2 with FcCOCH2COR, R = CF3 (Hfctfa), CH3 (Hfca), Ph (Hbfcm, Ph = phenyl) and Fc (Hdfcm, Fc = ferrocenyl). The single crystal structure of the fctfa rhodium(I) complex [C16H10F3FeO4Rh], monoclinic, C 2/c(15), a = 13.266(3) Å, b = 19.553(3) Å, c = 13.278(3) Å, β = 100.92(2)°, Z = 8 showed both rotational and translational displacement disorders for the CF3 group. An electrochemical study revealed that the formal reduction potential, E0′, for the electrochemically reversible one electron oxidation of the ferrocenyl group varied between 0.304 (for the fctfa complex) and 0.172 V (for the dfcm complex) versus Fc/Fc+ in a manner that could be directly traced to the group electronegativities, χR, of the R groups on the β-diketonato ligands, as well as to the values of the free β-diketones. Anodic peak potentials, Epa,Rh, for the dominant cyclic voltammetry peak associated with rhodium(I) oxidation were between 0.718 (bfcm complex) and 1.022 V (dfcm complex) versus Fc/Fc+. Coulometric experiments implicated a second, much less pronounced anodic wave for the apparent two-electron RhI oxidation that overlaps with the ferrocenyl anodic wave and that the redox processes associated with these two RhI oxidation waves are in slow equilibrium with each other.  相似文献   

8.
《Inorganica chimica acta》2004,357(10):2818-2826
[{Rh(cod)Cl}2] (cod=1,5-cyclooctadiene) reacts with o-(diphenylphosphino)benzaldehyde (PPh2(o-C6H4CHO)) (Rh:P=1:1) in the presence of aromatic diamines or 8-aminoquinoline (NN) to give acylhydride [Rh(Cl)(H){PPh2(o-C6H4CO)}(NN)] species. The oxidative addition of PPh2(o-C6H4CHO) in the presence of (NN) and PPh3 gives cationic species [Rh(H){PPh2(o-C6H4CO)} (PPh3)(NN)]+ containing mutually trans phosphorus atoms. When (NN)=8-aminoquinoline, a mixture of two isomers is obtained. These isomers differ in the nitrogen cis to the hydride, amino or quinolinic. By using Rh:PPh2(o-C6H4CHO)=1:2 stoichiometric ratios, oxidative addition of one PPh2(o-C6H4CHO) and P-coordination of another PPh2(o-C6H4CHO) occurs. The aldehyde group undergoes then a condensation reaction with the coordinated amine to afford new PNN terdentate ligands, phosphine-amino-imine when (NN)=diamine or phosphine-diimine when (NN)=8-aminoquinoline. These reactions give selectively the corresponding complexes [Rh(H){PPh2(o-C6H4CO)}(PNN)]+ containing trans phosphorus atoms and the hydride cis to the new imino group. X-ray diffraction studies of the PNN complexes are reported.  相似文献   

9.
Reaction of [Rh(CO)2I]2 (1) with MeI in nitrile solvents gives the neutral acetyl complexes, [Rh(CO)(NCR)(COMe)I2]2 (R=Me, 3a; tBu, 3b; vinyl, 3c; allyl, 3d). Dimeric, iodide-bridged structures have been confirmed by X-ray crystallography for 3a and 3b. The complexes are centrosymmetric with approximate octahedral geometry about each Rh centre. The iodide bridges are asymmetric, with Rh-(μ-I) trans to acetyl longer than Rh-(μ-I) trans to terminal iodide. In coordinating solvents, 3a forms mononuclear complexes, [Rh(CO)(sol)2(COMe)I2] (sol=MeCN, MeOH). Complex 3a reacts with pyridine to give [Rh(CO)(py)(COMe)I2]2 and [Rh(CO)(py)2(COMe)I2] and with chelating diphosphines to give [Rh(Ph2P(CH2)nPPh2)(COMe)I2] (n=2, 3, 4). Addition of MeI to [Ir(CO)2(NCMe)I] is two orders of magnitude slower than to [Ir(CO)2I2]. A mechanism for the reaction of 1 with MeI in MeCN is proposed, involving initial bridge cleavage by solvent to give [Rh(CO)2(NCMe)I] and participation of the anion [Rh(CO)2I2] as a reactive intermediate. The possible role of neutral Rh(III) species in the mechanism of Rh-catalysed methanol carbonylation is discussed.  相似文献   

10.
The rhodium(I) complexes TpmsRh(CO)2 (1) and TpmsRh(cod) (2) of the tripodal nitrogen ligand tris(pyrazolyl)methanesulfonate, Tpms=[(pz)3CSO3], catalyze the hydroformylation of 1-hexene. Addition of phosphine has a negative effect on the activity. The hydroformylation activity reaches a maximum at about 60 °C. At temperatures above 80 °C hydrogenation becomes an important secondary reaction. When the catalysis is performed at 60 °C in acetone with 1 or 2 as catalyst precursor all of the rhodium is recovered in the form of the rhodium(III) bis(acyl) complex TpmsRh(CO)(COC6H13)2 (9). A similar behaviour is observed with rhodium(I) complexes bearing the tripodal oxygen ligand LOMe=[(cyclopentadienyl)tris(dimethylphosphito-P) cobalt O,O,O″]. In this case all of the rhodium is transformed into LOMeRh(CO)(COC6H13)2 (10). These hitherto unknown bis(acyl) rhodium(III) complexes show the same catalytic activity as the rhodium(I) starting compounds.  相似文献   

11.
The synthesis of new β-diketonato rhodium(I) complexes of the type [Rh(FcCOCHCOR)(CO)2] and [Rh(FcCOCHCOR)(CO)(PPh3)] with Fc=ferrocenyl and R=Fc, C6H5, CH3 and CF3 are described. 1H, 13C and 31P NMR data showed that for each of the non-symmetric β-diketonato mono-carbonyl rhodium(I) complexes, two isomers exist in solution. The equilibrium constant, Kc, which relates these two isomers in an equilibrium reaction, are concentration independent but temperature and solvent dependent. ΔrG, ΔrH and ΔrS values for this equilibrium have been determined and a linear relationship between solvent polarity on the Dimroth scale and Kc exists. The relationship between RhP bond lengths, d(RhP), and 31P NMR peak positions as well as coupling constants 1J(31P103Rh) has been quantified to allow calculation of approximate d(RhP) values. Variations in d(RhP) for [Rh(RCOCHCOR′)(CO)(PPh3)] complexes have also been related to the group electronegativities (Gordy scale) of the terminal β-diketonato R groups trans to PPh3. A measure of the electron density on the rhodium centre of [Rh(RCOCHCOR′)(CO)(PPh3)] may be expressed in terms of the IR carbonyl stretching wave number, ν(CO), the sum of the group electronegativities of the R and R′ groups, (χR+χR′), or the observed pKa values of the free β-diketones RCOCH2COR. An empirical relationship between ν(CO) and either pKa or (χR+χR′) has also been quantified.  相似文献   

12.
Three kinds of crystalline compounds containing the nitrosylpentaamminechromium complexes [Cr(NO)(NH3)5]2+(A) were obtained: chloride ACl2 (red-orange), chloride perchlorate ACl(ClO4) (brown), and perchlorate A(ClO4)2 (green). The cause of the color change of the complex A with the change of outer sphere anions was sought using X-ray structural data of ACl2, ACl(ClO4), and A(ClO4)2. Crystal data: ACl2, orthorhombic, space group Cmcm, a=10.0236 (9) Å, b=9.098 (3) Å, c=10.357(1) Å, V=944.5 (5) Å3, Z=4; ACl(ClO4), tetragonal, space group P4/nmm, a=7.6986 (8) Å, c=9.9566(8) Å, V=590.1 (1) Å3,Z=2; A(ClO4)2, orthorhombic, space group Pnma, a=15.760 (2) Å, b=11.480(2) Å, c=7.920 (2) Å, V=1432.9 (4) Å3, Z=4. The complex cation in ACl2 has a distorted octahedral structure with a linear CrNO moiety. The short CrN (nitrosyl) distance of 1.692 (7) Å indicates the presence of multiple bonding between the chromium atom and the nitrogen atom in the nitrosyl group. The interatomic distances and angles within the complex cations hardly change with the change of the counter anions, while the distances between the complex cations in each crystal increase in the order ACl2<ACl(ClO4)<A(ClO4)2. The bulky perchlorate anions seems to separate the complex cations, while smaller chloride anions are not large enough to separate them. The distance (3.213(5) Å) between O(NO) and N(NH3 in the adjacent complex cation) is rather short in the crystal of ACl2, and there are six hydrogen bonds, where the NO group is surrounded by four NH3 ligands. The distance (4.002(5) Å) between O(NO) and N(NH3) is much longer in the crystal of A(ClO4)2, indicating the presence of no hydrogen bonding. In the crystal of ACl(ClO4) the distance (3.452(4) Å) between O(NO) and N(NH3) is in between those of ACl2 and A(ClO4)2. The presence of hydrogen bonding between O(NO) and N(NH3 in the adjacent complex cation) seems to cause the color change with the change of outer sphere anions.  相似文献   

13.
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   

14.
《Inorganica chimica acta》1988,144(1):109-121
The influence of the π acidity of the PR 3 ligands in Mo 2Cl 4(PR 3) 4 compounds on the strength of the Mo−Mo quadruple bond has been studied for twenty such compounds. The main result is that reasonable correlations exist between the δ → δ* transition energy (TE) and two accepted measures of π-acidity for phosphine ligands, namely Tolman's v̄(CO) IR criterion and Bodner's Δδ(CO) 13C NMR criterion. For a plot of TE versusv̄(CO) a correlation coefficient of −0.892 was obtained employing data for 15 compounds. For TE versus Δδ(CO), a correlation coefficient of 0.982 was obtained for 13 compounds (excluding three for which a steric factor may be important). An attempt to correlate the Mo−Mo bond lengths with π-acidity was unsuccessful since few compounds were obtained in satisfactory form for structure determination, and for the four compounds with known structure there was no correlation. In addition to the PMe 3 compound whose structure was previously known, those of the PEt 3, PMe 2Ph and PHPh 2 compounds have been determined and are reported here.  相似文献   

15.
Reaction between a mixture of cis-trans-[PtCl2(SMe2)2] and 1 equiv. AsPh3 in chloroform gives cis-[PtCl2(SMe2)(AsPh3)] crystallizing in P21/n with a=10.397(2), b=14.876(3), c=13.956(3) Å, β=90.86(3)° and Z=4. Selected geometrical parameters are PtAs 2.3531(10), PtS 2.262(2), PtCl (trans to S) 2.301(2), PtCl (trans to As) 2.328(2) Å and SPtAs 88.85(6), SPtCl(2) 90.77(8), AsPtCl(1) 91.07(6) and ClPtCl 89.42(7)°. cis-[PtCl2(AsPh3)2]·CHCl3 crystallizes in P21/c with a=20.557(4), b=9.5951(19), c=20.147(4) Å, β=96.77(3)° and Z=4. Selected geometrical parameters are PtAs(1) 2.3599(9), PtAs(2) 2.3770(9), PtCl(1) (trans to As(1)) 2.3515(18), PtCl(2) (trans to As(2)) 2.3251(18) Å and AsPtAs 97.87(3), As(1)PtCl(2) 88.67(5), As(2)PtCl(1) 84.30(5) and ClPtCl 89.32(7)°. By comparison with related structures from the literature the following trans influence series was established PMe2Ph>PPh3>AsPh3≈SbPh3>Me2SO≈SMe2≈SPh2>NH3≈olefin>Cl>MeCN.  相似文献   

16.
Two series of A-frame complexes, [Pd2(dppm)2(R)2(μ-X)]+ (R = Me and X = Cl, Br, I, H; R = Mes and X = Br, I), were investigated by cyclic voltammetry (CV). The 2-electron reduction potentials for the first series increase from I (−1.10), Br (−1.17), Cl (−1.25) to H (−1.65 V versus SCE, in CHCl3), as well as in the second series; Br (−1.35) and I (−1.38 V versus SCE, in THF). The nature of the LUMO where the electron reduction takes place is qualitatively addressed by DFT on the corresponding model complexes [Pd2(H2PCH2PH2)2(R)2(μ-X)]+. The LUMO and (LUMO + 1) of the halide derivatives exhibit the presence of Pd dx2-y2 atomic orbitals interacting in an anti-bonding fashion with the n-donor orbitals of X, P, and Me, explaining in part the observed reactivity upon reduction. The X-ray structure of [Pd2(dppm)2(Me)2(μ-Br)]+ compound exhibits the typical A-frame structure with a Pd?Pd non-bonding distance of 3.036(1) Å, and long Pd-Br bonds of 2.5623(5) and 2.5793(5) Å.  相似文献   

17.
Jin Zhao 《Inorganica chimica acta》2005,358(14):4201-4207
A series of compounds of the type CpMoR(CO)3 and RCpMo(CO)3CH3 (R = (CH2)3Si(OMe)3 or CH2Si(OEt)3), containing a (CH2)nSi(OR)3 functionality as a side chain, either fixed to the metal itself or to the cyclopentadienyl ligand are prepared and spectroscopically characterized. These molecules are applicable as precursors for both homogeneous and heterogeneous phase catalysts. Their homogeneous catalytic activity in the olefin epoxidation is in the same order of magnitude as that of their methyl and chloro analogues of the types CpMo(CO)3R and CpMo(CO)3Cl.  相似文献   

18.
The kinetics of rapid CO substitution by PPh3 in Co4(CO)12 and Rh4(CO)12 have been examined by stopped-flow and low temperature FT-IR methods. In Co4(CO)12 rapid (kobs ∼ 1.8 s−1) substitution of CO occurs after a 1–15 s induction period at 28 °C in C6H5Cl solvent by a catalytic process. Addition of PPh3 to Rh4(CO)12 yields Rh4(CO)11(PPh3) according to a predominantly second order rate law k1[Rh4- (CO)12] + k2[Rh4(CO)12][PPh3] with k1 = 25 ± 11 s−1 and k2 = 2.97 ± 0.27 X 104 M−1 s−1 at 28 °C. Substitution of a second CO ligand also occurs rapidly with k1 = 0.15 ± 0.09 s−1 and k2 = 6.54 ± 0.07 X 102 M−1 s−1 at 28 °C. The reactivity of Rh4(CO)12 toward associative substitution is 104– 1011 faster than for the Co and Ir analogues, In Rh4(CO)11(PPh3) the increase in CO substitution rates over Co and Rh analogues is 102–107. The ordering of associative substitution rates Co << Rh >>> Ir in these clusters exaggerates the trend seen in mononuclear metal complexes.  相似文献   

19.
The action of [Co(X)(NO)2]2 (X = Cl, Br, L) on [V(H)(CO)6?nLn] (L = 1/ndi- and tritertiary phosphine; n = 2, 3) in thf yields [V(CO)5?n(NO)L2] and [V(NO)2(thf)4]X as the two main products. Thf is easilty replaced by other ligands L′, leading to the complexes cis-[V(NO)2(thf)4?nL′n]X, where n = 1 to 4. In the case of L′= CNR (R = Cy, iPr, tBu), the species [VX(NO)2L′3] are formed. The presence of X in the first coordination sphere is established by the normal halogen dependence (Cl < Br < I) of 51V shielding.δ(51V) values have been obtained for the two series of complexes and compared with δ of other nitrosylvanadium species, including [VX(NO)L′4]X. for [V(NO)2L′4]br, 51V shielding increases in the sequence {O} < {S} < NR3 < NCMe < AsEt3 < SbEt3 < PEt2Ph < P(OMe)3 < CNR, reflecting a general increase of shielding as the polarizability of the ligand function increases and its electronegativity decreases. Superimposed effects arising from electronic influences (PEtPh2) < PMe3 < P(OMe)3 and steric conditions (chelate-4 ring < 7 ring < 6 ring < 5 ring) are also discussed. Steric factors are especially pronounced in the [V(CO)3(NO)Ph2P(CH2)mPPh3] series (m = 1–4). The thermo-labile parent compound, [V(CO)5NO], has been characterized by its δ(51V) = ?1489 ppm at 245 K.  相似文献   

20.
New chiral P,N-bidentate 1,3,2-diazaphospholidine ligands were obtained by one-step phosphorylation of (2S,3S)-2-(ferrocenylideneamino)-3-methylpentan-1-ol and (4S,5S)-(−)-2-methyl-4-hydroxymethyl-5-phenyl-2-oxazoline. Complexation of the new ligands with [Rh(CO)2Cl]2 and [Pd(allyl)Cl]2 was found to give chelate complexes [Rh(CO)Cl(η2-P,N)] and , correspondingly. With the new P,N-ligands, up to 70% ee was achieved in the asymmetric Pd-catalysed sulfonylation of 1,3-diphenyl-2-propenyl acetate with sodium p-toluenesulfinate. In the enantioselective alkylation of 1,3-diphenyl-2-propenyl acetate with dimethyl malonate, up to 91% enantioselectivity was achieved by using complexes as chiral catalysts.  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