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.
Upon the addition of Br2 to complexes (P-P)Pt(Ar)2, two different products were observed, depending on the bite angle of the bidentate phosphine ligand: a Pt(II) aryl bromide complex, the product of C-Br reductive elimination, and Pt(IV) oxidative addition complex. At high temperatures, the latter exclusively gave the product of the C-C reductive elimination.  相似文献   

2.
The platinum(IV) complexes fac-L(2)PtMe(3)(OR) (L(2) = bis(diphenylphosphino)ethane, o-bis(diphenylphosphino)benzene, R = carboxyl, aryl; L = PMe(3), R = aryl) undergo reductive elimination reactions to form carbon-oxygen bonds and/or carbon-carbon bonds. The carbon-oxygen reductive elimination reaction produces either methyl esters or methyl aryl ethers (anisoles) and L(2)PtMe(2), while the carbon-carbon reductive elimination reaction affords ethane and L(2)PtMe(OR). Choice of reaction conditions allows the selection of either type of coupling over the other. A detailed mechanistic study of the reductive elimination reactions supports dissociation of the OR(-) ligand as the initial step for the C-O bond formation reaction. This is followed by a nucleophilic attack of OR(-) upon a methyl group bound to the Pt(IV) cation to produce the products MeOR and L(2)PtMe(2). C-C reductive elimination proceeds from L(2)PtMe(3)(OR) by initial L (L = PMe(3)) or OR(-) (L(2) = dppe, dppbz) dissociation, followed by C-C coupling from the resulting five-coordinate intermediate. Our studies demonstrate that both C-C and C-O reductive elimination reactions from Pt(IV) are more facile in polar solvents, in the presence of Lewis acids, and for OR(-) groups that contain electron withdrawing substituents.  相似文献   

3.
4.
Platinum(IV) complexes containing monodentate sulfonamide ligands, fac-(dppbz)PtMe(3)(NHSO(2)R) (dppbz = o-bis(diphenylphosphino)benzene; R = p-C(6)H(4)(CH2)(3)CH(3) (1a), p-C(6)H(4)CH(3) (1b), CH(3) (1c)), have been synthesized and characterized, and their thermal reactivity has been explored. Compounds 1a-c undergo competitive C-N and C-C reductive elimination upon thermolysis to form N-methylsulfonamides and ethane, respectively. Selectivity for either C-N or C-C bond formation can be achieved by altering the reaction conditions. Good yields of the C-N-coupled products were observed when the thermolyses of 1a-c were conducted in benzene-d(6). In contrast, exclusive C-C reductive elimination occurred upon themolysis of 1a,b in nitrobenzene-d(5). When the thermolyses of 1a were performed in the presence of sulfonamide anion NHSO2R- in benzene-d(6), ethane elimination was completely inhibited and C-N reductive elimination products were formed in high yield. Mechanistic studies support a two-step reaction pathway involving initial dissociation of NHSO(2)R(-) from the platinum center, followed by nucleophilic attack of this anion on a methyl group of the resulting five-coordinate platinum(IV) cation to form MeNHSO(2)R and (dppbz)PtMe(2). C-C reductive elimination to form ethane occurs directly from the five-coordinate Pt(IV) cation.  相似文献   

5.
The reductive elimination of carbon-carbon bonds is one of the most fundamentally and synthetically important reaction steps in organometallic chemistry, yet relatively little is understood about the factors that govern the kinetics of this reaction. C-C elimination from complexes with the common d (6) six-coordinate configuration generally proceeds via prior ligand loss, which greatly complicates any attempt to directly measure the rates of the specific elimination step. We report the synthesis of a series of five-coordinate d (6) iridium complexes, ( (tBu)PCP)Ir(R)(R'), where R and R' are Me, Ph, and (phenyl-substituted) vinyl and alkynyl groups. For several of these complexes (R/R' = Ph/Vi, Me/Me, Me/Vi, Me/CCPh, and Vi/CCPh, where Vi = trans-CHCHPh) we have measured the absolute rate of C-C elimination. For R/R' = Ph/Ph, Ph/Me, and Ph/CCPh, we obtain upper limits to the elimination rate; and for R/R' = CCPh/CCPh, a lower limit. In general, the rates decrease (activation barriers increase) according to the following order: acetylide < vinyl approximately Me < Ph. Density functional theory (DFT) calculations offer significant insight into the factors behind this order, in particular the slow rates for elimination of the vinyl and, especially, phenyl complexes. The transition states are calculated to involve rotation of the aryl or vinyl group around the Ir-C bond, prior to C-C elimination, such that the group to which it couples can add to the face of the aryl or vinyl group. This rotation is severely hindered by the presence of the phosphino -t-butyl groups that lie above and below the plane of the aryl/vinyl group in the ground state. Accordingly, calculations predict dramatically different relative rates of elimination from the much less sterically hindered complexes ( (H)PCP)Ir(R)(R'). For example, the barrier to elimination from ( (H)PCP)Ir(Me) 2 is 20 kcal/mol, which is 2 kcal/mol greater than from the ( (tBu)PCP)Ir analogue. In contrast, the activation enthalpies calculated for vinyl-vinyl and phenyl-phenyl elimination from ( (H)PCP)Ir are remarkably low, only 2 and 9 kcal/mol, respectively; these values are 16 and 22 kcal/mol less than those of the corresponding ( (tBu)PCP)Ir complexes. Moreover, since these eliminations are very nearly thermoneutral, the barriers are calculated to be equally low for the reverse reactions [C-C oxidative addition to ( (H)PCP)Ir]. The absence of differences in intraligand CC bond lengths in the transition states relative to the ground states, combined with a comparison of calculated "face-on" and "planar" transition states for C-C coupling, suggests that the critical importance of the aryl/vinyl rotation is based on geometric or steric factors rather than electronic ones. Thus there is no evidence for participation of the pi or pi* orbitals of the aryl or vinyl groups in the formation of the C-C bond, although a small pi effect cannot be rigorously excluded. Likewise, the results do not support the hypothesis that the degree of directionality of the carbon-based orbital used for bonding to iridium (sp (3) > sp (2) > sp) plays an important role in this system in determining the barrier to reductive elimination.  相似文献   

6.
This communication describes the synthesis of a family of unusually stable palladium(IV) complexes containing two chelating 2-phenylpyridine ligands and two benzoates. These complexes undergo clean C-O bond-forming reductive elimination upon heating, and the mechanism of this catalytically relevant process has been studied in detail. Solvent effects, crossover experiments, Eyring plots (which show DeltaS of -1.4 +/- 1.9 and 4.2 +/- 1.4 in CDCl3 and DMSO, respectively), and Hammett analysis (which shows rho = -1.36 +/- 0.04 upon substitution of the para-benzoate substituent) all suggest that reductive elimination does not proceed via initial dissociation of a benzoate ligand. Instead, an unusual mechanism involving pre-equilibrium dissociation of the N-arm of the phenylpyridine ligand is proposed.  相似文献   

7.
Coupled mechanical forces are known to drive a range of covalent chemical reactions, but the effect of mechanical force applied to a spectator ligand on transition metal reactivity is relatively unexplored. Here we quantify the rate of C(sp2)–C(sp2) reductive elimination from platinum(ii) diaryl complexes containing macrocyclic bis(phosphine) ligands as a function of mechanical force applied to these ligands. DFT computations reveal complex dependence of mechanochemical kinetics on the structure of the force-transducing ligand. We validated experimentally the computational finding for the most sensitive of the ligand designs, based on MeOBiphep, by coupling it to a macrocyclic force probe ligand. Consistent with the computations, compressive forces decreased the rate of reductive elimination whereas extension forces increased the rate relative to the strain-free MeOBiphep complex with a 3.4-fold change in rate over a ∼290 pN range of restoring forces. The calculated natural bite angle of the free macrocyclic ligand changes with force, but 31P NMR analysis and calculations strongly suggest no significant force-induced perturbation of ground state geometry within the first coordination sphere of the (P–P)PtAr2 complexes. Rather, the force/rate behavior observed across this range of forces is attributed to the coupling of force to the elongation of the O⋯O distance in the transition state for reductive elimination. The results suggest opportunities to experimentally map geometry changes associated with reactions in transition metal complexes and potential strategies for force-modulated catalysis.

The influence of mechanical force on the rates of model reductive elimination reactions depends on the structure of the force-transducing ligand and provides a measure of geometry changes upon reaching the transition state.  相似文献   

8.
Cationic complexes of the type fac-[(L(2))Pt(IV)Me(3)(pyr-X)][OTf] (pyr-X = 4-substituted pyridines; L(2) = diphosphine, viz., dppe = bis(diphenylphosphino)ethane and dppbz = o-bis(diphenylphosphino)benzene; OTf = trifluoromethanesulfonate) undergo C-C reductive elimination reactions to form [L(2)Pt(II)Me(pyr-X)][OTf] and ethane. Detailed studies indicate that these reactions proceed by a two-step pathway, viz., initial reversible dissociation of the pyridine ligand from the cationic complex to generate a five-coordinate Pt(IV) intermediate, followed by irreversible concerted C-C bond formation. The reaction is inhibited by pyridine. The highly positive values for DeltaS()(obs) = +180 +/- 30 J K(-1) mol(-1), DeltaH(obs) = 160 +/- 10 kJ mol(-1), and DeltaV()(obs) = +16 +/- 1 cm(3) mol(-1) can be accounted for in terms of significant bond cleavage and/or partial reduction from Pt(IV) to Pt(II) in going from the ground to the transition state. These cationic complexes have provided the first opportunity to carry out detailed studies of C-C reductive elimination from cationic Pt(IV) complexes in a variety of solvents. The absence of a significant solvent effect for this reaction provides strong evidence that the C-C reductive coupling occurs from an unsaturated five-coordinate Pt(IV) intermediate rather than from a six-coordinate Pt(IV) solvento species.  相似文献   

9.
Stereoisomeric Pt(IV) complexes with threonine (ThrH = HOCH(CH3)CH(NH2)COOH, ??-amino-??-hydroxybutyric acid) were obtained. In the complexes trans-[Pt(S-ThrH)2Cl4] and trans-[Pt(R-ThrH)(S-ThrH)Cl4], the ThrH molecules act as monodentate ligands coordinated through the NH2 group. In the complexes cis- and trans-[Pt(S-Thr)2Cl2] and trans-[Pt(R-Thr)(S-Thr)Cl2], the deprotonated ligands are coordinated in a bidentate fashion through the NH2 and COO?-groups (R,S is the absolute configuration of the asymmetric carbon atom). All the complexes were identified using elemental analysis, IR spectroscopy, and 195Pt, 13C, and 1H NMR spectroscopy. The complexes trans-[Pt(S-ThrH)2Cl4] · 3H2O and cis-[Pt(S-Thr)2Cl2] · 2H2O were additionally characterized by X-ray diffraction.  相似文献   

10.
The complexes [cis-Rh(SC6F5)(PPh3)2(L)] (L = py, 3-Mepy, isoquin, N-Melm; py = pyridine, 3-Mepy = 3-methylpyridine, isoquin = isoquinoline, N-Melm = N-methylimidazole) readily undergo oxidative addition of HR (R = H, SC6F5, C2Ph) to give [RhH(R)(SC6F5)(PPh3)n(L)3−n] (n = 1, 2) whereas the complexes [cis-Rh(SC6F5)(PPh3)2(L′)] (L′ = 2-Mepy, 2,6-Me2py, quin; 2-Mepy = 2-methylpyridine; 2,6-Me2py = 2,6-dimethylpyridine, quin = quinoline) react only where R = C2Ph. Where conditions favour the formation of [RhH(R)(SC6F5)(PPh3)n(L′)3−n] reductive elimination of H2 (R = H) or C6F5SH (R = SC6F5, C2Ph) occurs.  相似文献   

11.
Electron-transfer series are described for three ferric complexes of the pentadentate ligand 4,8,11-trimethyl-1,4,8,11-tetraazacyclotetradecane-1-acetate (Me(3)cyclam-acetate) with axial chloride, fluoride, and azide ligands. These complexes can all be reduced coulometrically to their Fe(II) analogs and oxidized reversibly to the corresponding Fe(IV) species. The Fe(II), Fe(III), and Fe(IV) species have been studied spectroscopically and their UV-vis, M?ssbauer, EPR, and IR spectra are presented. The fluoro species [(Me(3)cyclam-acetate)FeF](n+) (n = 0, 1, 2) have been studied computationally using density functional theory (DFT), and the electronic structure of the Fe(IV) dication [(Me(3)cyclam-acetate)FeF](2+) is compared with that of the isoelectronic Fe(IV) oxo cation [(Me(3)cyclam-acetate)FeO](+); the different properties of the two species are mainly due to the significantly covalent Fe=O pi bonds in the latter.  相似文献   

12.
13.
Oxidation of a doubly cyclometallated platinum(II) complex results in two isomeric platinum(IV) complexes. Whereas the trans isomer is robust, being manipulable in air at room temperature, the cis isomer decomposes at -20 °C and above. Reductive coupling of an alkyl chloride at the cis isomer gives a new species which can be reoxidised. The independence of this coupling on additional halide rules out the reverse of an S(N)2 reaction, leaving a concerted process as the only sensible reaction pathway.  相似文献   

14.
The reactions of [Ag(NH=CMe2)2]ClO4 with cis-[PtCl2L2] in a 1:1 molar ratio give cis-[PtCl(NH=CMe2)(PPh3)2]ClO4 (1cis) or cis-[PtCl(NH=CMe2)2(dmso)]ClO4 (2), and in 2:1 molar ratio, they produce [Pt(NH=CMe2)2L2](ClO4)2 [L = PPh3 (3), L2= tbbpy = 4,4'-di-tert-butyl-2,2'-dipyridyl (4)]. Complex 2 reacts with PPh3 (1:2) to give trans-[PtCl(NH=CMe2)(PPh3)2]ClO(4) (1trans). The two-step reaction of cis-[PtCl2(dmso)2], [Au(NH=CMe2)(PPh3)]ClO4, and PPh3 (1:1:1) gives [SP-4-3]-[PtCl(NH=CMe2)(dmso)(PPh3)]ClO4 (5). The reactions of complexes 2 and 4 with PhICl2 give the Pt(IV) derivatives [OC-6-13]-[PtCl3(NH=CMe2)(2)(dmso)]ClO4 (6) and [OC-6-13]-[PtCl2(NH=CMe2)2(dtbbpy)](ClO4)2 (7), respectively. Complexes 1cis and 1trans react with NaH and [AuCl(PPh3)] (1:10:1.2) to give cis- and trans-[PtCl{mu-N(AuPPh3)=CMe2}(PPh3)2]ClO4 (8cis and 8trans), respectively. The crystal structures of 4.0.5Et2O.0.5Me2CO and 6 have been determined; both exhibit pseudosymmetry.  相似文献   

15.
The ease of H elimination from the 4- (beta-) position in a series of allylic complexes [Pd(5-C(6)F(5)-1-3-eta(3)-cyclohexenyl)XL](n+) (X, L = Br, N-, P-donor, C(6)Cl(2)F(3); n = -1, 0, +1) was compared by analyzing their decomposition products at 50 degrees C. Pd-H elimination does not occur for cationic complexes, whereas it is the main decomposition pathway for neutral and anionic complexes. In addition to the charge of the complex, the ease of this Pd-H elimination is determined by the trans influence of the ligands (aryl > PMe(3) > Br, N-donor).  相似文献   

16.
The nitrile ligands in trans-[PtX2(PhCN)2] (X = Cl, Br, I) undergo sequential 1,3 dipolar cycloadditions with nitrones R1R2C=N+(Me)-O(-) (R1 = H, R2 = Ph; R1 = CO2Et, R2 = CH2CO2Et) to selectively form the Delta4-1,2,4-oxadiazoline complexes trans-[PtX2(PhCN) (N=C(Ph)-O-N(Me)-CR1R2)] or trans-[PtX2(N=C(Ph)-O-N(Me)-CR1R2)2] in high yields. The reactivity of the mixed ligand complexes trans-[PtX2(PhCN)(N=C(Ph)-O-N(Me)-CR1R2)] towards oxidation and ligand substitution was studied in more detail. Oxidation with Cl2 or Br2 provides the Pt(IV) species trans-[PtX2Y2(PhCN)(N=C(Ph)-O-N(Me)-CH(Ph))] (X, Y = Cl, Br). The mixed halide complex (X = Cl, Y = Br) undergoes halide scrambling in solution to form trans-[PtX(4-n)Yn(PhCN)(N=C(Ph)-O-N(Me)-CH(Ph))] as a statistical mixture. Ligand substitution in trans-[PtCl2(PhCN)(N=C(Ph)-O-N(Me)-CR1R2)] allows for selective replacement of the coordinated nitrile by nitrogen heterocycles such as pyridine, DMAP or 1-benzyl-2-methylimidazole to produce mixed ligand Pt(II) complexes of the type trans- [PtX2(heterocycle)(N=C(Ph)-O-N(Me)-CR1R2)]. All compounds were characterised by elemental analysis, mass spectrometry, IR and 1H, 13C and 195Pt NMR spectroscopy. Single-crystal X-ray structural analysis of (R,S)-trans-[PtBr2(N=C(Ph)-O-N(Me)-CH(Ph))2] and trans-[PtCl2(C5H5N)(N=C(Ph)-O-N(Me)-CH(Ph))] confirms the molecular structure and the trans configuration of the heterocycles relative to each other.  相似文献   

17.
Carbon-carbon bond-forming reductive elimination of biphenyl is observed upon two-electron oxidation of the [ZrIVPh2(ap)2]2- dianion. Crossover experiments confirm that the C-C bond-forming step occurs at a single zirconium metal center. The reactivity is enabled by the participation of a redox-active amidophenolate ligand set.  相似文献   

18.
19.
Difluoro Pt(IV) complexes (P-P)Pt(Ar)2F2 were prepared and characterized. One of the fluoro ligands can be selectively removed from the Pt coordination sphere giving the cationic product with the phosphine ligand trans to the empty coordination site.  相似文献   

20.
The effect of different N-N spectator ligands on the reactivity of platinum(II) complexes was investigated by studying the water lability of [Pt(diaminocyclohexane)(H2O)2]2+ (Pt(dach)), [Pt(ethylenediamine)(H2O)2]2+ (Pt(en)), [Pt(aminomethylpyridine)(H2O)2]2+ (Pt(amp)), and [Pt(N,N'-bipyridine)(H2O)2]2+ (Pt(bpy)). Some of the selected N-N chelates form part of the coordination sphere of Pt(II) drugs in clinical use, as in Pt(dach) (oxaliplatin), or are models, regarding the nature of the amines, with higher stability in terms of substitution and hydrolysis of the diamine moiety, as in Pt(en) (cisplatin) and Pt(amp) (AMD473). The effect of pi-acceptors on the reactivity was investigated by introducing one (Pt(amp)) or two pyridine rings (Pt(bpy)) in the system. The pK(a) values for the two water molecules (viz., Pt(dach) (pK(a1) = 6.01, pK(a2) = 7.69), Pt(en) (pK(a1) = 5.97, pK(a2) = 7.47), Pt(amp) (pK(a1) = 5.82, pK(a2) = 6.83), Pt(bpy) (pK(a1) = 4.80, pK(a2) = 6.32) show a decrease in the order Pt(dach) > Pt(en) > Pt(amp) > Pt(bpy). The substitution of both coordinated water molecules by a series of nucleophiles (viz., thiourea (tu), L-methionine (L-Met), and guanosine-5'-monophosphate (5'GMP-) was investigated under pseudo-first-order conditions as a function of concentration, temperature, and pressure using UV-vis spectrophotometric and stopped-flow techniques and was found to occur in two subsequent reaction steps. The following k1 values for Pt(dach), Pt(en), Pt(amp), and Pt(bpy) were found: tu (25 degrees C, M(-1) s(-1)) 21 +/- 1, 34.0 +/- 0.4, 233 +/- 5, 5081 +/- 275; L-Met (25 degrees C) 0.85 +/- 0.01, 0.70 +/- 0.03, 2.15 +/- 0.05, 21.8 +/- 0.6; 5'GMP- (40 degrees C) 5.8 +/- 0.2, 3.9 +/- 0.1, 12.5 +/- 0.5, 24.4 +/- 0.3. The results for k2 for Pt(dach), Pt(en), Pt(amp), and Pt(bpy) are as follows: tu (25 degrees C, M(-1) s(-1)) 11.5 +/- 0.5, 10.2 +/- 0.2, 38 +/- 1, 1119 +/- 22; L-Met (25 degrees C, s(-1)) 2.5 +/- 0.1, 2.0 +/- 0.2, 1.2 +/- 0.3, 290 +/- 4; 5'GMP- (40 degrees C, M(-1) s(-1)) 0.21 +/- 0.02, 0.38 +/- 0.02, 0.97 +/- 0.02, 24 +/- 1. The activation parameters for all reactions suggest an associative substitution mechanism. The pK(a) values and substitution rates of the complexes studied can be tuned through the nature of the N-N chelate, which is important in the development of new active compounds for cancer therapy.  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