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 138 毫秒
1.
Abstract

Metal(II) perchlorate complexes with the ligands tri(2-pyridyl)phosphine, tri(2-pyridyl)phosphine oxide, and tri(2-pyridyl)arsine have the composition [M(TPX)2] (ClO4)2. Coordination occurs only through the nitrogens of the pyridines. In the case of Cu(II) and tri(2-pyridyl)phosphine oxide, two isomers were obtained. One isomer contains symmetrical tridentate tri(2-pyridyl)phosphine oxide ligands while the second isomer contains an unsymmetrical ligand. The unsymmetrical tri(2-pyridyl)phosphine oxide may be a bidentate ligand or a bridging tridentate. Weak axial interaction between a pyridyl group and a second Cu(II) ion is postulated in solution and may be present in the solid state.  相似文献   

2.
The 1H, 31P and 13C NMR spectra of cis-dialkyl(acetylacetonato)bis(tertiary phosphine)cobalt(III) complexes were obtained in several solvents. These complexes have an octahedral configuration with trans tertiary phosphine ligands. The coordinated tertiary phosphine ligands are partly dissociated in solution. One of the phosphine ligands in CoR2(acac)(PR3′)2 can be readily displaced with pyridine bases to give pyridine-coordinated complexes. From observation of the 1H and 31P NMR spectra several kinetic and thermodynamic data for exchange reactions and displacement reactions of tertiary phosphines were obtained.  相似文献   

3.
Rhodium-catalyzed hydroformylation of 1-octene in the presence of different phosphine and phosphine oxide ligands has been investigated. The molecular structure of new phosphine ligand, fluorenylidine methyl phenyl diphenylphosphine, was determined by single-crystal X-ray crystallography. Parameters such as different ligands, molar ratio of ligand to rhodium complex, ratio of olefin to rhodium complex, pressure of CO : H2 mixture, and time of the reaction were studied. The linear aldehyde was the main product when the phosphine ligands were used as auxiliary ligands while the selectivity was changed to the branched products when the related phosphine oxide ligands were used. Under optimized reaction conditions, in the presence of [Rh(acac)(CO)(Ph3P)]-di(1-naphthyl)phenyl phosphine oxide, conversion of 1-octene reached 97% with 87% selectivity of branched aldehyde.  相似文献   

4.
A series of aurocyanide and auricyanide complexes of phosphines, phosphine sulfides, and phosphine selenides were synthesized. These new complexes have the general formula [L n Au(CN) m ], where L could be Cy3P, (2-CN-Et)3P, Me3PS, Et3PS, Ph3PS, Me3PSe, or Ph3PSe. Auricyanide was reacted with L at 1?:?2 ratio. Products were characterized using elemental analysis, melting point, UV, IR, far-IR solution, and solid-state NMR spectroscopy. Phosphine ligands cause gold(III) reduction to gold(I); less redox tendency was found for phosphine sulfides and phosphine selenides. Tri-coordinate complexes [L2AuCN] were produced from phosphine ligands with gold-tetracyanide. IR and UV spectroscopic methods were used to identify gold oxidation state in the synthesized complexes.  相似文献   

5.
The catalytic performances of Co‐Rh/Fe3O4 catalysts modified with phosphine ligands (PPh3) and its analogues on dicyclopentadiene hydroformylation were evaluated. Among these catalysts, Co‐Rh/Fe3O4 modified with tris(p‐trifluoromethylphenyl)phosphine was determined to be effective for monoformyltricyclodecanes production, whereas Co‐Rh/Fe3O4 modified with PPh3 or tri‐p‐tolylphosphine was effective for the diformyltricyclodecanes production. To investigate the ligand effects, the complex catalyst system (Co‐Rh/Fe3O4 and phosphine ligand) was subjected to pretreatment with syngas and then characterized by thermogravimetry and differential thermal analysis (TG‐DTA). It was determined that the threshold decomposition temperature reflected the corresponding Rh‐phosphine interaction strength, affecting the catalytic selectivity toward different products. A weak Rh‐phosphine interaction was desirable to produce monoformyltricyclodecanes with fast reaction kinetics, whereas a strong Rh‐phosphine complex was required for the synthesis of diformyltricyclodecanes. In addition to the selectivity rule shown in the PPh3 series, experiments with other ligands also demonstrated similar selectivity trends.  相似文献   

6.
Depending on steric effects of phosphine and o-semiquinone (SQ) ligands bonded to the copper atom, (R3P)2CuSQ complexes have either pseudotetrahedral geometry with equivalent phosphine ligands or trigonal-pyramidal with one equatorial and one axial phosphine ligand. In an ESR study of the exchange dynamics of the phosphine ligands in trigonal-pyramidal complexes, the kinetic parameters of the exchange were determined. A model was proposed which explains distortions in the geometry of the complexes by nonvalence interatomic interactions on the periphery of the phosphine and SQ ligands.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 9, pp. 1986–1991, September, 1991.  相似文献   

7.
A set of phosphine complexes of the type W(CO)3(PX3)2(CH2CH2) (X=H, CH3, F, Cl, Br, and I) were investigated by density functional theory method (BP86) to examine the effect of the substituent X on the orientation of C-C vector of the ethylene ligand with respect to one of the metal-ligand bonds as well as the donation and the backdonation in the bonding ligands of phosphine and ethylene. When X=CH3, H, F, and Cl, the ethylene C-C vector prefers to be coplanar with metal-phosphine bonds, while for the ethylene complexes containing PBr3 and PI3 ligands, the structural preference is coplanarity of the ethylene and the metal-carbonyl bonds. The molecular orbital calculations and natural bond orbital analysis were used to examine the structural consequences derived from these complexes. It can be concluded that the structural preferences in the complexes have a clear relation to electronic effects of phosphine ligands. Our calculations for halide phosphine complexes, particularly for PBr3 and PI3, allow us to conclude that in addition to electronic effects, steric factors can also affect the orientation of the ethylene ligand in complexes.  相似文献   

8.
Reactions of triarylphosphines with fluoroantimony(III) triflates give phosphine antimony(III) complexes, which undergo spontaneous reductive elimination of fluorophosphonium cations. The resulting phosphine antimony(I) complexes catenate to give the first examples of cationic antimony bicyclic compounds, [(R3P)4Sb6]4+, featuring a bicyclo[3.1.0]hexastibine framework stabilized by four phosphine ligands. The unprecedented 14‐electron redox process illustrates the generality of the reductive catenation method.  相似文献   

9.
Substitution reactions of phosphine ligands, triphenylphosphine (PPh3), tri(m-chlorophenyl)phosphine (m-ClPPh3), tri(p-methoxyphenyl)phosphine (p-MeOPPh3) and tri(benzyl)phosphine (PBz3) with [M(CO)4(PCA)] (M?=?Cr, Mo and W, PCA?=?pyrazinecarboxamide) were found to be dependent on the type of metal and phosphine ligand. The complexes were characterized by elemental analysis, mass spectrometry, and IR and 1H NMR spectroscopy. UV–vis spectra of the complexes in different solvents showed bands due to metal-to-ligand charge transfer.  相似文献   

10.
The chiral phosphine‐oxazoline ligands 3 and 4 bearing 4‐alkoxymethyl substituents on the oxazoline ring with (R)‐configuration were prepared from L‐serine methyl ester in 66% and 33% yields, respectively. Along this synthetic pathway, the β‐hydroxylamides derived from L‐serine methyl ester and 2‐halobenzoyl chlorides were expediently converted to the corresponding oxazolines by using diethylaminosulfur trifluoride as the activation agent. Potassium diphenylphosphide was the reagent of choice for replacing the bromine atom on the phenyl ring, giving the desired oxazoline‐phosphine ligands 3 and 4 . Together with [Pd(η3‐allyl)Cl]2, ligands 3 and 4 induced an enantioselective allylic substitution reaction of 1,3‐diphenyl‐2‐pro‐penyl acetate by dimethyl malonate. Although ligands 3 and 4 exhibit the (R)‐configuration, differing from the (S)‐configuration of Pfaltz‐Helmchen‐Williams phosphine‐oxazoline ligands, all these ligands led to the same enantiotopic preference in the allylic substitution reaction. To facilitate the recovery and reuse of the phosphine‐oxazoline ligand, immobilization on Merrifield resin was attempted, albeit in low loading.  相似文献   

11.
Only [RuCl2(p‐cymene)(PR3)] complexes where the phosphine ligand, PR3, is both strongly basic and bulky proved to be effective catalysts for the controlled atom transfer radical polymerisation (ATRP) of methyl methacrylate and styrene. The best phosphine ligands were typically P(i‐Pr)3, P(cyclohexyl)2Ph, P(cyclohexyl)3, and P(cyclopentyl)3. Less basic and/or bulky phosphines led to ineffective systems for ATRP. Tricyclohexylarsine gave rise to a highly efficient catalyst system. However, related complexes in which the phosphine ligand was replaced by tricyclohexylstibine, nitrogen (piperidine and 4‐cyanopyridine) and carbon ligands (alkyl isocyanides) proved to be inefficient. The observation of a direct relationship between the p‐cymene lability (measured by TGA) and catalyst activity suggests that p‐cymene release is a prerequisite for the polymerisation process.  相似文献   

12.
Photolysis of [Me2SiSiMe2)[C5H4Fe(CO2)]2with a series of bis(phosphine)ligands Ph2P(CH2)n PPh2(n=1-4) leads to the formation of the corresponding diiron complexes with intramolecular and intermolecular bis(phosphine) substitution.When these complexes were heated in refluxing xylene.only in the complexes with intermolecular bis(phosphine )substitution the thermal rearrangement reaction occurred.  相似文献   

13.
Iron salts and N-heterocyclic carbene (NHC) ligands is a highly effective combination in catalysis, with observed catalytic activities being highly dependent on the nature of the NHC ligand. Detailed spectroscopic and electronic structure studies have been performed on both three- and four-coordinate iron(ii)–NHC complexes using a combined magnetic circular dichroism (MCD) and density functional theory (DFT) approach that provide detailed insight into the relative ligation properties of NHCs compared to traditional phosphine and amine ligands as well as the effects of NHC backbone structural variations on iron(ii)–NHC bonding. Near-infrared MCD studies indicate that 10Dq(T d) for (NHC)2FeCl2 complexes is intermediate between those for comparable amine and phosphine complexes, demonstrating that such iron(ii)–NHC and iron(ii)–phosphine complexes are not simply analogues of one another. Theoretical studies including charge decomposition analysis indicate that the NHC ligands are slightly stronger donor ligands than phosphines but also result in significant weakening of the Fe–Cl bonds compared to phosphine and amine ligands. The net result is significant differences in the d orbital energies in four-coordinate (NHC)2FeCl2 complexes relative to the comparable phosphine complexes, where such electronic structure differences are likely a significant contributing factor to the differing catalytic performances observed with these ligands. Furthermore, Mössbauer, MCD and DFT studies of the effects of NHC backbone structure variations (i.e. saturated, unsaturated, chlorinated) on iron–NHC bonding and electronic structure in both three- and four-coordinate iron(ii)–NHC complexes indicate only small differences as a function of backbone structure, that are likely amplified at lower oxidation states of iron due to the resulting decrease in the energy separation between the occupied iron d orbitals and the unoccupied NHC π* orbitals.  相似文献   

14.
The photochemistry of Mn2(CO)10 in the presence of phosphine ligands has been investigated. Steric hindrance by the phosphine molecules influences the products of the photolysis reaction. With ethyldiphenylphosphine disubstituted symmetrical species are obtained.  相似文献   

15.
New, improved phosphine ligands for palladium-catalyzed butadiene telomerization with methanol have been discovered. Using high throughput experimentation and an Electrospray Ionization-Mass Spectrometry (ESI-MS) investigation of telomerization catalysts solutions, we have identified phosphines of the type P(C6H4-2-OMe)2(C6H5−n(X)n), where X is an electron-withdrawing group, as high selectivity, high activity phosphine ligands for butadiene telomerization with methanol to 1-methoxy-2,7-octadiene (MOD-1). These ligands were designed to mitigate anaerobic oxidation of phosphines by alkylation which was shown by Electrospray Ionization-Mass Spectrometry (ESI-MS) studies to correlate with catalyst death and palladium precipitation in working telomerization catalyst solutions. The best phosphine-promoted catalysts achieve selectivities to desired 1-methoxy-2,7-octadiene of 94% at high butadiene conversion, significantly improved over those achieved commercially by triphenylphosphine.  相似文献   

16.
We developed a new series of single‐component air‐ and moisture‐stable catalysts for alkyne polymerization based on nickelocene complexes containing phosphine and alkynyl ligands. Chlorine, phosphine and alkynyl ligands exhibited great influence on the catalytic activity of the nickelocene complexes. Highly soluble polymers with fairly high molecular weight (Mw 23 400) were obtained in high yields (85%) in the homogeneous polymerization of p‐diethynylbenzene initiated with five nickelocene acetylides (π‐C5H5)LNi(C≡CR) (L = PPh3, PBu3; R = p‐C6H4C≡CH, C6H5, H) under mild conditions.  相似文献   

17.
The synthetic aspects of chemistry of ligands based on naphthalene peri-substituted by heavier Group 15 elements (P, As, Sb, Bi) or Group 16 elements (S, Se, Te) are discussed in this review. An overview of coordination chemistry of these ligands is also given. In general, the area is dominated by bis(phosphines) Nap(PR2)2 and dithiolates Nap(SR)2 (Nap = naphthalene-1,8-diyl), and most of the ligands act with chelating rigid C3-backbones. Whilst all known bis(phosphine) complexes with Ni, Pd and Pt contain unmodified Nap(PR2)2 moieties, the reactions with a variety of metal carbonyls sometimes result in P–C bond cleavage within the ligand. A range of gold complexes with Nap(PR2)2 ligands have been investigated for material applications. NapP2 ligands other than phosphines are also described, these include 1,2-diphosphaacenaphthenes, bis(phosphonites) and bis(phosphine oxides). Group 16 peri-dichalcogenolates used as ligands include NapS2, NapSe2 and NapSSe systems, but no tellurium congeners. Heterodentate ligands discussed in this review include those with NapPN, NapPO, NapPS, NapPF, NapPC and NapSN motifs. Ligands with heavier Group 15 donor atoms (NapAs2, NapSb2) are also reported. All possible oxides of the dithioles (monooxide to tetraoxide) as ligands are also discussed. Areas of interest for further work are outlined.  相似文献   

18.
In tris(tert-butyl)­phosphine selenide, C12H27PSe, all the methyl ligands are disordered over two sites in the ratio 70/30. The mol­ecule displays crystallographic C3 symmetry. The bond angles at the P atom are distorted tetrahedral [C—P—C 110.02 (5)° and Se=P—C 108.91 (5)°]. The P—C and P=Se bond lengths are 1.908 (1) and 2.1326 (6) Å, respectively. A comparison of the structural data of the complete series of tris(tert-butyl)­phosphine chalcogenides (tBu3PO, tBu3PS, tBu3PSe and tBu3PTe) with the corresponding data of other phosphine chalcogenides substituted by smaller organic groups shows the great influence of the bulky tert-butyl ligands.  相似文献   

19.
The reactivity of secondary phosphine oxides containing bulky organic fragments in hydrophosphinylation reactions has been investigated using several rhodium based catalysts. Upon heating in a focused microwave reactor, HP(O)(2-C6H4Me)2 adds to prototypical terminal alkynes affording a complex mixture containing 1,2 and 1,1-addition products. Addition of a second ortho-substituent (HP(O)Mes2) completely suppresses the hydrophosphinylation reaction for alkyl and aryl substituted alkynes. Variations in the temperature, catalyst loading, solvent, and microwave power were unable to induce an addition reaction in the case of HP(O)Mes2. While this secondary phosphine oxide did not participate in the hydrophosphinylation reaction, it promoted the polymerization of phenylacetylene. HP(O)R2 substrates are not commonly thought of as innocent ligands for rhodium complexes in reactions involving alkynes due to facile hydrophosphinylation. While this is certainly true for diphenylphosphine oxide, the chemistry presented herein suggests that HP(O)Mes2 and related bulky secondary phosphine oxides have great potential as valuable ligands for rhodium catalyzed transformations involving alkynes due to their lack of reactivity towards the addition reaction.  相似文献   

20.
Abstract

In this article, five diiron 1,2-dithiolate complexes containing phosphine ligands are reported. Treatment of complex [Fe2(CO)6(μ-SCH2CH2S)] (1) with the phosphine ligands tris(4-methylphenyl)phosphine, tris(4-methoxyphenyl)phosphine, tris(3-chlorophenyl)phosphine, tris(3-methylphenyl)phosphine, or 2-(diphenylphosphino)biphenyl in the presence of Me3NO·2H2O as the decarbonylating agent afforded the target products [Fe2(CO)5(L)(μ-SCH2CH2S)] [L?=?P(4-C6H4CH3)3, 2; P(4-C6H4OCH3)3, 3; P(3-C6H4Cl)3, 4; P(3-C6H4CH3)3, 5; Ph2P(2-C6H4Ph), 6] in 80–93% yields. Complexes 26 have been characterized by elemental analysis, spectroscopy, and X-ray crystallography. Additionally, the electrochemical properties were studied by cyclic voltammetry.  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