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 31 毫秒
1.
An X-ray diffraction analysis is carried out for the complex [Pd(Acac)(PPh3)2]BF4 (I), which is a precursor of the active complexes of styrene dimerization and norbornene additive polymerization in the system [(Acac)Pd(PPh3)2]BF4-BF3 · OEt2. In complex I the palladium atom is coordinated by two oxygen atoms of the acetylacetonate ligand and two phosphorus atoms of the triphenylphosphine ligands at the vertices of the distorted square.  相似文献   

2.
The reactions of the individual cationic complex [(PPh3)3Ni]BF4with unsaturated hydrocarbons (ethene, propene, and styrene), as well as the reaction of the -alkyl Ni(I) complex (PPh3)2Ni–CH(COOC2H5)2(obtained in situ) with ethene, were studied using EPR, UV, and NMR spectroscopy. It was found that stable dimeric carbocationic -alkyl Ni(I) complexes are intermediates in the reactions of unsaturated hydrocarbons with cationic Ni(I) complexes. The transformations of unsaturated hydrocarbons on cationic Ni(I) complexes were explained in terms of an experimentally justified ionic coordination mechanism. A stable mononuclear Ni(I) complex with a Ni–C -bond was synthesized and characterized using EPR. An organonickel(I) complex with - and -bonded carbon atoms was identified using EPR and NMR methods.  相似文献   

3.
Neutral nickel and palladium σ-acetylide complexes [Ni(CCPh)2(PBu3)2] and [Pd(CCPh)2(PBu3)2] are novel initiators for the polymerization of styrene in CHCl3 over a range of polymerization temperature from 40 to 60 °C. Between them, the nickel catalyst exhibited much higher activity than the palladium catalyst. The polystyrene obtained with Ni(II) initiator was a syndio-rich atactic polymer and its weight-average molecular weight reached 279 000. The mechanism of the polymerization was discussed and a radical mechanism was proposed.  相似文献   

4.
The catalytic activity of a series of indenylnickel(Ⅱ) halides: (1-R-Ind)Ni(PPh3)X (R=ethyl, cyclopentyl and benzyl, while X=Cl, Br and I), towards styrene polymerization was studied in the presence of NaBPh4 and PPh3. The catalytic property of these halides was related to the substituent group on the indenyl ligand and the halogen atom bonded to the metal atom. Among them, the (1-Et-Ind)Ni(PPh3)Cl/NaBPha/PPh3 system showed the highest activity for the polymerization of styrene, and the polystyrene obtained was a syndio-rich (rr triad) atactic polymer with Mn values in the range of 103--104. The mechanism of the styrene polymerization initiated by the (1-Et-Ind)Ni(PPh3)Cl/NaBPha/PPh3 system was studied.  相似文献   

5.
The catalytic properties and formation mechanism of alkene dimerization-active complexes in systems based on Ni(PPh3)4 and boron trifluoride etherate are considered. The nature of the modifying action of Brønsted acids on the properties of metal complex catalysts for propylene dimerization is reported. The interaction between Ni(PPh3)4 and BF3 · OEt2 is influenced by water. Depending on the water concentration, the reaction can proceed via formally one-electron oxidation to yield cationic Ni(I) complexes or via two-electron oxidation to yield Ni(II) hydrides. The catalytically active species in alkene dimerization and oligomerization in these systems are Ni(II) hydrido complexes.  相似文献   

6.
The influence of the order of introduction of promoters (complex protonic acids) on the formation of active complexes in the Ni(PPh3)4/BF3 · OEt2 catalytic system and the activity of these systems in ethylene oligomerization have been studied. The activity of the systems in which nickel exists mainly as cationic Ni(I) complexes is more than one order of magnitude higher than the activity of the systems where nickel exists mainly in the form of Ni(II) hydride complexes. The role of alcohols as promoters in the Ni(PPh3)4/BF3 · OEt2 catalytic system is elucidated. The alcohols are the source of Ni(II) hydrides and, more importantly, the source of strong Brønsted acids, which efficiently ensure the coordinative unsaturation of the cationic Ni(I) complexes.  相似文献   

7.
The nickel nitrosyl compound [BseMe]Ni(PPh3)(NO) has been synthesized by the reaction of Ni(PPh3)2(NO)Br with potassium bis(2-seleno-1-methylimidazolyl)hydroborate, [BseMe]K. X-ray diffraction studies demonstrate that (i) the B–H group of the [BseMe] ligand interacts with the nickel center and (ii) the nitrosyl ligand is bent, with Ni–N–O bond angles of 149.1(3)° and 153.1(3)° for the two crystallographically independent molecules. The bent nature of the nitrosyl ligand in [BseMe]Ni(PPh3)(NO) is in marked contrast to the linearity observed for the tris(2-seleno-1-mesitylimidazolyl)hydroborato counterpart [TseMes]NiNO (180.0°). Density functional theory geometry optimization calculations demonstrate that the Ni?H–B interaction is not responsible for causing the nitrosyl ligand to bend, but rather the difference between [TseMes]NiNO and [BseMe]Ni(PPh3)(NO) is due to the [TseMes] ligand allowing the former molecule to adopt a structure with C3 symmetry. In contrast, the steric and electronic asymmetry of [Se2P] donor array of the [BseMe] and PPh3 ligand combination prevents [BseMe]Ni(PPh3)(NO) from having C3 symmetry and the nitrosyl ligand bends to stabilize the occupied M–N σ antibonding orbital.  相似文献   

8.
In this work, we examined the synthesis of novel block (co)polymers by mechanistic transformation through anionic, cationic, and radical living polymerizations using terminal carbon–halogen bond as the dormant species. First, the direct halogenation of growing species in the living anionic polymerization of styrene was examined with CCl4 to form a carbon–halogen terminal, which can be employed as the dormant species for either living cationic or radical polymerization. The mechanistic transformation was then performed from living anionic polymerization into living cationic or radical polymerization using the obtained polymers as the macroinitiator with the SnCl4/n‐Bu4NCl or RuCp*Cl(PPh3)/Et3N initiating system, respectively. Finally, the combination of all the polymerizations allowed the synthesis block copolymers including unprecedented gradient block copolymers composed of styrene and p‐methylstyrene. © 2018 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2019 , 57, 465–473  相似文献   

9.
By merit of dual catalysis of the cationic rare‐earth complex [(η5‐Flu‐CH2‐Py)Ho(CH2SiMe3)2(THF) (Flu = fluorenyl, Py = pyridyl) for the living polymerizations of butadiene (BD) and styrene (St), the crystalline styrene‐butadiene‐styrene (SBS) triblock copolymers consisting of elastic polybutadiene (PBD) sequences with suitable 1,4 regularity (about 70%) and crystalline syndiotactic polystyrene (sPS, [rrrr] > 99%) sequences were successfully synthesized through sequential addition of St, BD, and St monomers. The catalytic system showed high polymerization activities for St and BD in a controlled manner. The crystalline styrene‐b‐(ethylene‐co‐butylene)‐b‐styrene (SEBS) triblock copolymers were obtained by hydrogenation of the above SBS copolymers. The observation of a strong endothermic peak at 266 °C in their differential scanning calorimetry (DSC) curves confirmed the existence of the sPS blocks in the crystalline SEBS different from the industrial product Kraton SEBS‐1652. Thermal degradation temperature of the crystalline SEBS (418 ± 2 °C) indicated the well thermostability and process window of this polymer. © 2016 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 55, 1243–1249  相似文献   

10.
The initiation and catalysis of isobutylene polymerization from several new metallocene and nonmetallocene initiator-catalysts that contain the noncoordinating anions (NCA), B(C6F5)4 and RB(C6F6)3, is reported. Application of these initiator-catalysts is extended to styrenics and vinyl ethers. The NCA does not contribute to termination and can be used in low concentrations compared with conventional Lewis acids. These qualities provide for isobutylene polymerizations that yield low Mn oligomers or high Mn polymer, dependent upon the initiator and polymerization conditions. Mechanistic aspects of initiation, transfer and termination as well as the participation of adventitious water are considered for each class of initiator-catalyst. The influence of the NCA on the stereoregularity of cationic styrene polymerization is also considered. NCAs do not cause the stereospecific carbocationic polymerization of styrene. We suggest that under conditions not conducive to carbocationic polymerization, NCA/metallocenes mediate the coordination polymerization of styrene. © 1997 John Wiley & Sons, Inc.  相似文献   

11.
The reaction of the hydridometal complex [RuClH(CO)(PPh3)3] with 1,4-diphenyl-butadi-1,3-yne has been investigated and found to proceed with monoinsertion to give a coordinatively unsaturated σ-vinyl complex [Ru-{C(CCPh)CHPh} Cl(CO)(PPh3)2], which is also the major product of the reaction of [RuClH(CO) (PPh3)3] with [Hg(CCPh)2].  相似文献   

12.
The Ni-methyl complex (η5-C5H5)Ni(CH3)(PPh3) (1) reacted with B(C6F5)3 to give an unstable contact ion-pair complex with a μ-methyl bridge between the Ni and B atoms. Formation of the B-CH3 bond was confirmed by the reaction of this complex with PPh3 to give [(η5-C5H5)Ni(PPh3)2][B(CH3)(C6F5)3] which was structurally characterized. Spontaneous decomposition of the contact ion-pair complex yielded (η5-C5H5)Ni(C6F5)(PPh3) which is very stable and does not show any reactions with norbornene with or without added B(C6F5)3. 19F NMR study showed that the polynorbornene obtained by the catalysis of 1/B(C6F5)3 system has the C6F5 end-group. A series of reactions, which includes CH3/C6F5 exchange between the Ni and B centers with concomitant dissociation of PPh3 to accept coordination of a norbornene monomer, is proposed as the route to active species that can initiate vinyl polymerization of norbornene.  相似文献   

13.
Complexes of methyl methacrylate and methacrylonitrile with Lewis acids (SnCl4, AlCl3, and BF3) were copolymerized with styrene at ?75°C under irradiation with a high-pressure mercury lamp in toluene solution. The resulting copolymers consisted of equimolar amount of methyl methacrylate or methacrylonitrile and styrene, regardless of the molar ratio of monomers in the feed. NMR spectroscopy showed the copolymers to have an alternate sequence. The tacticities of the copolymers varied with the complex to have an alternate sequence. The tacticities of the copolymers varied with the complex species: the copolymer from the SnCl4 complex system had a higher cosyndiotactieity, while those from the AlCl3 and the BF3 complex systems showed coisotacticity to predominate over cosyndiotacticity. NMR spectroscopic investigation of the copolymerization system indicated the presence of a charge-transfer complex between the styrene and the methyl methacrylate coordinated to SnCl4. The concentration of the charge-transfer complex was estimated to be about 30% of monomer pairs at ?78°C at a 1:1 molar ratio of feed. The growing end radicals were identified as a methyl methacrylate radical for the AlCl3 complex–styrene system and a styrene radical for the SnCl4 complex–styrene system by the measurement of the ESR spectra of the copolymerization systems under or after irradation with a high-pressure mercury lamp. The tacticity of the resulting polymer appears to be controlled by the structure of the charge transfer complex. In the case of the SnCl4 complex a certain interaction of SnCl4 with the growing end radical seems to be a factor controlling the polymer structure. These copolymerizations can be explained by an alternating charge-transfer complex copolymerization scheme.  相似文献   

14.
The binuclear complex [Rh2(μ-S(CH2)2NMe2)2(cod)2] 1 (cod=1,5-cyclooctadiene) was anchored to a sulfonic exchange resin through the residual amine groups. The reaction of the immobilized complex with CO and PPh3 yielded the catalytically active complex [Rh2(μ-S(CH2)2NHMe2)2(CO)2(PPh3)2]2+ supported in the polymer matrix. When methanol was used as solvent, the metal complex loaded cationic resin behaved as a multifunctional catalyst, since it was active in the hydroformylation of styrene and the subsequent formation of the acetals, directly rendering 1,1-dimethoxy-2-phenylpropane in 85% selectivity. Furthermore, the immobilized catalyst can be separated from the reaction mixture and recycled. A homogeneous model of the supported catalyst was generated by reacting complex 1 with HTsO, PPh3, and CO. Thus, the methanol soluble complex [Rh2(μ-S(CH2)2NHMe2)2(CO)2(PPh3)2](TsO)2 was also found to be active in the hydroformylation of styrene yielding identical selectivity in the branched isomer to that of the immobilized catalyst, although the latter is much slower (20-fold) than the homogeneous catalyst.  相似文献   

15.
The neutral complexes (η5-C5H5NiXL (X = Cl, L = PPh3 (I); L = PCy3 (II); X = Br, L = PPh3 (III); L = PCy3 (IV); X = I, L = PPh3 (V); L = PCy3 (VI)) have been obtained by treating NiX2L2 with thallium cyclopentadienide. The same reaction in the presence of TlBF4 gives cationic derivatives [(η5-C5H5)NiL2]BF4 (L = 2PPh2Me (VII); L = dppe (VIII)), whereas mononuclear complexes containing two different ligands (L2 = PPh3 + PCy3 (IX)) or dinuclear [(η5-C5H5)Ni(PPh3)]2dppe(BF4)2 (X) are obtained from the reaction of III with TlBF4 in the presence of a different ligand. Reduction of cationic complexes with Na/Hg gives very unstable nickel(I) derivatives (η5-C5H5)NiL2, which could not be isolated purely. Similar reduction of neutral complexes under CO gives a mixture of decomposition products containing [(η5-C5H5)Ni(CO)]2 and nickel(o) carbonyls, whereas in the presence of acetylenes, dinuclear [(η5-C5H5)Ni]2(RCCR′) (R = R′ = Ph; R = Ph, R′ = H) are obtained.  相似文献   

16.
The cathodic behaviour of electrochemically generated nickel(II) has been investigated in acetonitrile in the presence of triphenylphosphine at a platinum electrode. An appropriate combination of voltammetric, spectrophotometric and NMR findings has allowed us to establish that Ni(II) is present in solution as [Ni(PPh3)2 (CH3CN)42+.]. For the reduction of this species an EECE mechanism is proposed which is consistent with the data. It undergoes an irreversible two-electron reduction giging the Ni(0) complex [Ni(PPh3)4] which reacts quickly with the depolarizer. In this last homogeneous redox reaction the not previously reported Ni(I) complex [Ni(PPh3)4+] is obtained. The degree of reversibility of the redox processes involved has been discussed taking into account the structure, the coordination number and the nature of the ligands in both the redox partners.  相似文献   

17.
Reactions of aluminium tribromide with the Ni(0) phosphine and phosphite complexes are studied by EPR method. AlBr3was found to cause the oxidation of the transition metal in the (PPh3)4Ni complex to the univalent state with the formation of the tetracoordinated (PPh3)3NiBr complex. With an excess of AlBr3, the phosphine ligands are eliminated from the coordination sphere of Ni(I), and the coordinatively unsaturated complexes are destroyed to give the colloidal nickel. In the reaction of (P(OEt)3)4Ni with AlBr3, Ni(0) is also oxidized to Ni(I), but the acido ligand is not eliminated even with a 15-fold excess of the Lewis acid. The activity of catalytic systems on the basis of the Ni(0) phosphine complexes and the Lewis acids in the low-molecular oligomerization reactions of olefines is determined by the cationic coordinatively unsaturated Ni(I) complexes formed in these systems.  相似文献   

18.
Oxidative addition of N-alkyl-2-halopyrimidinium cations to [Pd(PPh3)4] gives straightforward access to the cationic complexes [(PPh3)2(NHC)PdX]BF4 (3a,b) with pyrimidine-derived NHC-ligands. The new complexes were fully characterized including X-ray crystallography.  相似文献   

19.
The known compound 4′-(carboxyphenyl)-2,2′:6,2″-terpyridine (LH) was prepared and complexed with RuCl3.3H2O. The resulting complex [Ru(LH)Cl3] was then allowed to react separately with 2,2′-bipyridine (bpy), 1,10-phenanthroline (phen), triphenylphosphine (PPh3) and 1,2-bis-(diphenylphosphino)ethane (dppe). The compositions of corresponding complexes [Ru(LH)bpyCl](BF4) 1, [Ru(LH)phenCl](BF4) 2, [Ru(LH)(PPh3)(CH3CN)2] (BF4)23 and [Ru(LH)(dppe)Cl](BF4) 4 were assigned on the basis of their FAB-mass spectra, elemental analysis, spectroscopic (IR, NMR) data and X-ray diffraction measurements. The diamagnetic, cationic complexes displayed strong MLCT transitions in the visible region with significant shift in MLCT band energy corresponding to the strength of substituted ligands. The redox behaviour of the complexes was investigated using cyclic voltammetry measurements. Among all the complexes, 3 efficiently catalyzed the synthesis of propargylamine via three components coupling reaction.  相似文献   

20.
The heterogeneous catalytic polymerization of styrene vapor with a tetrakis(acetonitrile)palladium(II) tetrafluoroborate, [Pd(CH3CN)4][BF4]2, thin film has been demonstrated. The catalyst is deposited by nebulization of dilute solutions onto a quartz crystal microbalance (QCM) and then exposed to styrene vapor in controlled environments. The use of QCM allows in situ monitoring of catalyst deposition and polymer growth kinetics. The polymerization process appears to involve the entire catalyst film rather than polymerization only at the catalyst film surface. The styrene vapor polymerization occurs rapidly after a short induction time needed for monomer dissolution and catalyst activation. The narrow molecular weight distribution of the produced polymer suggests that the deposited film acts as a single site catalyst. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 1930–1934, 2005  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