首页 | 官方网站   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.
Reactions of (CO)5Re(Br), (η5‐C5H5)Ru(Cl)(PPh3)2, and [Pt(μ‐Cl)(C6F5)(S(CH2CH2‐)2)]2 with the alkyne‐containing phosphine Ph2P(CH2)6C≡CCH3 give the bis(phosphine) complexes fac‐(CO)3Re(Br)(Ph2P(CH2)6C≡CCH3)2 ( 5 ), (η5‐C5H5)Ru(Cl)(Ph2P(CH2)6C≡CCH3)2 ( 6 ), and trans‐(Cl)(C6F5)Pt(Ph2P(CH2)6C≡CCH3)2 ( 7 ). Alkyne metatheses with the catalyst (t‐BuO)3W(≡C‐t‐Bu) (10–15 mol %, chlorobenzene, 80 °C) give the seventeen‐membered metallamacrocycles fac‐(CO)3Re(Br)(Ph2P(CH2)6CC(CH2)6P Ph2) ( 8 ), (η5‐C5H5)Ru(Cl)(Ph2P(CH2)6CC(CH2)6P Ph2) ( 9 ), and trans‐(Cl)(C6F5)Pt(PPh2(CH2)6CC(CH2)6P Ph2) ( 10 ). 31P NMR analyses show 90–75% conversions to 8 – 10 (59–47% isolated after chromatography). The identity of 8 was confirmed by a crystal structure, and 10 was hydrogenated over Pd/C to fac‐(CO)3Re(Br)(Ph2P(CH2)6CC(CH2)6P Ph2) ( 12 , 87%), which was crystallographically characterized earlier. A catalyst derived from Mo(CO)6/4‐chlorophenol effects a slower conversion of 7 to 10 at 140 °C. In the case of 5 , a mer, trans isomer of 8 is isolated ( 11 , 44%), as established by NMR and IR data. In 10 – 12 , the diphosphines span trans positions. These results, together with previous examples involving group VIII metallocenes, establish the wide viability of the title reaction.  相似文献   

2.
The mixed pincer palladacycles (Me2NCH2(Cl)CCCH2CH2Y‐κNCY)PdCl ( 1 , Y=PPh2; 2 , OPPh2) and (t‐BuSCH2CH2CC(Cl)(o‐NC5H4)‐κSCN)PdCl 3 have been obtained in high yields by chloropalladation of heterosusbstituted alkynes Me2NCH2CCCH2CH2PPh2, Me2NCH2CCCH2CH2OPPh2 and t‐BuSCH2CH2CC(o‐NC5H4), respectively. The molecular structures of 1 and 3 have been ascertained by means of X‐ray diffraction analysis. The catalytic properties of these mixed donor group pincer‐type palladacycles have been evaluated in the arylation of olefins (Heck reaction). The pincer palladacycle 1 is highly active for the coupling of aryl iodides and aryl bromides with n‐butyl acrylate. In contrast it is only moderately active for the coupling of aryl chlorides substituted with electron‐withdrawing groups and inactive for the coupling of electron neutral and electron deactivated aryl chlorides.  相似文献   

3.
The new complexes RuHCl(PPh2CH2CHRNH2)2 and RuHCl(PPh2CH2CHRNH2)(R‐ binap), R=H (Pgly), R=Me [(R)‐Pala] were prepared by the substitution of the PPh3 ligands in RuHCl(PPh3)3 or RuHCl(PPh3)[(R)‐binap] with beta‐aminophosphines derived from amino acids. The complex trans‐RuHCl(Pgly)[(R)‐binap] has been characterized by X‐ray crystallography. The complex trans‐RuHCl[(S)‐Ppro]2 where (S)‐Ppro is derived from proline was also prepared and characterized by X‐ray crystallography. These were used as catalyst precursors in the presence of a base (KOPr‐i or KOBu‐t) for the hydrogenation of various ketones and imines to the respective alcohols and amines with H2 gas (1–11 atm) at room temperature. Acetophenone was hydrogenated to (S)‐1‐phenylethanol in low ee (up to 40%) when catalyzed by the enantiomerically pure complexes. These complexes are especially active in the hydrogenation of sterically congested and electronically deactivated ketones and imines and are selective for the hydrogenation of CO bonds over CC bonds.  相似文献   

4.
The oxorhenium(V) chelates [ReOCl(N,O‐L)(PPh3)] [N,O‐L=(OCH2CH2)N(CH2CH2OH)(CH2COO) ( 2 ), (OCH2CH2)N(CH2COO)(CH2COOCH3) ( 3 )] and [ReOCl2(N,O‐L)(PPh3)] [N,O‐L=C5H4N(COO‐2) ( 4 ) C5H3N(COOCH3‐2)(COO‐6) ( 5 )] have been prepared by reaction of [ReOCl3(PPh3)2] ( 1 ), in refluxing methanol, with N,N‐bis(2‐hydroxyethyl)glycine [bicine; N(CH2CH2OH)2(CH2COOH)], N‐(2‐hydroxyethyl)iminodiacetic acid [N(CH2CH2OH)(CH2COOH)2], picolinic acid [NC5H4(COOH‐2)] or 2,6‐pyridinedicarboxylic acid [NC5H3(COOH‐2,6)2], respectively, with ligand esterification in the cases of 3 and 5 . All these complexes have been characterized by IR and multinuclear NMR spectroscopy, FAB+‐MS, elemental and X‐ray diffraction structural analyses. They act as catalysts, in a single‐pot process, for the carboxylation of ethane by CO, in the presence of potassium peroxodisulfate K2S2O8, in trifluoroacetic acid (TFA), to give propionic and acetic acids, in a remarkable yield (up to ca. 30%) and under relatively mild conditions, with some advantages over the industrial processes. The picolinate complex 4 provides the most active catalyst and the carboxylation also occurs, although much less efficiently, by the TFA solvent in the absence of CO. The selectivity can be controlled by the ethane and CO pressures, propionic acid being the dominant product for pressures about ca. 7 and 4 atm, respectively (catalyst 4 ), whereas lower pressures lead mainly to acetic acid in lower yields. These reactions constitute an unprecedented use of Re complexes as catalysts in alkane functionalization.  相似文献   

5.
Homogenous polymerization of methyl methacrylate using Pd(II)- and Ni(II)-based acetylide complexes as initiators has been investigated. M(PR'3)2(CCR)2 (M=Pd, Ni; R'=PPh3, Pn-Bu3; R=Ph, CH2OH, CH2OOCCH3, CH2OOCPh, CH2OOCPhOH-o) were found to be a novel type of effective initiators for the polymerization of methyl methacrylate. Among them, Pd(C CPh)2(PPh3)2 (PPP) shows the highest activity in the MMA polymerization and the PMMA obtained is a syndiotactic polymer with high number-average molecular weight (M n) of 14.1 × 104. Some features and kinetic behavior of MMA polymerization initiated by PPP were studied in detail. The polymerization reaction is first-order with respect to both [PPP] and [MMA]. Radical polymerization mechanism is proposed.  相似文献   

6.
Soluble, thermally stable, and easily processable platinum(II) polyyne polymers of the form trans-[–Pt(PBu3)2C≡CRC≡C–]n (R = 2,8-disubstituted dibenzothiophene, 2,8-disubstituted and 3,7-disubstituted dibenzothiophene-S,S-dioxide units) have been prepared in good yields by CuI-catalyzed polymerization involving the dehydrohalogenating coupling of trans-[PtCl2(PBu3)2] and HC≡CRC≡CH. We report their optical absorption and photoluminescence spectra and compare the results with the monomeric model complexes trans-[Pt(Ph)(PEt3)2C≡CRC≡CPt(Ph)(PEt3)2]. The different electronic properties and linkage geometry of the central R group lead to new organometallic materials with distinct photophysical traits. The polymer with more conjugated 3,7-disubstituted dibenzothiophene-S,S-dioxide mainly displays fluorescence band, while its isomer with less conjugated 2,8-disubstituted spacer effectively enhances the intersystem crossing rate and strong phosphorescence emission can be detected even at room temperature.  相似文献   

7.
An N‐propargylamide monomer, CH?CCH2NHCOC(CH3)2CH2CH3 (monomer 9), was polymerized in the presence of (nbd)Rh+B?(C6H5)4 (nbd represents norbornadiene) in CH2Cl2, CHCl3, tetrahydrofuran or dimethylformamide, to provide polymers with moderate number‐average molecular weights (Mn = 8700–12 100 g mol?1) in high yields (≥92%). The resulting poly(N‐propargylamide) (polymer 9) dissolves almost completely in CHCl3 (>95%). According to the UV‐visible spectra, measured at various temperatures, polymer 9 forms relatively stable helices over a wide temperature range (35–65 °C). Moreover, it exhibits reversible conformational transitions from an ordered helix to a random coil. On copolymerization of monomer 9 with CH?CCH2NHCO(CH2)3CH3 (monomer 4) or CH?CCH2NHCO(CH2)7CH3 (monomer 8), the solubility of polymer 9 improves noticeably. All the copolymers form helices under the experimental conditions. From the viewpoint of monomers 4 and 8, copolymerization with monomer 9 is favorable in terms of the copolymers forming helices. These findings reveal that the helical content and thermodynamic stability of the helices formed in the copolymers are likely to be controlled by selecting a suitable comonomer and by adjusting the composition of the copolymer. Copyright © 2007 Society of Chemical Industry  相似文献   

8.
Reaction of ligand, L-CH2Ph with Pd(CH3CN)2Cl2 in acetonitrile in presence of PPh3 and subsequent addition of NaClO4 yielded Pd(II) complex, [Pd(L)(PPh3)](ClO4) 1 due to the SC (CH2Ph) bond cleavage during complexation. Complex 1 was characterized by spectral analysis and the structure was authenticated by single crystal X-ray diffraction. The diffraction analysis revealed that the monoanionic ligand binds with palladium (II) in (N, N, S) fashion in a distorted square planar geometry where the fourth position is occupied by one tri phenyl phosphine group. One ClO4 ion satisfies the charge of the former aggregate, [Pd(L)(PPh3)]+. The complex [Pd(L)(PPh3)](ClO4), 1 exhibits a very good catalytic activity towards CC bond formation.  相似文献   

9.
A novel approach based on conjugation interruption was developed for a luminescent and thermally stable platinum(II) polyyne polymer trans-[–Pt(PBu3)2C≡C(C6H4)CH2(C6H4)C≡C–] n (1) containing the diphenylmethane chromophoric spacer. Particular attention was focused on the photophysical properties of this group 10 polymetallayne and comparison was made to its binuclear model complex trans-[Pt(Ph)(PEt3)2C≡C(C6H4)CH2(C6H4)C≡CPt(Ph)(PEt3)2] (2) and their closest group 11 gold(I) and group 12 mercury(II) neighbors, [MC≡C(C6H4)CH2(C6H4)C≡CM] (M = Au(PPh3) (3), HgMe (4)). The regiochemical structures of these angular-shaped compounds were studied by various spectroscopic analyses. Upon photoexcitation, each of them emits an intense purple-blue fluorescence emission in the near UV region in dilute fluid solutions at room temperature. Harvesting of organic triplet emissions harnessed through the strong heavy-atom effects of group 10–12 transition metals was examined. These metal-containing phenyleneethynylenes spaced by the conjugation-breaking CH2 unit were found to have high optical gaps and high-energy triplet states. The influence of metal and sp3-hybridized methylene conjugation-interrupters on the intersystem crossing rate and the spatial extent of the lowest singlet and triplet excitons was fully elucidated. Our investigations indicate that high-energy triplet states in these materials intrinsically give rise to very efficient phosphorescence with fast radiative decays. Dedicated to Professor Didier Astruc in recognition of his outstanding contribution to metallodendrimers and polymers.  相似文献   

10.
The reaction of the Cu(II) bis N,O‐chelate‐complexes of L‐2,4‐diaminobutyric acid, L‐ornithine and L‐lysine {Cu[H2N–CH(COO)(CH2)nNH3]2}2+(Cl)2 (n = 2–4) with terephthaloyl dichloride or isophthaloyl dichloride gives the polymeric complexes {‐OC–C6H4–CO–NH–(CH2)n–CH(nh2)(COO)Cu(OOC)(NH2)CH–CH2)n–NH‐}x 1 – 5 . From these the metal can be removed by precipitation of Cu(II) with H2S. The liberated ω,ω′‐N,N′‐diterephthaloyl (or iso‐phthaloyl)‐diaminoacids 6 – 10 react with [Ru(cymene)Cl2]2, [Ru(C6Me6)Cl2]2, [Cp*RhCl2]2 or [Cp*IrCl2]2 to the ligand bridged bis‐amino acidate complexes [Ln(Cl)M–(OOC)(NH2)CH–(CH2)nNH–CO]2–C6H4 11 – 14 .  相似文献   

11.
Summary Polymerization of N-substituted 2-propynamides proceeded in the presence of Pd(II) catalysts such as [Pd(CH3CN)4](BF4)2, PdCl2(PhCN)2-n-BuLi, and PdCl2(nbd)-n-BuLi. The molecular weights of the obtained polymers ranged 8,500 to 14,000, depending on the catalysts. From the 1H NMR spectra, these polymers were found to have alternating double bonds in the main chain. Received: 25 December 2000/Revised version: 17 January 2001/Accepted: 19 January 2001  相似文献   

12.
The novel phosphine Pd/Te dimer [(np3)PdTe]2, np3=N(CH2CH2PPh2)3, has been synthesized by reaction of (np3)Pd with TePEt3 and characterized by X-ray diffraction studies.  相似文献   

13.
New polymer-bound hydrogenation catalysts were made by complexing PdCl2, RhCl3·3H2O, or NiCl2 with anthranilic acid anchored to chloromethylated polystyrene. The Pd(II) and Ni(II) polymers were reduced to the corresponding Pd(O) and Ni(O) catalysts with NaBH4. In the hydrogenation of methyl sorbate, these polymer catalysts were highly selective for the formation of methyl 2-hexenoate. The diene to monoene selectivity decreased in the order: Pd(II), Pd(O), Rh(I), Ni(II), Ni(O). Kinetic studies support 1,2-reduction of the Δ4 double bond of sorbate as the main path of hydrogenation. In the hydrogenation of soybean esters, the Pd(II) polymer catalysts proved superior because they were more active than the Ni(II) polymers and produced lesstrans unsaturation than the Rh(I) polymers. Hydrogenation with Pd(II) polymers at 50~100 C and 50 to 100 psi H2 decreased the linolenate content below 3% and increasedtrans unsaturation to 10~26%. The linolenate to linoleate selectivity ranged from 1.6 to 3.2. Reaction parameters were analyzed statistically to optimize hydrogenation. Recycling through 2 or 3 hydrogenations of soybean esters was demonstrated with the Pd(II) polymers. In comparison with commercial Pd-on-alumina, the Pd(II) polymers were less active and as selective in the hydrogenation of soybean esters but more selective in the hydrogenation of methyl sorbate. Presented at ISF-AOCS Meeting, New York, April 1980.  相似文献   

14.
Regioselective Markovnikov‐type addition of PhSH to alkynes (HC≡C‐R) has been performed using easily available nickel complexes. The non‐catalytic side reaction leading to anti‐Markovnikov products was suppressed by addition of γ‐terpinene to the catalytic system. The other side reaction leading to the bis(phenylthio)alkene was avoided by excluding phosphine and phosphite ligands from the catalytic system. It was found that catalytic amounts of Et3N significantly increased the yield and selectivity of the catalytic reaction. Under optimized conditions high product yields of 60–85% were obtained for various alkynes [R=n‐C5H11, CH2NMe2, CH2OMe, CH2SPh, C6H11(OH), (CH2)3CN]. The X‐ray structure of one of the synthesized products is reported.  相似文献   

15.
Ammonium persulfate-initiated cyclopolymerization of maleic acid (HO2CC=CCO2H) (MA) with diallylamine (–NR2, R = CH2 = CH-CH2) derivatives (DADs): R2NH+CH2CO2 ( I ), R2NCH2CO2Na+ ( II ), R2NCH2CO2Et ( III ), R2NH+(CH2)3CO2 ( IV ), R2N(CH2)3CO2Et ( V ), R2NH+(CH2)3SO3 ( VI ) and R2N(CH2)3SO3Na+ ( VII ) gave a series of new pH-responsive alternate copolymers: –[(DAD-alt-MA)]n– VIII - XIV , respectively. Homopolymers XV and XVI of the corresponding monomers IV and V ·HCl were also synthesized. The evaluation of the synthesized polymers as potential antiscalants in reverse osmosis (RO) plants was examined. An effective scale inhibitor must arrest the formation of scale for a residence time (≈15 min) of feed water in the RO chamber. R2NH+(CH2)3CO2 ( IV )-alt-MA copolymer XI at a concentration of 5 ppm imparted 99% scale inhibition for 360 min, while at 2.5 and 1 ppm concentrations, the CaSO4 scale inhibitions were found to be 99% and ≈ 100% during 120 and 20 min, respectively, at 40°C. Homopolymers XV and XVI , at a 20-ppm concentration, demonstrated ≈ 95% efficiency in inhibiting mild steel corrosion in 1 M HCl. In some important oilfield applications, requirement of simultaneous scale and corrosion inhibition make these polymers potential candidates for desalination and oilfield applications.  相似文献   

16.
The reactivity of mononuclear platinum and palladium dithiolato species as building blocks for homo- and heterobimetallic complexes is described. The heterobimetallic cationic complex [(PPh3)2Pt(μ-S(CH2)2S)Pd(dppb)](BF4)2 was obtained by reaction of [Pt(S(CH2)2S)(PPh3)2] with [PdCl2(Ph2P(CH2)4PPh2)] in presence of AgBF4.  相似文献   

17.
New azophenol polymers (P1, P2 and P3) were synthesised by the oxidative polycondensation (OP) reaction of three different azophenol monomers in aqueous alkaline medium with NaOCl as the oxidant. The monomers and the polymers were characterised by elemental analyses, and UV‐visible, Fourier transform infrared, 1H NMR and 13C NMR spectroscopic studies, which revealed that the polymers synthesised by OP are composed of oxyphenylene (C? O? C) and phenylene (C? C) units. The polymers obtained are soluble in dimethylformamide and dimethylsulfoxide. Average molecular weights of the polymers were determined by gel permeation chromatography. Additionally, P2 and P3 are soluble in water and methanol. On the basis of thermogravimetric analyses, 5 and 50% weight‐loss temperatures of the polymers were found to be 218, 700 (P1), 263, 609 (P2) and 100, 809 °C (P3), respectively, suggesting a high thermal stability. Thermal analyses using differential scanning calorimetry revealed that the azophenol polymers are highly amorphous, and melting peaks were not observed in the heating cycles. This suggests that all the polymers are highly amorphous. The azophenol polymers show a reversible transcistrans isomerisation process. These properties of the polymer could be promising for their technological usage. Copyright © 2007 Society of Chemical Industry  相似文献   

18.
The ruthenium (II) diene complexes [Ru(X)(Cl)(nbd)(dppb)] (X = Cl, H; nbd = 2,5-norbornadiene; dppb = PPh2(CH2)4PPh2) have been prepared and characterized spectroscopically. The X-ray crystal structure of RuCl2(nbd)(dppb) (crystal data at 22°C: space group P1, a = 10.896 (1) Å, b = 15.168(2) Å, c = 10.829 (1) Å, α = 103.02(1)°, β = 107.08(1)°, γ = 81.65(1)°, Z = 2, R = 0.054 for 6420 reflections) shows an octahedral geometry at Ru, with the chloro ligands slightly distorted from a trans configuration (Cl)(1)-Ru-C1(2) = 168.4°); the unit cell contains two molecules of the complex and one molecule of benzene. Reaction of this complex with H2, in presence of Proton Sponge (PS, 1,8-bis(dimethylamino)naphthalene) as base, is complicated by initial dissociation of nbd, and [Ru2Cl5(dppb)2]-PSH+ is the major product. A minor product, the hydrido(diene) complex trans-RuCl(nbd)(dppb) 5 , characterized spectroscopically, is more effectively synthesized from (a) trans-Ru(H)Cl(nbd)(PPh3)2, 1 , and dppb, or (b) reaction of RuCl2(dppb)-(PPh3) with H2 in presence of nbd and PS. Complex 5 is unreactive toward H2 or CO while 1 has been shown previously to give η2-H2 and norbornenoyl derivatives, respectively; the differences in reactivity are discussed.  相似文献   

19.
Pt(II) and Ni(II) complexes of N,N-bis(diphenylphosphinomethyl)aminopropyltriethoxysilane [(CH3CH2O)3Si(CH2)3N(CH2PPh2)2] (DIPAPTES) and silica supported [SiO2–(DIPAPES)] ligands have been synthesized under nitrogen atmosphere using Schlenk method and characterized by using atomic absorption, FT-IR, NMR (1H and 31P) and elemental analysis techniques. All the complexes were used as catalysts for the oxidation of 2-methyl naphthalene (2MN) to 2-methyl-1,4-naphthoquinone (vitamin K3, menadione, 2MNQ) using hydrogen peroxide as a clean and cheap oxidant. [Pt–(DIPAPTES)Cl2] and supported complex [SiO2–(DIPAPES)–PtCl2] showed medium catalytic activity whereas Ni(II) complexes did not show any catalytic activity for the selective oxidation of 2-methyl naphthalene to 2-methyl-1,4-naphthoquinone.  相似文献   

20.
An associative mechanism has been computationally characterized for the Stille cross‐coupling of vinyl bromide and trimethylvinylstannane catalyzed by PdL2 (L=PMe3, AsMe3) with or without dimethylformamide as coordinating ligand. All the species along the catalytic cycles that start from both the cis‐ and the trans‐PdL(Y)(vinyl)Br complexes (Y=L or S; L=PMe3, AsMe3 or PH3; S=DMF) have been located in the gas phase and in the presence of polar solvents. Computations support the central role of species trans‐PdL(DMF)(vinyl)Br which react by ligand dissociation and stannane coordination in the rate‐limiting transmetalation step via a puckered four‐coordinate (at palladium) transition state comprised of Pd, Br, Sn and sp2 C atoms. A donating solvent may enter the catalytic cycle assisting isomerization of cis‐PdL2(vinyl)Br to trans‐PdL(DMF)(vinyl)Br complexes via a pentacoordinate square pyramidal Pd intermediate. In keeping with experimental observations, the activation energies of the catalytic cycles with arsines as Pd ligands are lower than those with phosphines. Polytopal rearrangements from the three‐coordinate T‐shaped Pd complexes resulting from transmetalation account for the isomerization and the C C bond formation on the reductive elimination step.  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-23

京公网安备 11010802026262号