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.
Isoprene, 1,3-butadiene and 2,3-dimethyl-1,3-butadiene react with HFe(CO)4SiCl3 by addition of the Fe---H function to the diene. Isoprene appears to add predominantly 1,4 and 2,3-dimethyl-1,3-butadiene appears to add 1,2, while 1,3-butadiene may add both ways. In the case of isoprene and 1,3-butadiene loss of CO from the addition compound gives a stable π-allyl- Fe(Co)3SiCl3 product. Either cis- or trans-1,3-pentadiene is reduced to pentene by HFe(CO)4SiCl3.  相似文献   

2.
Planar nickel(II) complexes involving N‐(2‐Hydroxyethyl)‐N‐methyldithiocarbamate, such as [NiX(nmedtc)(PPh3)] (X = Cl, NCS; PPh3 = triphenylphosphine), and [Ni(nmedtc)(P‐P)]ClO4(P‐P = 1,1‐bis(diphenylphosphino)methane(dppm); 1,3‐bis(diphenylphosphino)propane (1,3‐dppp); 1,4‐bis(diphenylphosphino)butane(1,4‐dppb) have been synthesized. The complexes have been characterized by elemental analyses, IR and electronic spectroscopies. The increased νC–N value in all the complexes is due to the mesomeric drift of electrons from the dithiocarbamate ligands to the metal atom. Single crystal X‐ray structure of [Ni(nmedtc)(1,3‐dppp)]ClO4·H2O is reported. In the present 1,3‐dppp chelate, the P–Ni–P angle is higher than that found in 1,2‐bis(diphenylphosphino)ethane‐nickel chelates and lower than 1,4‐bis(diphenylphosphino)butane‐nickel chelates, as a result of presence of the flexible propyl back bone connecting the two phosphorus atoms of the complex.  相似文献   

3.
Treatment of a benzene or a CH2Cl2 solution of bis(N,N-dimethylcarbamoylseleno)methanes with SnCl4 afforded β-1,3,5-triselenanes, and the key intermediates, acylselonium ions and selenoaldehydes, were successfully trapped by using allyltrimethylsilane or 2,3-dimethyl-1,3-butadiene to obtain the allylation products or the cycloadducts, respectively.  相似文献   

4.
A series of trifluoromethanesulfonate (OTf) salts of N-heterocyclic phospheniums (NHP) bearing phenyl (1a), para-methoxyphenyl (1b), 2,6-diisopropylphenyl (1c) and mesityl (1d) substituents is reported. The compounds are made by a modification to a literature procedure that improves the overall yields for and by 15 and 23%, respectively. Two unwanted side-products in the synthesis of , the diammonium salt, [(2,6-iPr-C6H3)N(H)2CH2CH2N(H)2(2,6-iPr-C6H3)]Cl2 (4) and the bisphosphine (2,6-iPr-C6H3)N(PCl2)CH2CH2N(PCl2)(2,6-iPr-C6H3) (5), are crystallographically characterized, as is the intermediate cyclic chlorophosphine, C1PN(4-OMe-C6H4)CH2CH2N(4-OMe-C6H4) (3b). The phenyl-substituted NHP is fully characterized, including by X-ray crystallography, for the first time; this compound contains a short P-O contact of 2.1850(14) A. Cycloaddition reactions of with 2,3-dimethyl-1,3-butadiene give the expected spirocyclic phospholeniums, 7,8-dimethyl-1,4-diaryl-1,4-diaza-5-phopshoniaspiro[4.4]non-7-ene, as their OTf salts (6a-d), while reactions with N,N'-dimesityl-1,4-diaza-1,3-butadiene give, except in the case of , which is too bulky to react, the aza analogues, 1,4-dimesityl-6,9-diaryl-1,4,6,9-tetraaza-5-phosphoniaspiro[4.4]non-2-ene (7a, 7b and 7d). The sterically congested is in thermal equilibrium with and free diazadiene, and undergoes a substitution reaction with 2,3-dimethyl-1,3-butadiene to give .  相似文献   

5.
A new polyoxometalate-templated manganese-Schiff-base compound 1, {[Mn(L)2]2[PMo12O40][Cl] · DMF · 2CH3CN · CH3OH} n (where L is 1,4-bis(4-imidazolyl)-2,3-diaza-1,3-butadiene, DMF is N,N-dimethylformamide) has been synthesized by introducing the Metal-Schiff-base into polyoxometalates(POMs) at room temperature, and structurally characterized by elemental analysis, infrared spectroscopy, and single-crystal X-ray diffraction analysis. The result of the single crystal X-ray diffraction suggested that the compound has the packing of the Mn(II)-Schiff-base cation layer and Keggin anion layer.  相似文献   

6.
The reaction of carbonylative addition of alkyl alkynes to aniline derivatives has been successfully achieved by a catalytic system formed of Pd(OAc)2 and a suitable bidentate phosphine ligand. The reaction led mainly to gem‐α,β‐unsaturated amides ( 3 ) with Pd(OAc)2/1,3‐bis(diphenylphosphino)propane/p‐toluenesulfonic acid/CO as the catalytic system. However, the reaction catalyzed by Pd(OAc)2/1,4‐bis(diphenylphosphino)butane/H2/CO in CH2Cl2 as a solvent affords trans‐α,β‐unsaturated amides ( 4 ) as the major product. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

7.
Olefin-diene copolymerizations in the presence of C2 symmetric zirconocene rac-[CH2(3-tert-butyl-1-indenyl)2]ZrCl2/MAO catalytic system have been reported and rationalized by experimental and molecular modeling studies. Ethene gives 1,2-cyclopropane and 1,2-cyclopentane, 1,3-cyclobutane, and 1,3-cyclopentane units in copolymerization with 1,3-butadiene, 1,4-pentadiene, and 1,5-hexadiene, respectively. Propene-1,3-butadiene copolymerizations lead to 1,2 and 1,4 butadiene units and to a low amount of 1,2-cyclopropane units.  相似文献   

8.
The cationic polymerizations of dimethyl-1,3-butadienes with various catalysts in methylene chloride and toluene have been investigated. The activity of catalysts decreased in the order WCl6 > AcClO4 > SnCl4·TCA > BF3OEt2. The homopolymerization rate of dimethyl-1,3-butadienes with WCl6, AcClO4, and SnCl4·TCA decreased in the order 1,3-dimethyl-1,3-butadiene > 2,3-dimethyl-1,3-butadiene > 1,2-dimethyl-1,3-butadiene > 2,4-hexadiene. The polymers prepared with WCl6, SnCl4.TCA, and BF3OEt2 were rubberlike polymers or white powders, whereas those prepared with AcClO4 were oily oligomers. The 1,4-propagation increased in the order 1,2-dimethyl-1,3-butadiene < 1,3-dimethyl-1,3-butadiene < 2,3-dimethyl-1,3-butadiene < 2,4-hexadiene. This order may indicate that the steric effect of methyl group determine primarily the microstructure of the polymer. The relative reactivity of dimethyl-1,3-butadienes toward a styryl cation decreased in the order 1,3-dimethyl-1,3-butadiene > 1,2-dimethyl-1,3-butadiene > 2,3-dimethyl-1,3-butadiene > 2,4-hexadiene. This order may be explained in terms of the stability of the resulting allylic cation.  相似文献   

9.
The reaction between 2,3-dichloromaleic acid dialkylester (alkyl=CH3 and C2H5) and diphenyl(trimethylsilyl)phosphine, leading to diphenylphosphine substituted esters of maleic and fumaric acid has been studied. With a molar ratio 1:1 of the components 2-chloro-3-diphenylphosphinomaleic acid dimethylester (3) and-diethylester are obtained as colourless crystalline compounds. From a 1:2 reaction however only bis(diphenylphosphino)fumaric acid dimethylester (colourless crystals) and-diethylester (yellow) can be crystallized, the latter in a partially oxydized form. The presence of bis(diphenylphosphino)maleic acid diester in the oily part of the reaction products has been proved by its chelating with Ni2+ and Pd2+ to complexes of the compositionMeCl2·(PP). Pure bis(diphenylphosphino)maleic acid dimethylester (4) could be synthesized by alcoholysis and following methylation of bis (diphenylphosphino)maleic anhydrid. Contrary to this easily chelating and air stable compound the corresponding fumaric acid diesters give no complexes with the metals examined as far and are very sensitive towards oxygen. This sensitivity decreases strongly after oxydation to 2-diphenylphosphino-3-diphenylphosphorylfumaric acid diester, the diethylester of which could be crystallized in pure form.Characteristic vibration bands, uv/vis-absorption and31P-nmr peaks are discussed.The result of X-ray diffraction data of3 and4 are reported and conformation, bond lengthes and bond angles of these molecules are given.  相似文献   

10.
In the title compound, diaqua­bis(1,4‐di‐4‐pyrid­yl‐2,3‐diaza‐1,3‐butadiene)dimethanolzinc(II) bis­(perchlorate) 1,4‐di‐4‐pyrid­yl‐2,3‐diaza‐1,3‐butadiene methanol 1.72‐solvate 1.28‐hydrate, [Zn(C12H10N4)2(CH4O)2(H2O)2](ClO4)2·C12H10N4·1.72CH4O·1.28H2O, determined at ca 110 K, the Zn cation and the extended dipyridyl ligand both lie across inversion centres in space group P. The structure consists of a network arrangement of the constituent species stabilized by a combination of coordination, hydrogen bonding and π–π forces. Uncoordinated methanol and water solvent mol­ecules occupy the otherwise void spaces within and between the networks.  相似文献   

11.
Reaction of lithiated bis(diphenylphosphino)amine, [(C6H5)2P]2N? Li+, with K2PtCl4 or PdCl2 (in the presence of trimethylphosphine) yields the homoleptic bis[bis(diphenylphosphino)amide] complexes I and II, respectively. With NiCl2/(CH3)3P the chloro-bridged dinuclear complex III is obtained. A symmetrical bonding of the Ph2P-·N-·PPh2 anion to the metal through the phosphorus atoms is indicated for these diamagnetic, deep-yellow (I, II) or red (III) complexes by 31P NMR spectroscopy (J(PtP) 1812 Hz for I). I and II are dissolved in CF3COOH with protonation at the nitrogen atoms to give bis(diphenylphosphino)amine complexes IV and V, respectively (J(PtP) 2080 Hz for IV). IV and V are 1:2 electrolytes in CH3NO2. Methylation of I-III with CH3OSO2F leads to the bis(diphenylphos-phino)methylamine complexes VI-VIII, of which the palladium compound VII has been structurally characterized by single crystal X-ray diffraction. VII contains a planar CNP2PdP2NC skeleton and is thus based on planar ligand arrays both at Pd and at the two N atoms.  相似文献   

12.
Some 1,4‐phenylene‐bis[1,2,4]oxadiazolo‐[5,4‐d][1,5]benzothiazepine derivatives ( 4a , 4b , 4c ) were synthesized by 1,3‐dipolar cycloaddition reaction of benzohydroximinoyl chloride with 1,4‐phenylene‐bis(4‐aryl)‐2,3‐dihydro[1,5]benzothiazepine ( 2a , 2b , 2c ); meanwhile, compounds 2a , 2b , 2c also occurred ring contraction under acylating condition to obtain bis[2‐aryl‐2′‐(β‐1,4‐phenylenevinyl)‐3‐acetyl]‐2,3‐dihydro[1,5]benzothiazoles ( 3a , 3b , 3c ). The structures of some novel compounds were confirmed by IR, 1H‐NMR, elemental, and X‐ray crystallographic analysis.  相似文献   

13.
Rhodium‐catalyzed 1,4‐addition of lithium 5‐methyl‐2‐furyltriolborate ([ArB(OCH2)3CCH3]Li, Ar=5‐methyl‐2‐furyl) to unsaturated ketones to give β‐furyl ketones was followed by ozonolysis of the furyl ring for enantioselective synthesis of γ‐oxo‐carboxylic acids. [Rh(nbd)2]BF4 (nbd=2,5‐norbornadiene) chelated with 2,2′‐bis(diphenylphosphino)‐1,1′‐binaphthyl (binap) or 2,3‐bis(diphenylphosphino)butane (chiraphos) gave high yields and high selectivities in a range of 91–99 % ee at 30 °C in a basic dioxane/water solution. The corresponding reaction of unsaturated esters, such as methyl crotonate, had strong resistance under analogous conditions, but the 1,4‐adduct was obtained in 70 % yield and with 94 % ee when more electron‐deficient phenyl crotonate was used as the substrate.  相似文献   

14.
The palladium(0)‐catalyzed polyaddition of bifunctional vinyloxiranes [1,4‐bis(2‐vinylepoxyethyl)benzene ( 1a ) and 1,4‐bis(1‐methyl‐2‐vinylepoxyethyl)benzene ( 1b )] with 1,3‐dicarbonyl compounds [methyl acetoacetate ( 4 ), dimethyl malonate ( 6 ), and Meldrum's acid ( 8 )] was investigated under various conditions. The polyaddition of 1 with 4 was carried out in tetrahydrofuran with phosphine ligands such as PPh3 and 1,2‐bis(diphenylphosphino)ethane (dppe). Polymers having hydroxy, ketone, and ester groups in the side groups ( 5 ) were obtained in good yields despite the kinds of ligands employed. The number‐average molecular weight value of 5b was higher than that of 5a . The polyaddition of 1b and 6 was affected by the kinds of ligands employed. The corresponding polymer 7b was not obtained when PPh3 and 1,2‐bis(diphenylphosphino)ferrocene were used. The polyaddition was carried out with dppe as the ligand and gave polymer 7b in a good yield. The molecular weight of the polymer obtained from 1b and 8 was much higher than those of polymers 5b and 7b . The polyaddition with Pd2(dba)3 · CHCl3/dppe as a catalyst (where dba is dibenzylideneacetone) produced polymer 9b in a 92% yield (number‐average molecular weight = 45,600). The stereochemistries of all the obtained polymers were confirmed as an E configuration by the coupling constant of the vinyl proton. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 2487–2494, 2002  相似文献   

15.
The smooth reaction of 3‐chloro‐3‐(chlorosulfanyl)‐2,2,4,4‐tetramethylcyclobutanone ( 3 ) with 3,4,5‐trisubstituted 2,3‐dihydro‐1H‐imidazole‐2‐thiones 8 and 2‐thiouracil ( 10 ) in CH2Cl2/Et3N at room temperature yielded the corresponding disulfanes 9 and 11 (Scheme 2), respectively, via a nucleophilic substitution of Cl? of the sulfanyl chloride by the S‐atom of the heterocyclic thione. The analogous reaction of 3‐cyclohexyl‐2,3‐dihydro‐4,5‐diphenyl‐1H‐imidazole‐2‐thione ( 8b ) and 10 with the chlorodisulfanyl derivative 16 led to the corresponding trisulfanes 17 and 18 (Scheme 4), respectively. On the other hand, the reaction of 3 and 4,4‐dimethyl‐2‐phenyl‐1,3‐thiazole‐5(4H)‐thione ( 12 ) in CH2Cl2 gave only 4,4‐dimethyl‐2‐phenyl‐1,3‐thiazol‐5(4H)‐one ( 13 ) and the trithioorthoester derivative 14 , a bis‐disulfane, in low yield (Scheme 3). At ?78°, only bis(1‐chloro‐2,2,4,4‐tetramethyl‐3‐oxocyclobutyl)polysulfanes 15 were formed. Even at ?78°, a 1 : 2 mixture of 12 and 16 in CH2Cl2 reacted to give 13 and the symmetrical pentasulfane 19 in good yield (Scheme 5). The structures of 11, 14, 17 , and 18 have been established by X‐ray crystallography.  相似文献   

16.
bis(alkoxycarbonyl) complexes of platinum of the type [Pt(COOR)2L] [L = 1,2-bis(diphenylphosphino)ethane (dppe), 1,3-bis(diphenylphosphino)propane (dppp), l,4-bis(diphenylphosphino)butane (dppb), 1,1'-bis(diphenylphosphino)ferrocene (dppf) or 1,2-bis-(diphenylphosphino)benzene (dpb); R = CH3, C6H5 or C2H5] were obtained by reaction of [PtCl2L] with carbon monoxide and alkoxides. Palladium and nickel complexes gave only carbonyl complexes of the type [M(CO)L] or [M(CO)2L]. The new complexes were characterized by chemical and spectroscopic means. The X-ray structure of [Pt(COOCH3)2(dppf] · CH3OH is also reported. The reactivity of some alkoxycarbonyl complexes was also investigated.  相似文献   

17.
A new organometallic-inorganic hybrid compound {[Mn(L)2]2[PMo12O40][NO3] · 4CH3OH · 3H2O} n (I), where L is 1,4-bis(4-imidazolyl)-2,3-diaza-1,3-butadiene) was synthesized in situ from [PMo12O40]3?, Mn2+ and L reaction in an H2O-DMSO-CH3OH mixture at room temperature, and the product was structurally characterized by elemental analysis, infrared spectroscopy, and single-crystal X-ray diffraction analysis. Compound I is constructed from alternating layers of the metal-Schiff-base cation and Keggin anions based on electrostatic forces and hydrogen bond interactions.  相似文献   

18.
1,3-Cyclooctadiene reacts with trimethylaluminum and potassium according to the equation 4K + 4Al(CH3)3 + 2C8H12
3K[Al(CH3)4]+K[(C8H12)2Al] to give potassium bis(3,8-cis-cyclooctenyl) aluminate. The compound can be described as a bicyclo derivative of cyclooctene formed by the 1,4 addition to 1,3-cyclooctadiene. The structure of the complex was determined from 4500 unique data measured by single crystal X-ray diffractometer techniques. Full matrix least squares refinement gave final agreement factors of R1 = 0.043 (observed data) and R2 = 0.045 (all data) in the monoclinic space group P21/c (a = 9.658(6), b = 12.220(8) and c = 14.150(9)Å; β = 113.85(1)°; V = 1527.43 Å; Z = 4 for ?calc = 1.23 g cm?3).  相似文献   

19.
1H and 31P NMR spectroscopy are used to determine the nature of the species present in catalytically active solutions prepared by treating [RhCl(C2H4)2]2 with diphosphines and [Rh(norbornadiene)diphosphine]BF4 with hydrogen (diphosphine = 1,3-bis(diphenylphosphino)propane (dppp) and isopropylidene-2,3-dihydroxy-1,4-bis(diphenylphosphino)butane (diop)).  相似文献   

20.
A chiral cyclic carbonate, 4‐vinyl‐1,3‐dioxolan‐2‐one was used as racemic substrate in asymmetric hydroformylation. The catalysts were formed in situ from “pre‐formed” PtCl2(diphosphine) and tin(II) chloride. (2S,4S )‐2,4‐Bis(diphenylphosphinopentane ((S,S )‐BDPP)), (S,S )‐2,3‐O‐izopropylidine‐2,3‐dihydroxy‐1,4‐bis(diphenylphosphino)butane ((S,S )‐DIOP)), and (R )‐2,2′‐bis(diphenylphosphino)‐1,1′‐binaphthyl ((R )‐BINAP)) were used as optically active diphosphine ligands. The platinum‐containing catalytic systems provided surprisingly high activity. The hydroformylation selectivities of up to 97% were accompanied by perfect regioselectivity towards the dioxolane‐based linear aldehyde. The enantiomeric composition of all components in the reaction mixture was determined and followed throughout the reaction. The unreacted 4‐vinyl‐1,3‐dioxolan‐2‐one was recovered in optically active form. The kinetic resolution was rationalized using the enantiomeric composition of the substrate and the products.  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