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 359 毫秒
1.
Dichlorocobalt(III) complexes of (2S,5S,9S)-trimethyltriethylenetetraamine (L1) and (2S,5R,9S)-trimethyltriethylenetetraamine (L2) have been prepared. Both L1 and L2 coordinate to the cobalt(III) ion to give three isomers: Λ-cis-α, Δ-cis-β, trans isomers for L1 and Δ-cis-α, Δ-cis-β, trans isomers for L2. Each of the trans-dichloro complexes of the two ligands have been isomerized stereospecifically to the cis-α-dichloro complex in methanol, and each of the cis-α-dichloro complexes stereospecifically to the trans-diaqua complex in water. Both the geometrical and optical inversions took place at the same time in the observed stereospecific isomerizations.  相似文献   

2.
The monomeric cis-dioxomolybdenum(VI) complexes [MoO2(oep-saldpen)] and [MoO2Cl2(oep-H2saldpen)], with a tetradentate [N2(imine)O2] and bidentate [N2(imine)] salen-type ligand functionalised with two pyrrole derivative pendant arms [oep-H2saldpen = 1,2-diphenylethylenebis(3-oxyethylpyrrole)salicylideneimine], were synthesised and characterised by 1H NMR, IR and Raman spectroscopy. The solid-state structure of the free ligand oep-H2saldpen was determined by single crystal X-ray diffraction. Assignment of the vibrational spectra of the molybdenum complexes was supported by carrying out ab initio calculations for the possible isomers using [MoO2(salen)] and [MoO2Cl2(H2salen)] as model compounds [H2salen = N,N′-ethylenebis(salicylideneimine)]. The oep-saldpen complexes were examined as catalysts for the epoxidation of cyclooctene, (R)-(+)-limonene, styrene, α-pinene, and cis and trans-β-methylstyrene, with tert-butyl hydroperoxide as the oxidant. Both complexes exhibited high selectivity for the epoxidation reaction, with the bis(chloro) complex being always the more active of the two.  相似文献   

3.
H. Martens  G. Hoornaert  S. Toppet 《Tetrahedron》1973,29(24):4241-4249
The IR and UV spectroscopic data of a series of alkyl substituted β-chlorovinyl ketones have been determined. Some of the IR and UV spectroscopic criteria, developed for the S-cis or S-trans conformation of ordinary α,β-unsaturated ketones, seem to be applicable for the β-chlorovinyl ketones: the S-cis conformer shows a low ratio (ri) of the integrated intensities of the carbonyl and double bond stretching vibrations and a relatively low UV ε value; whereas high ri and ε values are found for the S-trans conformer; non-planarity increases the ri value and decreases the ε value.The prefered conformations as determined by the IR and UV spectral data have led to the presentation of benzene solvation models for a series of β-chlorovinyl ketones. These models made it possible to assign thecis ortrans structure to α,β-dialkyl-β-chlorovinyl ketones on the basis of the NMR aromatic solvent induced shifts of the β-alkyl group: high solvent shifts to a higher field (0·31–0·66 ppm) are found forcis-β-chlorovinyl ketones, whereas low solvent shifts (0·040·15) are found for thetrans-β-chlorovinyl ketones. Assignments based on the chemical shifts alone can lead to erroneous interpretations.  相似文献   

4.
《Polyhedron》1991,10(23-24)
In a direct synthesis reaction of the tris(aminocarboxylato)cobalt(III) complexes with S-arginine all four theoretically possible tris(S-arginine)cobalt(III) diastereomers were obtained as tripositive complex ions. Besides, one out of 24 theoretically possible isomers of di-μ-hydroxo-tetrakis(S-arginine)dicobalt(III) tetrapositive ion was obtained. Geometrical and absolute configurations of the tris(S-arginine)cobalt(III) isomers were determined by electronic and CD spectroscopy. The molecular structure of the (-)589-anti(N)-Δ-cis(N),cis(O)-Λ-cis(N),cis(O)-di-μ-hydroxo-tetrakis(S-arginine)dicobalt(III) ion was solved by X-ray crystal structure analysis. The complexes obtained represent the first examples of cationic aminocarboxylato complexes of this type. For the first time the formation of the di-μ-hydroxo-tetrakis(aminocarboxylato)dicobalt(III) complex has been observed as a reaction concurrent to the formation of tris(aminocarboxylato)cobalt(III) complexes. Finally, the stereoselectivity of S-arginine in the tris(S-arginine)cobalt(III) isomers synthesis was discussed.  相似文献   

5.
Iron-catalyzed highly regio- and enantioselective organic transformations with generality and broad substrate scope have profound applications in modern synthetic chemistry; an example is herein described based on cis-FeII complexes having metal- and ligand-centered chirality. The cis-β FeII(N4) complex [FeII(L)(OTf)2] (L = N,N′-bis(2,3-dihydro-1H-cyclopenta-[b]quinoline-5-yl)-N,N′-dimethylcyclohexane-1,2-diamine) is an effective chiral catalyst for highly regio- and enantioselective alkylation of N-heteroaromatics with α,β-unsaturated 2-acyl imidazoles, including asymmetric N1, C2, C3 alkylations of a broad range of indoles (34 examples) and alkylation of pyrroles and anilines (14 examples), all with high product yields (up to 98%), high enantioselectivity (up to >99% ee) and high regioselectivity. DFT calculations revealed that the “chiral-at-metal” cis-β configuration of the iron complex and a secondary π–π interaction are responsible for the high enantioselectivity.

A cis-β FeII complex having metal- and ligand-centered chirality catalyzes highly regio- and enantioselective alkylation of indoles (at the N1, C2, or C3 position), pyrroles and anilines with α,β-unsaturated 2-acyl imidazoles (48 examples, up to 99% ee).  相似文献   

6.
A general method for the preparation of unsymmetricaltrans-2-allyl-6-alkyl(aryl)-1,2,3,6-tetrahydropyridines6 based on a combination of 1,2-addition of RLi to pyridine andtrans-6-allylation with triallylborane in the presence of methanol was elaborated. It was shown thattrans-piperideines (6 (R=Alk, Ph) isomerize into the correspondingcis-2-allyl-6-alkyl(phenyl)-3-piperideines14 on heating with triallylborane followed by deboronation of aminoborane (16) with methanol and an alkali. The stereochemistry of compounds6 and14 was determined by two-dimensional NOE spectroscopy. A possible mechanism of the formation oftrans-amines6 and their transformation intocis-isomers14 is discussed. Alkaloids (±)-epidihydropinidine (trans-2-methyl-6-propylpiperidine2a, 70%) and (±)-dihydropinidine (cis-2-methyl-6-propylpiperidine1d, 71%) were synthesized by hydrogenation of compound6a (R=Me) and14a (R=Me), respectively, over Raney nickel. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 467–474, March, 1998.  相似文献   

7.
The effects of the variables of head group structure and salt concentration on microemulsions formed in mixtures of water, alkyl ethylene glycol ethers (CkOC2OCk), andn-alkyl β- -glucopyranosides (CmβG1) are explored. Phase behavior of mixtures containing an anomer of the surfactant (n-alkyl α- -glucopyranoside, CmαG1), or surfactants with long head groups (n-alkyl maltopyranosides, CmG2), or NaCl or NaClO4as electrolyte are systematically reported as a function of temperature and composition. The substitution ofn-alkyl α- -glucopyranosides forn-alkyl β- -glucopyranosides causes precipitation under some conditions in all mixtures studied. These solubility boundaries begin in the water–surfactant binary mixture at the Krafft boundary, then extend to high concentrations of both surfactant and oil. Increasing the effective length of the surfactant head group by adding CmG2to water–CkOC2OCk–CmβG1mixtures moves the phase behavior dramatically up in temperature when even small amounts of CmG2are used. Adding a lyotropic electrolyte, NaCl, to water–CkOC2OCk–CmβG1mixtures moves the phase behavior down in temperature, while the hydrotropic electrolyte NaClO4moves the phase behavior up in temperature.  相似文献   

8.
[TiCl2(salen)] (1) reacts with AlMe3 (1:2) to give the heterometallic Ti(III) and Ti(IV) complexes [Ti{(μ-Cl)(AlMe2)}{(μ-Cl)(AlMe2X)}(salen)] (X=Me or Cl) (2) and [TiMe{(μ-Cl)(AlCl2Me)}(salen)] (3). Addition of diethyl ether to 3 affords [Ti(Me)Cl(salen)] (4). The analogous reaction of [TiBr2(salen)] (5) gives the crystallographically characterised [Ti{(μ-Br)(AlMe2)}{(μ-Br)(AlMe2X)}(salen)] (X=Me or Br) (6) and [Ti(Me)Br(salen)] (7) in a single step, whilst the comparable reaction of [TiCl2{(3-MeO)2salen}] (8) with AlMe3 yields [Ti(Me)Cl{(3-MeO)2salen}] (9) with no evidence of titanium(III) species. Reactivity of both halide and methyl groups of 4 has been probed using magnesium reduction, SbCl5 and AgBF4 halide abstraction and SO2 insertion reactions. Hydrolysis of [Ti(Me)X(L)] complexes affords μ-oxo species [TiX(L)]2(μ-O) [X=Cl, L=salen (13); X=Br, L=salen (14); X=Cl, L=(3-MeO)2salen (15)].  相似文献   

9.
The stability of foams formed with the protein β-lactoglobulin as a function of increasing concentration of the lipid analogue -α-lysophosphatidylcholine were investigated using a microconductivity technique. The drainage, surface diffusion and thickness properties of thin liquid films (foam lamallae) were also studied using optical microscopy including epi-illumination, fluorescence recovery after photobleaching and film interferometry techniques. In addition, the surfactant binding properties of the protein were examined. The addition of small quantities of -α-lysophosphatidylcholine to β-lactoglobulin (molar ratio, R < 7:1) increased the foam stability, whereas a slightly higher concentration of surfactant in the mixture (R = 10) caused foam destabilisation. The explanation of these observations is based on changes in the composition and structure of the adsorbed interfacial layers of the thin films caused by competitive displacement of the protein by the surfactant.  相似文献   

10.
A simple and green method that uses [Ru(Me3tacn)Cl3] ( 1 ; Me3tacn=N,N′,N′′‐trimethyl‐1,4,7‐triazacyclononane) as catalyst, aqueous H2O2 as the terminal oxidant, and Al2O3 and NaCl as additives is effective in the cis‐dihydroxylation of alkenes in aqueous tert‐butanol. Unfunctionalized alkenes, including cycloalkenes, aliphatic alkenes, and styrenes (14 examples) were selectively oxidized to their corresponding cis‐diols in good to excellent yield (70–96 %) based on substrate conversions of up to 100 %. The preparation of cis‐1,2‐cycloheptanediol (119 g, 91 % yield) and cis‐1,2‐cyclooctanediol (128 g, 92 % yield) from cycloheptene and cyclooctene, respectively, on the 1‐mol scale can be achieved by scaling up the reaction without modification. Results from Hammett correlation studies on the competitive oxidation of para‐substituted styrenes (ρ=?0.97, R=0.988) and the detection of the cycloadduct [(Me3tacn)ClRuHO2(C8H14)]+ by ESI‐MS for the 1 ‐catalyzed oxidation of cyclooctene to cis‐1,2‐cyclooctanediol are similar to those of the stoichiometric oxidation of alkenes by cis‐[(Me3tacn)(CF3CO2)RuVIO2]+ through [3+2] cycloaddition (W.‐P. Yip, W.‐Y. Yu, N. Zhu, C.‐M. Che, J. Am. Chem. Soc. 2005 , 127, 14239).  相似文献   

11.
N.G. Bisset  B.C. Das  J. Parello   《Tetrahedron》1973,29(24):4137-4148
From the leaves ofStrychnos icaja Baill. a further seven alkaloids have been isolated and their structures determined: 16-hydroxystrychnine (1e), 21,22-α-epoxy-4-hydroxy-3-methoxy-N-methyl-sec.-pseudostrychnine (5c), 21,22-α-epoxy-4-hydroxy-N-methyl-sec.-pseudostrychnine5a), 21,22-α-epoxy-2-methoxy-N-methyl-sec.-pseudostrychnine (5b), 21,22-α-epoxy-14-hydroxy-N-methyl-sec.-pseudostrychnine (6a), 21,22-α-epoxy-4,14-dihydroxy-N-methyl-sec.-pseudostrychnine (6b), and 14-hydroxy-N-methyl-sec.-pseudostrychnine (7). Jaminet's alkaloid B' is shown to be impure 21,22-α-epoxy-14-hydroxy-2,3-dimethoxy-N-methyl-sec.-pseudostrychnine (6c).  相似文献   

12.
The (E)-α-selenylvinylstannanes react with acyl halides in presence of a catalytic amount of Pd(PPh3)4 to give the corresponding (Z)-α-selenyl-α,β-unsaturated ketones in good yield.  相似文献   

13.
An efficient synthesis of enantiomerically pure cis-3-substituted prolines is reported. Key steps involve the stereoselective organocuprate addition to the (E)-α,β-unsaturated ester 1, obtained from the Garner's aldehyde, and expedient oxidation–cyclization sequences.  相似文献   

14.
Immobilised Mn(salen) catalysts with two different linkages were studied in the asymmetric epoxidation of cis/trans‐β‐methylstyrene using NaClO as oxidant. The immobilised Mn(salen) complexes inside nanopores can lead to different catalytic behaviour compared with that of homogeneous Jacobsen catalyst. The rigidity of the linkage was found to be a key factor affecting the catalytic performance of immobilised catalysts. The immobilised catalyst with a rigid linkage exhibited comparable chemical selectivity, enantioselectivity and cis/trans ratio of product formation to that obtained with homogeneous Jacobsen catalysts. In contrast, the immobilised catalyst with a flexible linkage gave remarkably lower chemical selectivity, enantioselectivity and inverted cis/trans ratio compared with the results obtained with the homogeneous Jacobsen catalyst and the immobilised catalyst with rigid linkage. Thus, for immobilised Mn(salen) catalysts, a rigid linkage connecting active centres to the support is essential to obtain activity and enantioselectivity as high as those obtained in homogeneous systems.  相似文献   

15.
The acid-induced dissociation of copper(II) porphyrin complexes with a high degree of meso- and β-substitution in the ligand was studied. The copper(II) complexes with octaethylporphyrins containing 0, 1, 2, 3, and 4 meso-phenyl substituents were ranked with respect to the stability. Octaβ-alkyl and unsymmetrical meso-phenyl substitution affects not only the quantitative characteristics of dissociation of the complexes but also the kinetic relationships of the process.__________Translated from Zhurnal Obshchei Khimii, Vol. 75, No. 3, 2005, pp. 497–503.Original Russian Text Copyright © 2005 by Kosareva, Klyueva, Lomova, Suslova.  相似文献   

16.
The present study analyzes the complex of bioactive compounds from rose hips pulp powder (RHP) obtained after separating the seeds from Rosa canina L. in order to obtain the oil. The extract prepared from RHP was characterized in terms of the total content of polyphenols, flavonoids, cinnamic acids, flavonols, carotenoids, but also the content of individual polyphenols and carotenoids, antioxidant activity, and CIELab color parameters. The effects of some salts, potentially present in foods, and pH variations were examined to predict possible interactions that could occur when adding rosehip pulp as a food component. The results turned out to be a high content of polyphenols, carotenoids and antioxidant activity. The main phenolic components are procyanidin B1, chlorogenic acid, epicatechin, procyanidin B2, gallic acid, salicylic acid, and catechin. The carotenoid complex includes all-trans-β-carotene, all-trans-lycopene, zeaxanthin, α-cryptoxanthin, β-cryptoxanthin, rubixanthin, cis-β-carotene, cis-γ-carotene and cis-lycopene. The addition of CaCl2 and NaCl to the RHP extract reduced the antioxidant activity and the strong acidic environment (pH to 2.5) decreased the antioxidant activity by 29%. The addition of rose hip powder to gingerbread has improved its general characteristics, and increased its antioxidant activity and microbiological stability, the effects of 4% RHP being the most important.  相似文献   

17.
Chlorodiphenylphosphine and 2,2′-biphenylylenephosphorochloridite react with 2-hydroxy-2′-(1,4-bisoxo-6-hexanol)-1,1′-biphenyl to yield the new α,ω-bis(phosphorus-donor)-polyether ligands, 2-Ph2PO(CH2CH2O)2–C12H8-2′-OPPh2 (1) and 2-(2,2′-O2C12H8)P(CH2CH2O)2–C12H8-2′-P(2,2′-O2C12H8) (2). These ligands react with Mo(CO)4(nbd) to form the monomeric metallacrown ethers, cis-Mo(CO)4{2-Ph2PO(CH2CH2O)2–C12H8-2′-OPPh2} (cis-3) and cis-Mo(CO)4{2-(2,2′-O2C12H8)P(CH2CH2O)2–C12H8-2′-P(2,2′-O2C12H8)} (cis-4), in good yields. The X-ray crystal structures of cis-3 and cis-4 are significantly different, especially in the conformation of the metal center and the adjacent ethylene group. The very different 13C-NMR coordination chemical shifts of this ethylene group in cis-3 and cis-4 suggest that the solution conformations of these metallacrown ethers are also quite different. Both metallacrown ethers undergo cistrans isomerization in the presence of HgCl2. Although the cistrans equilibrium constants for the isomerization reactions are nearly identical, the isomerization of cis-3 is more rapid. Phenyl lithium reacts with cis-3 to form the corresponding benzoyl complexes but does not react with either trans-3 or cis-4. Both the slower rate of cistrans isomerization of cis-4 and its lack of reaction with PhLi are consistent with weaker interactions between the hard metal cations and the carbonyl oxygens in both trans-3 and cis-4.  相似文献   

18.
A series of heterocyclic trans-dichloro-β-diketonato-cis-diorganoantimony(V) compounds of the type R2SbCl2X (R2 = (CH2)4, (CH2)5, o,o′−C6H4C6H4, o,o′−C6H4CH2C6H4; X = Acac, Dpm) has been synthesized. The stereochemistry of these compounds has been deduced from PMR spectroscopic and molecular dipole moment data. Since the cis-dichloro-β-diketonato-trans-diorganoantimony(V) compounds R2SbCl2Acac (R = Me, Et, Ph) were known previously, a set of both cis- and trans-diorgano main group organometallic complexes has thus been made available, which allows a comparative study of the influence of stereochemistry on the strength of metal—ligand interactions in this type of octahedral d10 metal complex. β-Diketonate—ligand exchange reactions have been studied by PMR spectroscopy, and a marked influence of stereochemistry observed. trans-Dichloro-β-diketonato-cis-diorganoantimony(V) compounds undergo ligand exchange only slowly, if at all, whereas cis-dichloro-β-diketonato-trans-diorganoantimony(V) compounds react instantaneously. Both PMR chemical shift data and IR spectroscopic data point to the occurrence of a stronger antimony-β-diketonate interaction in trans-dichloro-β-diketonato-cis-diorganoantimony than in cis-dichloro-β-diketonato-trans-diorganoantimony compounds. This can be understood in terms of the hybridization of the antimony valence orbitals. The results are in line with the assumption that Sb---O bond rupture is the rate-determining step in β-diketonate ligand exchange.  相似文献   

19.
Summary: A binaphthyl‐bridged salen dichlorozirconium (IV ) complex that displays an octahedral structure with a trans‐O, cis‐N, and cis‐Cl arrangement was synthesized and tested as a precatalyst for ethylene and α‐olefin polymerization. While use of methylaluminoxane (MAO) cocatalyst afforded poor catalytic activity, activation by mixtures of aluminium alkyls such as AliBu3 and either MAO or [CPh3][B(C6F5)4] resulted in reasonable polymerization activities for ethylene, propene, and higher α‐olefins. Quite unexpectedly, while the polymerization of propene results in the production of a high‐molecular‐weight stereoirregular polymer, highly isotactic polymers are obtained under similar conditions from polymerization of 1‐butene, 1‐pentene, and 1‐hexene.

Polymerization employing the binaphthyl‐bridged salen dichlorozirconium (IV ) complex gave unexpected different stereospecificities for the polymerization of propene and higher α‐olefins, to yield ultrahigh‐molecular‐weight atactic poly(propylene) and highly isotactic polymers, respectively.  相似文献   


20.
The dielectric properties of mixed monolayers of per-(6-amino-2,3-di-O-hexyl) β-CD hydrochloride (NH3-β-CD-OC6) and 1,2 dipalmitoyl, 3-sn-phosphatidic acid (DPPA) have been assessed using surface potential measurements at constant area. From the comparison of these surface potential (ΔV) versus surface density (δ) relationships with those of surface pressure (π) against surface density (δ) it was apparent that the increase in the NH3-β-CD-OC6 content in mixed films gave rise to a gradual increase in the saturation value of the surface potential (ΔVmax). This potential for pure DPPA was found to be equal to 396 mV and for pure CD 554 mV. The ΔVmaxvalues reflect the onset of reorientation effects that arrive at molar areas before the collapse of these films. Independently of reorientation effects, the obtained results strongly indicate that the dipolar term contributing to the overall ΔVvalue was for NH3-β-CD-OC6 due to the hydration of its NH+3group. For both DPPA and NH3-β-CD-OC6 molecules the contribution of the electric double layer (Ψ) was calculated and was found for DPPA and NH3-β-CD-OC6 to be equal to −249 and +252 mV, respectively. These calculated Ψ values made possible the evaluation of dipole moments for NH3-β-CD-OC6 and DPPA monolayers which revealed a marked difference in dipolar properties between these two film forming components. In contrast to DPPA which exhibited a decrease in the surface dipole moment (μ) with the decrease inA, NH3-β-CD-OC6 displayed an increase in μwith the decrease inAforAvalues above 580 Å2. Below this value μdecreases with decreasing molecular area and this variation arises from a change in the polarity of the electric double layer arising from interactions with the complementary anion. The differences in dielectric properties between the two film forming molecules have been attributed to modification, during compression, in the structure of the interfacial water bound to the cyclodextrin.  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