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 914 毫秒
1.
Many natural and synthetic organic amphiphiles have been discovered, prepared, and/or characterized. Aggregation behavior has been studied for many of these monomers. Very few of these many amphiphiles are organometallic compounds and almost nothing is known about this novel group of amphiphiles. Under appropriate circumstances, aggregation of certain organometallic amphiphile monomers can be controlled by use of redox switching. Compounds which are not amphiphilic can be made so and amphiphilic compounds can be deprived of this property by appropriate alterations in the metal ion's oxidation state. Redox-switched aggregation-deaggregation behavior is described for bis(hexadecycloxy)-1,10-phenanthrolinium perchlorate (2), its nickel complex (3), [CH3(CH2)15NH2]2Ag+AcO (4) and bis(hexadecylamine)copper dichloride (5), chloride acetate (6) and diacetate (7), all of the general formula: [CH3(CH2)15NH2]2Cu2+X22. Two completely new switching mechanisms are disclosed for the first time which involve electron transfer from dilute aqueous sugar solutions or from solid zinc metal which are effective in collapsing [CH3(CH2)15NH2]2Cu2+X22− vesicles.  相似文献   

2.
The complex Ir(CH3) (CO) (CF3SO3)2 (dppe) (1) has been synthesized from the reaction of Ir(CH3)I2(CO) (dppe) and silver triflate. Methane and IrH(CO) (CF3SO3)2 (dppe) (2) are formed when a methylene chloride solution of 1 is placed under 760 torr dihydrogen. Conductivity studies indicate that methylene chloride solutions of complexes 1 and 2 are weak electrolytes and only partially ionized at concentrations above 1 mM. Complex 2 is an effective hydrogenation catalyst for ethylene and 1-hexene while acetone hydrogenation is inhibited by the formation of [IrH2(HOCH(CH3)2) (CO) (dppe)] (OTf) (3). Linear dimerization and polymerization of styrene occurs via a carbocationic mechanism initiated by triflic acid elimination from 2. Treatment of an acetonitrile solution of Ir(CH3)I2(CO) (dppe) with silver hexafluorophosphate produces the solvent promoted carbonyl insertion product [Ir(C(O)CH3) (NCCH3)3 (dppe)] [PF6]2 (7) which readily undergoes deinsertion in methylene chloride to form [Ir(CH3) (CO) (NCCH3)2 (dppe)] [PF6]2 (8) and acetonitrile.  相似文献   

3.
The square-planar bis chelate complexes Ni(R-sal)2 (= bis(N-alkyl)salicylaldiminato)nickel(II)) with R = (CH2)2Ph (I; Ph = phenyl), (CH2)3Ph (II), (CH2)4Ph (III) and (CH2)2(4-hydroxyphenyl) (IV) were prepared and characterized. ComplexesII and III meet the steric requirements for intramolecular aromatic ring stacking. Stopped-flow spectrophotometry was used to study the kinetics of ligand substitution in complexesI–IV by H2salen (=N,N′-disalicylidene-ethylenediamine) in acetone. For the substitution of the two bidentate ligands in Ni(R-sal)2 only one step is kinetically observed which follows a second-order rate law, rate =k[H2salen] [Ni(R-sal)2], with k = 43.4 (I), 64.0 (II), 87.0 (III) and 49.5 (IV) M−1 s−1 at 298 K. It is found, therefore, that the size of k does not change significantly upon lengthening of the alkane chain in Ni(Ph(CH2)nsal)2 from n = 2 to 4 and that there is no kinetic evidence for intramolecular stacking interactions. The equilibrium constants and thermodynamic parameters for the formation of the bis adductsIII·(py)2 and III·(MeOH)2 in acetone are reported.  相似文献   

4.
The dicarbonylation reaction of E-β-deuteriostyrene to syndiotactic poly(1-oxo-2-phenyltrimethylene) as well as to dimethyl-2-phenylbutanedioate and dimethyl-2,5-diphenyl-4-oxoheptanedioate using Pd(CF3COO)2/2,2′-bipyridine as the catalyst precursor in the presence of 1,4-benzoquinone in methanol takes place stereospecifically in a syn-fashion with complete retention of the label. The same result was found for the dicarbonylation to dimethyl 2-phenylbutanedioate catalyzed by [Pd(CF3COO)2(Diop)]. In the absence of the oxidant the latter catalytic system produces methyl 2- and 3-phenylpropionates for which some scrambling of deuterium is observed when using either -deuteriostyrene or CH3OD as the labeled substrate. [Pd(CH3CN)4][BF4]2 modified with different ligands catalyses the formation of E-1,5-diphenylpent-1-en-3-one or of E-1,4-diphenylpent-1-en-3-one in tetrahydrofuran as the solvent. The label distribution using E-β-deuteriostyrene as the substrate (or styrene in the presence of dideuterium) suggests that in the synthesis of ketones catalyzed by [Pd(p-CH3C6H4SO3)2(Dppp)]·2H2O the regioselectivity of the first inserted olefin unit does not determine the ketone regioisomer; rather which regioisomeric product preferentially forms depends on the rate of carbon monoxide insertion in either the branched or linear metal-hydrocarbyl intermediate. β-Hydrogen elimination is very rapid both after the first and the second olefin insertion.  相似文献   

5.
Four 9,10-anthraquinones (AQ) mono- or bis-substituted with the -NH(CH2)2 NH(CH2)2OH group were studied. 1-AQ, 1,5-AQ and 1,8-AQ but not 1,4-AQ (100°M) generated pBR322 plasmid DNA single strand breaks in the presence of purified NADPH dependent cytochrome P450 reductase. 1-AQ, 1,5-AQ and 1,8-AQ (at 100 °M) stimulated hydroxyl radical formation in MCF-7 S9 cell fraction (as measured by dimethyl pyrolline N-oxide spin trapping) and MCF-7 DNA strand breaks as measured by alkaline filter elution. In contrast 1,4-AQ did not stimulate hydroxyl radical formation and produced considerably less strand breaks in MCF-7 cells compared to the other AQ's. It would appear that the position of the -NH(CH2)2 NH(CH2)2OH groups on the chromophore is an important determinant in the metabolic activation of cytotoxic anthraquinones. This may contribute to the cytotoxicity (ID50 values) of 1-AQ (0.06 °M), 1-8-AQ (0.5 °M) and 1,5-AQ (12.3 °M) but not the 1,4-AQ (1.2 °M).  相似文献   

6.
The ruthenium(III) complex [(Cp*)RuCl2]2 (Cp*=permethylcyclopentadienyl) catalyzes polymerization of propiolic acid to give a mixture of poly(propiolic acid), [---CH=C(COOH)---]n (1), and cyclic trimers, 1,2,4- and 1,3,5- benzenetricarboxylic acids. GPC analysis shows MN and MW values of the polymer of 4.0 × 103 and 4.3 × 103, respectively. Reaction of propiolic acid in the presence of the Ru(II) complex, (Cp*)RuCI(L) (L=1,5-cyclooctadiene and norbornadiene), gives the cyclic trimers rather than 1. [(Cp*)RuCl2]2 catalyzes polymerization of acetylenedicarboxylic acid and of propargyl alcohol to give the corresponding poly(acetylene) derivatives, [---C(COOH)=C(COOH)---]n (2) and [---CH=C(CH2OH)---]n (3), respectively. Polymerization of ethyl propiolate, 2-butyn-1,4-diol, phenylacetylene and (trimethylsilyl)acetylene using [(Cp*)RuCl2]2 gives the corresponding polymers [---CH=C(COOEt)---]n (4), [---C(CH2OH)=C(CH2OH)---]n (5), [---CH=CPh---]n (6) and [---CH=C(SiMe3)---]n (7) in low yields.  相似文献   

7.
2-(Diethylphosphonate)-nitrosopropane (DEPNP), prepared by oxidation of the corresponding aminophosphonate, was found to essentially exist as monomer in both water and organic solvents. The mechanisms of its degradation under 80°C heating or visible light exposure were studied by EPR spectroscopy: its decomposition gave rise to paramagnetic by-products, which have been identified as DEPNP / ·C(CH3)2[P(O)(OC2H5)2] and DEPNP / ·P(O)(OC2H5)2 spin adducts. Despite this drawback, DEPNP was successfully used as spin trapping agents to scavenge various carbon — and phosphorus-centred free radicals both in aqueous and organic media, giving rise to intense EPR spectra characteristic of the species trapped.  相似文献   

8.
A series of borane and monoiodoborane derivatives of bis(diphenylphosphino)alkanes. (C6H5)2P--- (CH2)n---P(C6H5)2 in which n has values of 2 through 4 has been synthesized. Only compounds with the formulae [(C6H5)2P]2(CH2)n · (BH3)2 and (C6H5)2P]2CH2)n · BH2I were isolable, the latter being boronium iodides. The compounds were characterized by their melting points, elemental analyses, molar conductivities, infrared spectroscopy, and 1H and 11B nuclear magnetic resonance spectroscopy. The relationship between the length of the carbon chain and the 11B NMR chemical shift is discussed.  相似文献   

9.
Addition of (Cp*2YH)2 (4) to 2-methyl-1,4-pentadiene produced the yttrium-alkyl-alkene chelate complex Cp*2YCH2CH2CH2C(CH3)=CH2 (2) in which a disubstituted alkene is complexed to the metal center. Evidence for coordination of the alkene unit of 2 comes from the 1H and 13C NMR chemical shifts of the vinyl units and from observation of nOe effects between Cp* protons and vinyl hydrogens. The disubstituted alkene ligand of 2 is weakly bound, and evidence for an equilibrium with substantial amounts of complex 3 with a free alkene was obtained from variable temperature 1H NMR spectroscopy.  相似文献   

10.
An efficient one-pot catalytic method to obtain 4,6-dimethyl-2-hydroxyacetophenone (A) is reported, the reaction proceeds via the intermolecular auto-condensation of 2,4-pentanedione using samarium(III) acetylacetonate (Sm(AcAc)3) as promoter. A novel complex [Sm(CH3COO)3(H2O)2](H2O)2 (I) was isolated from the reaction media. The structure of I was determined by X-ray crystallography showing that the central atom is ennea-coordinated (monocapped square-antiprism geometry). This complex I also shows activity in the named autocondensation reaction.  相似文献   

11.
The photochemistry of the diphosphino Pt(II) hydrides [LPtH2] (L=(t-Bu)2P(CH2)2P(t-Bu)2 (7); L=(t-Bu)2P(CH2)3P(t-Bu)2 (8);L=(t-Bu)(Ph)P(CH2)2P(Ph)(t-Bu) (9)) is reported. The primary photoevent is the dissociation of H2 and formation of the 14-e [LPt] species. These coordinatively unsaturated intermediates provide a versatile entry point into the C---H bond activation of hydrocarbons. [LPt] reacts with benzene in an oxidative addition reaction to yield [LPt(H)(C6H5)] complexes. The importance of the metal centre and ancillary ligation in the C---H bond activation is discussed.  相似文献   

12.
[Fe(TIM)(CH3CN)2](PF6)2 (1) (TIM = 2,3,9,10-tetramethyl-1,4,8,11-tetraazacyclodeca-1,3,8,10-tetraene) forms a complex with NO reversibly in CH3CN (53±1% converted to the NO complex) or 60% CH3OH/40% CH3CN (81±1% conversion). Quantitative NO complexation occurs in H2O or CH3OH solvents. The EPR spectrum of [Fe(TIM)(solvent)NO]2+ in frozen 60/40 CH3OH/CH3CN at 77 K shows a three line feature at g=2.01, 1.99 and 1.97 of an S=1/2FeNO7 ground state. The middle line exhibits a three-line N-shf coupling of 24 G indicating a six-coordinate complex with either CH3OH or CH3CN as a ligand trans to NO. In H2O [Fe(TIM)(H2O)2]2+ undergoes a slow decomposition, liberating 2,3-butanedione, as detected by 1H NMR in D2O, unless a π-acceptor axial ligand, L=CO, CH3CN or NO is present. An equilibrium of 1 in water containing CH3CN forms [Fe(TIM)(CH3CN)(H2O)]2+ which has a formation constant KCH3CN=320 M−1. In water KNOKCH3CN since NO completely displaces CH3CN. [Fe(TIM)(CH3CN)2]2+ binds either CO or NO in CH3CN with KNO/KCO=0.46, sigificantly lower than the ratio for [FeII(hemes)] of 1100 in various media. A steric influence due to bumping of β-CH2 protons of the TIM macrocycle with a bent S=1/2 nitrosyl as opposed to much lessened steric factors for the linear Fe---CO unit is proposed to explain the lower KNO/KCO ratio for the [Fe(TIM)(CH3CN)]2+ adducts of NO or CO. Estimates for formation constants with [Fe(TIM)]2+ in CH3CN of KNO=80.1 M−1 and KCO=173 M are much lower than to hemoglobin (where KNO=2.5×1010 M−1 and KCO=2.3×107) due to a reversal of steric factors and stronger π-backdonation from [FeII(heme)] than from [FeII(TIM)(CH3CN)]2+.  相似文献   

13.
Affinity probes for the noncompetitive blocker or picrotoxinin site of the γ-aminobutyric acid (GABA)-gated chloride channel were designed for four types of applications: photoaffinity reagents to covalently label the binding site; fluorescent probes for receptor analysis; biotinylated compounds and agarose/sepharose conjugates for affinity chromatography; ligand-protein/enzyme conjugates for immunoassay. These 5e-tert-butyl-2e-[4-(substituted-ethynyl)phenyl]-1,3-dithianes were optimized by structure-activity studies for potency as inhibitors of 3H ethynylbicycloorthobenzoate binding to bovine brain membranes, measured as the concentration for 50% inhibition (IC50). Preferred compounds are 5e-(CH3)3CCH(CH2S)2CH-2e-C6H4-4-CCCH2OCH2C(O)R, wherein R confers the following properties and 1C50 values: R = SCH2CH2SCH2C(O)C6H4-4-N3, photo-affinity, 9 nM; R = NHCH2CH2NHC(O)C6H2-2-OH,5-1,4-N3, photoaffinity, 105 nM; R = SCH2CH2S-4-benzofurazan-7-NO2, fluorescent, 13 nM; R = SCH2CH2SCH2-5-fluorescein, fluorescent, 27 nM; R = NHCH2CH2NH[C(O)(CH2)5NH]2-biotin, affinity chromatography, 190 nM. The most potent photoaffinity ligand (IC50 9 nM) was labeled at 7 Ci mmol−1 by reacting the appropriate thiol with 3H 4-azidophenacyl bromide (obtained by alumina-catalyzed tritium exchange of its enolizable hydrogens). The first steps have been taken in using the NCB site for affinity chromatography of the GABAA receptor in CHAPS-solubilized bovine brain membranes with the dithiane-biotin probe and an avidin-acrylic bead system or with an analogous dithiane-agarose/sepharose column eluting with GABA or dithiane as above (R = OH). A protein conjugate of a related dithiane-monosulfone elicited production of specific antisera in rabbits. These findings illustrate the diversity and utility of new affinity probes prepared in the alkynylphenyldithiane series.  相似文献   

14.
A series of zirconium(IV) complexes, [ZrX2(XDK)], where XDK is the constrained carboxylate ligand m-xylylenediamine bis(Kemp's triacid imide), were prepared and structurally characterized. The solid state structure of the mononuclear carboxylate alkyl complex [Zr(CH2Ph)2(XDK)] reveals that one benzyl group is bonded in an η2-fashion to the metal center. The reactivity of [Zr(CH2Ph)2(XDK)] displays its electrophilic character toward nucleophiles strong enough to displace the η2-benzyl group. Thus, weak sigma donor ligands such as CO, alkynes and anilines do not react, whereas strong sigma donors, such as pyridines and isocyanides, rapidly form the monoadduct [Zr(CH2Ph)2(4-tert-butylpyridine)(XDK)] and [Zr{η2-2,6-Me2PhNCCH2Ph}2(XDK)], an η2-iminoacyl derivative, respectively. Attempts to prepare zirconium amido complexes with H2XDK generally afforded the eight-coordinate [Zr(XDK)2] complex but use of the small amido ligand precursorZr(NMe2)4 allowed [Zr(NMe2)2(4-tert-butylpyridine)(XDK)] to be isolated in good yield.  相似文献   

15.
The chloro complexes trans-[Pt(Me)(Cl)(PPh3)2], after treatment with AgBF4, react with 1-alkynes HC---C---R in the presence of NEt3 to afford the corresponding acetylide derivatives trans-[Pt(Me) (C---C---R) (PPh3)2] (R = p-tolyl (1), Ph (2), C(CH3)3 (3)). These complexes, with the exception of the t-butylacetylide complex, react with the chloroalcohols HO(CH2)nCl (n = 2, 3) in the presence of 1 equiv. of HBF4 to afford the alkyl(chloroalkoxy)carbene complexes trans-[Pt(Me) {C[O(CH2)nCl](CH2R) } (PPh3)2][BF4] (R = p-tolyl, N = 2 (4), N = 3 (5); R=Ph, N = 2 (6)). A similar reaction of the bis(acetylide) complex trans-[Pt(C---C---Ph)2(PMe2Ph)2] with 2 equiv. HBF4 and 3-chloro-1-propanol affords trans-[Pt(C---CPh) {C(OCH2CH2CH2Cl)(CH2Ph) } (PMe2Ph)2][BF4] (7). T alkyl(chloroalkoxy)-carbene complex trans-[Pt(Me) {C(OCH2CH2Cl)(CH2Ph) } (PPh3)2][BF4] (8) is formed by reaction of trans-[Pt(Me)(Cl)(PPh3)2], after treatment with AgBF4 in HOCH2CH2Cl, with phenylacetylene in the presence of 1 equiv. of n-BuLi. The reaction of the dimer [Pt(Cl)(μ-Cl)(PMe2Ph)]2 with p-tolylacetylene and 3-chloro-1-propanol yields cis-[PtCl2{C(OCH2CH2CH2Cl)(CH2C6H4-p-Me}(PMe2Ph)] (9). The X-ray molecular structure of (8) has been determined. It crystallizes in the orthorhombic system, space group Pna21, with a = 11.785(2), B = 29.418(4), C = 15.409(3) Å, V = 4889(1) Å3 and Z = 4. The carbene ligand is perpendicular to the Pt(II) coordination plane; the PtC(carbene) bond distance is 2.01(1) Å and the short C(carbene)-O bond distance of 1.30(1) Å suggests extensive electronic delocalization within the Pt---C(carbene)---O moietry.  相似文献   

16.
The ready substitution of coordinated trifluoromethanesulfonate on pentaamminechromium(III) has been applied to the facile synthesis of a range of complexes of neutral ligands, [Cr(NH3)5(L)]3+ (L = OH2, OHCH3, OS(CH3)2, OP(OCH3)3, OC(NH2)2, OC(NHCH3)2, OC(CH3) · N(CH3)2, OCH · NH2, OCH · N(CH3)2, NCCH3, NH3 and imidazole). The complexes have been characterized by microanalysis, electronic and infrared spectroscopy, and the lability of the neutral ligand towards acid hydrolysis determined, and compared with cobalt(III) analogues.  相似文献   

17.
Lipase catalysis was successfully employed to synthesize high molecular weight poly(butylene succinate) (PBS). Attempts to copolymerize succinic acid with 1,4-butanediol were unsuccessful due to phase separation of the reactants. To circumvent this problem, monophasic reaction mixtures were prepared from diethyl succinate and 1,4-butanediol. The reactions were studied in bulk as well as in solution. Of the organic solvents evaluated, diphenyl ether was preferred, giving higher molecular weight products. After 24 h in diphenyl ether, polymerizations at 60, 70, 80, and 90 degrees C yielded PBS with M(n) of 2000, 4000, 8000, and 7000, respectively. Further increase in reaction time to 72 h resulted in little or no further increase in M(n). However, increasing the reaction time produced PBS with extraordinarily low M(w)/M(n) due to the diffusion and reaction between low-molecular weight oligomers and chains that occurs at a greater frequency than interchain transesterification. Time-course studies and visual observation of polymerizations at 80 degrees C revealed PBS precipitates at 5 to 10 h, limiting the growth of chains. To maintain a monophasic reaction mixture, the polymerization temperature was increased from 80 to 95 degrees C after 21 h. The result was an increase in the PBS molecular weight to M(w) = 38 000 (M(w)/M(n) = 1.39). This work paves the way for the synthesis of PBS macromers and polymers that contain variable quantities of monomers with chemically sensitive moieties (e.g., silicone, epoxy, vinyl). Furthermore, this study established the feasibility of using lipase catalysis to prepare polyesters from alpha,omega-linear aliphatic diethyl ester/diol monomers with less than six carbons.  相似文献   

18.
Binding characteristics of the selective V2 antagonist radioligand [3H]desGly-NH29-d(CH2)5[D-Ileu2,Ileu4]AVP to rat kidney were determined. Binding was specific, saturable and reversible. The peptide bound to a single class of high-affinity binding sites with Bmax 69.4±6.8 fmol/mg protein and KD 2.8±0.3 nM. AVP and other related peptides displaced [3H]desGly-NH29-d(CH2)5[D-Ileu2,Ileu4]AVP binding. The order of potency of inhibition was desamino-D-AVP > AVP > d(CH2)5[D-Ileu2,Ileu4]AVP > oxytocin > d(CH2)3[Tyr(Me)2]AVP > d(CH2)5[sarcosine7]AVP, which is typical of a selective V2 radioligand. Autoradiographic localization of [3H]desGly-NH29-d(CH2)5[D-Ileu2,Ileu4]AVP binding sites in kidney showed dense binding in the inner and outer medulla with less binding in the cortex, which is consistent with known renal V2 receptor distribution.  相似文献   

19.
Lithiation of [p-But-calix[4]-(OMe)2(OH)2] (1), followed by reaction with TiCl3(thf)3 or TiCl4(thf)2, led to the corresponding titanium-calix[4]arene complexes [p-But-calix[4]-(OMe)2(O)2]TiCl] (2) and [p-But-calix[4]-(OMe)2(O)2]TiCl2] (3), respectively. Reaction of 1 with TiCl4(thf)2 results in demethylation of the calix[4]arene and the obtention of [p-But-calix[4]-(OMe)2(O)3]TiCl] (4), whose hydrolysis led to [p-But-calix[4]-(OMe)(OH)3] (6). The preparation of 6 can be carried out as a one-pot synthesis. Both 2 and 4 undergo alkylation reactions using conventional procedures, thus forming surprisingly stable organometallic species, namely [p-But-calix[4]-(OMe)2(O)2Ti(R)] (R = Me (7); CH2Ph (8), p-MeC6H4 (9) and [p-But-calix[4]-(OMe)(O)3Ti(R)] (R = Me (10); CH2Ph (11); p-MeC6H4 (12)). Complexes 7 and 9 undergo a thermal oxidative conversion into 10 and 12, occurring with the demethylation of one of the methoxy groups. A solid state structural property of 9 and 12 has been revealed by X-ray analysis showing a self-assembly of the monomeric units into a columnar polymer, where the p-tolyl substituent at the metal functions as a guest group for an adjacent titanium-calixarene. Reductive alkylation of 3 with Mg(CH2Ph)2 gave 8 instead of forming the corresponding dialkyl derivative. Two synthetic routes have been devised for the synthesis of the Ti(III)-Ti(III) dimer [p-But-calix[4]-(OMe)(O)3Ti]2] (13): the reduction of 4 and the reaction of TiCl3(thf)3 with the lithiated form of 6. A very strong antiferromagnetic coupling is responsible for the peculiar magnetic behavior of 13. The proposed structures have been supported by the X-ray analyses of 4, 9, 12 and 13.  相似文献   

20.
Ionic liquids are now recognized as solvents for use in lipase-catalyzed reactions; however, there still remains a serious drawback in that the rate of reaction in an ionic liquid is slower than that in a conventional organic solvent. To overcome this problem, attempts have been made to evolve phosphonium ionic liquids appropriate for lipase-catalyzed reaction; several types of phosphonium salts have been prepared and their capability evaluated for use as solvent for the lipase-catalyzed reaction. Very rapid lipase PS-catalyzed transesterification of secondary alcohols was obtained when 2-methoxyethyl(tri-n-butyl)phosphonium bis(trifluoromethanesulfonyl)imide ([MEBu3P][NTf2]) was used as solvent, affording the first example of a reaction rate superior to that in diisopropyl ether.  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