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.
The first preparation of enantiomerically pure 1,1-binaphthyl-2,2-diboronic acid (by resolution) is reported. Optimization of the cross-coupling conditions was found to be crucial to achieve good yields in Suzuki arylation in approaches from both 2,2-diiodide or 2,2-diboronic acid (52-56%, with model p-tolyl reagents). Stereochemical results in these reactions were dramatically different: almost complete racemization starting from the 2,2-diiodide versus complete conservation of stereogenic information from the 2,2-diboronic acid. This novel synthetic approach, a stereoconservative Suzuki arylation of the diboronic acid, should be a valuable method for the synthesis of a new group of 2,2-diarylated (including functionalized) binaphthyl derivatives.  相似文献   

2.
The intercalation process of 2,2′-bipyridine into layered MnPS3 is studied with powder X-ray powder diffraction (XRD) technology as the monitoring tool. From the XRD results, it is found that the absence or presence of acid greatly influences the existing form and the arranged orientation of the guest. Two series of the reactions are carried out. In Series A, only MnPS3 and 2,2′-bipyridine are present. While in Series B, a variety of acetic acid is added. During the intercalation of Series A, there coexist four phases: the 00l phase (with lattice spacing of 6.47 Å) is pristine MnPS3; the 00l′ phase (with lattice spacing of 9.81 Å), indicating the parallel orientation of the 2,2′-bipyridine molecular ring to the layer; the 00l″ phase (with lattice spacing of 12.20 Å), indicating the perpendicular orientation of the 2,2′-bipyridine molecular ring to the layer of the host, which is only an intermediate phase for the formation of the 00l′′′ phase; the 00l′′′ phase (with the lattice spacing of 15.33 Å), indicating the existence of the complex cation [Mn(bipy)3]2+ coming from the in situ coordination of the inserted guest with intralayered Mn2+ ions between the interlayer space of host. As the intercalation proceeds, the 00l, 00l′ and 00l″ phases finally disappear, and 00l′′′ phase is intensified and a complete intercalate is obtained. In Series B, due to the presence of the acid, the formation of the complex cation [Mn(bipy)3]2+ is inhibited, and the amount of the acid in the intercalation plays a key role in the formation of the guest. With the increase of the acid, the protonated bipyridine becomes the main existing form of the guest, which is arranged in the perpendicular orientation of molecular ring to the layer. From the experimental evidences, the possible intercalation mechanisms are proposed and the novel intercalation phenomenon of in situ coordination of the inserted 2,2′-bipyridine with Mn2+ of the host is elucidated.  相似文献   

3.
A series of cobalt(II) complexes having terpyridine derivatives such as 2,2:6,2″-terpyridine (1), 4,4,4″-tBu3-2,2:6,2″-terpyridine (2), 5,5″-Me2-2,2:6,2″-terpyridine (3), 6,6″-Me2-2,2:6,2″-terpyridine (4) and 6,6″-(3,5-Me2C6H3)2-2,2:6,2″-terpyridine (5) was synthesized. The structures of 1, 3, and 4 were confirmed by X-ray crystallography. The coordination sphere around the cobalt center in 1 can be described as pseudo square pyramidal. On the other hand, complex 4 has pseudo trigonal bipyramidal structure. Upon activation with d-MAO (dried-methylaluminoxane), these complexes showed high activities for the polymerization of norbornene (NBE). In particular, polymerization of NBE with 4/d-MAO system at room temperature resulted in quantitative yield within several hours to give the polymers with relatively narrow molecular weight distributions and controlled molecular weight. The polymerizations of NBE with these cobalt catalyst systems proceeded in vinyl addition polymerization, which was confirmed by 1H NMR spectra of the resulting polymers.  相似文献   

4.
The coordination of two 5-substituted-2,2-bipyridines L (L1=5-methyl-2,2-bipyridine, L2=5,5-dimethyl-2,2-bipyridine) to palladium was studied. The neutral complexes [Pd(L)Cl2] and [Pd(L)(Me)Cl], and the cationic complexes obtained after chlorine abstraction [Pd(L)2][BAr4]2 and [Pd(L)(Me)(NCMe)][BAr4] (Ar=3,5-(CF3)2-C6H3), respectively, were isolated and characterized by NMR and FAB mass spectroscopy. The complex [Pd(L2)(L3)][BAr4]2 (L3=2,2-bipyridine) bearing different ligands, was prepared for comparison purposes. The activity of the monocationic and dicationic complexes as catalytic precursors in the CO/4-tert-butylstyrene copolymerization was compared with that of related well-known catalysts containing the unsubstituted 2,2-bipyridine as nitrogen ligand, to evaluate the influence of the substituents in 5- and 5,5-position. The presence of one or two substituents on the nitrogen ligand has a positive effect on productivity using both types of precursors. No influence was observed on the polymer properties in terms of molecular weight and tacticity. Analysis of the reactivity of the methyl-palladium complexes towards carbon monoxide shows further differences depending on the nitrogen ligand.  相似文献   

5.
Atom transfer radical polymerization conditions with copper(I) bromide/2,2-bipyridine (Cu/2,2-bpy) as the catalyst system were employed for the homopolymerization and random copolymerization of 1-phenoxycarbonyl ethyl methacrylate (PCMA) with methyl methacrylate (MMA). Temperature studies indicated that the polymerizations occurred smoothly in bulk at 110 °C. Poly(PCMA)(polydispersity index=1.27) homopolymer was characterized and then used as macroinitiator for increasing its molecular weight. The homopolymerization of PCMA was also carried out under free radical conditions using 2,2-azobisisobutyronitrile as an initiator.The monomer and polymers were characterized by FT-IR and 1H and 13C-NMR techniques. The glass transition temperatures, the solubility parameters and average-molecular weights of the polymers were determined. Thermal stabilities of the polymers were given as compared with each other by using TGA curves. Thermal degradation products of poly(PCMA)s obtained by ATRP and free radical polymerization were compared with each other by using 1H-NMR technique.  相似文献   

6.
Dilithiation of optically active 2,2-dibromo-1,1-binaphthyl 2 with t-BuLi followed by carboxylation of the resulting dilithio-intermediate 3 with CO2 gave optically active 1,1-binaphthyl-2,2-dicarboxylic acid 1, which was further transformed to its dicyano derivative 4. Both of these transformations were carried out in a one-pot operation and the products were obtained in excellent yields with no observable racemization.  相似文献   

7.
A series of hyperbranched copolyimides (HBPI)s based on commercially available monomers 4,4′-oxydiphthalic anhydride (ODPA), 2,4,6-triaminopyrimidine (TAP) and 4,4′-oxydianiline (ODA) were prepared. The synthesis involved the formation of hyperbranched polyamic acid (PAA) precursors in the first step and the thermal imidization of cast thin PAA films in the second step. Two basic types of HBPIs were prepared by controlling the molar ratio of ODPA and an amine mixture of TAP and ODA. When the molar ratio was 1:1, the amine-terminated HBPIs were obtained; with the molar ratio of 2:1 anhydride-terminated HBPIs were prepared. Degree of branching was estimated by 1H and 13C NMR analysis. It was found that approximately 48% of TAP units presented in ODPA:TAP:ODA = 1:0.75:0.25 HBPI macromolecules create the branching unit. Amine-terminated HBPIs showed moderate weight-average molecular weights and these values rather higher than for the anhydride-terminated HBPIs. With increasing ODA comonomer content in amine-terminated HBPIs increased their molecular weight and thermal and mechanical stability, whereas in anhydride-terminated HBPIs these trends were opposite. Amine-terminated HBPIs generally exhibited higher thermal stability than the anhydride-terminated ones. Gas permeability coefficients of both HBPIs types increased with increasing content of ODA comonomer. Prepared membranes exhibited high separation performance and have a potential to be utilized in industrial gas separation applications.  相似文献   

8.
Two new diacid monomers, 2,2′-sulfide bis(4-methyl phenoxy acetic acid) and 2,2′-sulfoxide bis(4-methyl phenoxy acetic acid) were successfully synthesized by refluxing the 2,2′-sulfide bis(4-methyl phenol) and 2,2′-sulfoxide bis(4-methyl phenol) with chloroacetonitrile in the presence of potassium carbonate, and subsequent basic reduction. Two novel series of poly(sulfide-ether-amide)s and poly(sulfoxide-ether-amide)s with aliphatic units in the main chain were prepared from diacids with various diamines.The polyamides were obtained in quantitative yields and their inherent viscosities were in the range of 0.43-0.89 dl g−1 at a concentration of 0.5 g dl−1 in N,N-dimethylacetamide (DMAc) solvent at 25 °C. They showed good thermal stability. The temperature for 10% weight loss in argon atmosphere was in the range of 350-415 °C. The polymers showed glass transition temperatures between 228 and 261 °C. Almost all of the polyamides were readily soluble in a variety of polar solvents such as N-methyl-2-pyrrolidone (NMP) and dimethyl sulfoxide (DMSO).  相似文献   

9.
A highly reproducible and sensitive signal-on electrogenerated chemiluminescence (ECL) biosensor based on the DNAzyme for the determination of lead ion was developed. The ECL biosensor was fabricated by covalently coupling 5′-amino-DNAzyme-tagged with ruthenium bis (2,2′-bipyridine) (2,2′-bipyridine-4,4′-dicarboxylic acid)-ethylenediamine (Ru1-17E′) onto the surface of graphite electrode modified with 4-aminobenzoic acid, and then a DNA substrate with a ribonucleotide adenosine hybridized with Ru1-17E′ on the electrode. Upon binding of Pb2+ to the Ru1-17E′ to form a complex which catalyzed the cleavage of the DNA substrate, the double-stranded DNA was dissociated and thus led to a high ECL signal. The signal linearly increases with the concentration of Pb2+ in the range from 5.0 to 80 pM with a detection limit of 1.4 pM and a relative standard derivation of 2.3%. This work demonstrates that using DNAzyme tagged with ruthenium complex as an ECL probe and covalently coupling method for the fabrication of the ECL biosensor with high sensitivity, good stability and significant regeneration ability is promising approach.  相似文献   

10.
Photo-responsive spiropyran-based compounds, such as, 1′,3′,3′-trimethyl-6-hydroxy-spiro(2H-1-benzopyran-2,2′-indoline) [OHSP], its monomer, such as 1′,3′,3′-trimethyl-6-methacryloyloxy-spiro(2H-1-benzopyran-2,2′-indoline) [MOSP] and its copolymers with methyl methacrylate [MMA] were synthesised using conventional synthetic routes. The copolymerisation was carried out either in tetrahydrofuran [THF] or in toluene using 2,2′-azobisisobutyronitrile [AIBN] as an initiator. The structures of these materials were investigated using 1H and 13C NMR spectroscopy. DEPT-135, HCCOSW and COSY45 NMR experiments were used to assign and interpret the complex structure of spiropyran based materials.  相似文献   

11.
A novel crown ether which incorporates the 2,2′-biimidazole moiety was prepared by cyclization of 1,1′-dibenzyl-1H,1′H-[2,2′]biimidazolyl-4,4′-dicarboxylic acid and 4,7,10-trioxa-1,13-tridecanediamine followed by removal of the benzyl groups. The diacid has been obtained by hydrolysis of the diester previously prepared by alcoholysis of the corresponding dicyano biimidazole. The cyano group is introduced by a palladium-catalyzed procedure starting from the corresponding dibromo biimidazole. The macrocyclic structure of the N-dibenzylated derivative of the receptor has been studied by X-ray diffraction. Binding constants for 1:1 biimidazole–anion complexation (Kassoc) are on the order of 105 M−1 for H2PO4 and Cl.  相似文献   

12.
Four studies of the 1H NMR spectrum for the aromatic protons of 4-fluoroaniline between 1958 and 1974 give three very different solutions to the second-order, AA′BB′X, spectrum. A re-evaluation of the second-order spectrum was done at 300 MHz. Simultaneous simulations of the 1H NMR spectrum and 19F NMR spectrum for 4-fluoroaniline were done using WINDNMR-Pro, and a new set of parameters for the six coupling constants was obtained from the optimized simulations. This new set of parameters was used as a basis to evaluate the AA′BB′X spectrum for the aromatic protons in N4-(4′-fluorophenyl)succinamic acid and in N4-(4′-fluorophenyl)-3,3-difluorosuccinamic acid by simultaneous simulations of the 1H NMR spectrum and 19F NMR spectrum for each using WINDNMR-Pro. Formation of the amide bond results in small, but significant, changes in the values for the set of parameters in both molecules. These results confirm that second-order analyses as an AA′BB′X system are required for derivatives of 4-fluoroaniline, rather than first-order analyses that have been used in previous reports.  相似文献   

13.
Optically active 2,2,4,4,6,6-hexafluorobiphenyl-3,3-dicarboxylic acid was obtained through its brucine salt. The half-life time for racemization was determined at various temperatures and the torsional barrier for racemization was calculated to be 25.4 kcal/mol. These results prove, contrary to textbook knowledge, that ortho,ortho-tetrafluoro substituted biphenyls are resolvable.  相似文献   

14.
彭宗海  付海燕  马梦林  陈华  李贤均 《催化学报》2010,31(12):1478-1482
 以 3-溴苯甲醚为原料合成了新型双膦配体 6,6′-二甲氧基-2,2′-二 (二-N-咔唑基膦)-1,1′联苯, 并将该配体与钯组成的配合物用于对溴苯甲醚和苯硼酸的 Suzuki 偶联反应, 考察了溶剂、碱、底物/催化剂摩尔比、膦/钯摩尔比对偶联反应的影响. 结果表明, 该催化体系在 1,4-二氧六环中催化对溴苯甲醚和苯硼酸的 Suzuki 偶联反应得到 99% 的分离产率. 同时, 该催化体系用于其它芳基溴和苯硼酸的 Suzuki 偶联反应也表现出很好的催化性能, 即使芳基溴有较大的空间位阻或具有取代基也能获得很好的结果.  相似文献   

15.
Nakano S  Matumoto Y  Yoshii M 《Talanta》2005,68(2):312-317
A novel flow-injection spectrophotometric method has been developed for the determination of manganese(II) at sub-nanogram/ml levels. The method is based on its catalytic effect on the oxidation of N,N′-bis(2-hydroxy-3-sulfopropyl)tolidine (HSPT) by periodate. The catalytic effect of manganese(II) was enhanced by the presence of 2,2′-bipyridine as an activator. By monitoring the change in absorbance of the oxidation product of HSPT at 670 nm, manganese(II) ranging 0.02-3.0 ng ml−1 could be determined with the relative standard deviations of less than 2%. The interfering ions were effectively suppressed by the addition of 2,2′-iminodiethanol and citric acid. The proposed method is directly applicable to the determination of manganese in lake and river water samples.  相似文献   

16.
A series of novel bifluorene based systems was synthesised by a convergent approach by means of a Suzuki cross-coupling between 7,7′-bis-(4,4,5,5-tetramethyl-[1,3,2]dioxaborolan-2-yl)-9,9,9′,9′-tetraoctyl-2,2′-bifluorene and suitable aryl-bromides. All the oligomers have been characterized by 1H, 13C NMR, FT-IR, UV-vis, PL spectroscopy and mass analyses. In particular, it has been demonstrated that the presence of strong electron donor (amines) or withdrawing (carboxylic esters) groups causes a bathochromic shift of the optical properties with respect to those of unsubstituted molecules. The effects of these functional groups on the HOMO-LUMO energy levels were investigated by cyclic voltammetry. Remarkably, the LUMO energy level of 7,7′-bis-[5′-carbodecaoxy-2,2′-bithiophen-5-yl]-9,9,9′,9′-tetraoctyl-2,2′-bifluorene (−3.07 eV) is strongly influenced by the presence of the ester functional group.  相似文献   

17.
Thin films of porphyrin-containing polyimide were produced by high vacuum co-evaporation of 4,4′-hexafluoroisopropylidene diphthalic anhydride (6FDA), 3,3′-diaminodiphenyl sulfone (DDS) and 5,10,15,20 meso-tetraphenyl porphyrin (TPP). The films were characterized by FT-IR analysis, optical absorption and emission spectroscopy. FT-IR analysis shows that the film matrix is comprised of only unreacted monomers. The conversion of monomers to polyamic acid and the following condensation to polyimide were studied by curing the samples at temperatures up to 240 °C. The amount of polyamic acid increases from room temperature to 120 °C, while at higher temperature it starts to condense to polyimide. Optical analysis shows that TPP is incorporated in the film matrix and its chemical state is determined by the interaction with the monomers, polyamic acid and polyimide. After curing the TPP molecules are finely dispersed in the polyimide matrix and their absorption and fluorescence properties are wholly preserved.  相似文献   

18.
A specific recognition material for bisphenol A (BPA) was prepared by using a covalent imprinting technique. A chloroform solution containing bisphenol A dimethacrylate as a template, ethylene glycol dimethacrylate as a cross-linking agent and 2,2′-azobis(isobutyronitrile) as an initiator was polymerized by UV initiation. When BPA was removed from the resulting polymer by hydrolysis of the ester bonds with aqueous sodium hydroxide, carboxylic acid residues were generated in the polymer. After the polymer was packed into a stainless steel column, retention factors of BPA and related compounds were measured. The imprinted polymer adsorbed BPA and structurally related compounds such as 4,4′-dihydroxybenzophenone, bis(4-hydroxyphenyl)sulfone and 4,4′-dihydroxybiphenyl. A typical association constant (Ka) was calculated to be 1.72×105 M−1 by Scatchard analysis. Interestingly, 17α- and 17β-estradiol were also bound to the imprinted polymer (Ka=1.68×105 M−1), indicating that the polymer could be used as artificial receptors for screening the compounds having estrogenic action.  相似文献   

19.
Fang Fang 《Tetrahedron letters》2009,50(48):6672-1951
A new family of achiral 3,3′,5,5′-tetrasubstituted-2,2′,6,6′-tetrahydroxy biphenyl ligand 4 was developed. The axial chirality of the ligand could be induced by the chelation of 2,2′,6,6′-tetrahydroxy groups with (R)-BINOL-Ti(OiPr)2 to form an axially chiral bimetallic titanium catalyst 9. Compared with (R)-BINOL-Ti(OiPr)2 catalyst, this novel catalyst 9 exhibited excellent activity and enantioselectivity for the carbonyl-ene reaction of methylstyrene and ethyl glyoxylate. 3,3′,5,5′-Tetrasubstituted groups showed a remarkable effect on both enantioselectivity and yield. With 9d prepared from 3,3′,5,5′-tetramethyl-2,2′,6,6′-tetrahydroxy biphenyl 4d as the catalyst, the best result, up to 97.6% ee and 99% yield, was obtained. Additionally, the bimetallic catalyst 9 also showed better catalytic capability than the corresponding monometallic catalyst.  相似文献   

20.
Ruthenium(II) complexes bearing a redox-active tridentate ligand 4′-(2,5-dimethoxyphenyl)-2,2′:6′,2′′-terpyridine (tpyOMe), analogous to terpyridine, and 2,2′-bipyridine (bpy) were synthesized by the sequential replacement of Cl by CH3CN and CO on the complex. The new ruthenium complexes were characterized by various methods including IR and NMR. The molecular structures of [Ru(tpyOMe)(bpy)(CH3CN)]2+ and two kinds of [Ru(tpyOMe)(bpy)(CO)]2+ were determined by X-ray crystallography. The incorporation of monodentate ligands (Cl, CH3CN and CO) regulated the energy levels of the MLCT transitions and the metal-centered redox potentials of the complexes. The kinetic data observed in this study indicates that the ligand replacement reaction of [Ru(tpyOMe)(bpy)Cl]+ to [Ru(tpyOMe)(bpy)(CH3CN)]2+ proceeds by a solvent-assisted dissociation process.  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