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 51 毫秒
1.
Catalysis of the solvolysis of organophosphorus esters by polymers of aliphatic amines, imidazole, pyridine, 2,2′-bipyridine, and their copper(II) complexes was studied using diisopropyl fluorophosphate (DEP) as a model substrate. The polymeric catalysts were synthesized either by (1) derivation of available polymers, including polyethylenimine, polyvinyl amine, polyvinyl alcohol, polystyrene, and poly-4-vinylpyridine or (2) by polymerization of functionalized monomers such as 4(5)-vinylimidazole and 4-vinyl-4′-methyl-2,2′-bipyridine. Polymer hydrophilicity was controlled by partial quaternization of amine groups with different alkyl halides. The greatest catalytic activity was exhibited by copper(II) complexes of polymers containing the 2,2′-bipyridine group. At pH 7.6 and 3.7 × 10?3M, the most active of these catalysts reduced the half-life of DFP from 800 to 9 min. The rate was largely independent of the pH in the range 6.5–8.5 but was limited by the aqueous solubility of the catalyst. Heterogeneous catalysis by some polymers was observed but was less effective. A Lineweaver-Burk plot of V0?1 versus [DFP]?1 for a soluble polymeric 2,2′-bipyridine-copper(II) catalyst was linear. There was no correlation between catalysis of solvolysis of DFP and the carboxylic ester, p-nitrophenyl acetate.  相似文献   

2.
The grafting of poly(methyl methacrylate) (PMMA) onto biosynthesized γ-poly(glutamic acid) (γ-PGA) initiated by pendant azo groups introduced onto γ-PGA was performed. The introduction of pendant azo groups onto γ-PGA was achieved by the reaction of carboxyl groups of γ-PGA with azo initiators having hydroxyl or amino groups, such as 2,2-azobis[2-(hydroxymethyl)propionitrile] (AHP), 2,2′-azobis[2-methyl-N-(2-hydroxyethyl)propionamide] (AMHP), and 2,2′-azobis[2-(2-imidazolin-2-yl)propane] (AIP), using N,N′-dicyclohexylcarbodiimide. The amount of pendant AHP groups introduced onto γ-PGA was estimated to be 0.15 mmol/g. Untreated γ-PGA failed to initiate the polymerization of MMA. On the contrary, the polymerization of MMA was found to be initiated in the presence of γ-PGA having azo groups: the polymerization rate was proportional to the square root of the concentration of γ-PGA having pendant azo groups. During the polymerization PMMA was grafted onto γ-PGA; the percentage of grafting of PMMA onto γ-PGA obtained from the graft polymerization initiated by pendant AHP, AMHP, and AIP groups was evaluated to be 65.0, 53.1, and 29.0%, respectively. Differential scanning calorimetric analysis shows that the endotherm transition point of γ-PGA at 220°C disappears by the grafting of PMMA onto the polymer. © 1993 John Wiley & Sons, Inc.  相似文献   

3.
Polysiloxanes containing pendant tris(2,2′-bipyridine)ruthenium(II) complex (Ru(bpy)32+) were prepared by reaction of polysiloxane-pendant 2,2′-bipyridine (PSiO-bpy) with cis-Ru(bpy)2Cl2. In methanol solution, the polymer pendant Ru(bpy)32+ showed absorption maximum at 456nm and emission maximum at around 609nm, both of which are shifted to longer wavelength than the monomeric Ru(bpy)32+. The lifetime τ0 of the excited polymer complex with low Ru(bpy)32+ content was almost the same as that of the monomeric one in methanol (830ns), but τ0 of the polymer with higher complex content was shorter because of a concentration quenching. In a solid state, τ0 was much shorter (306–503ns) than that in a methanol solution contrary to the conventional polymeric system. Higher complex content in the polymer film caused higher glass transition temperature (Tg), but shorter τ0. These results indicate concentration quenching in the polymer film. The excited polymer pendant Ru(bpy)32+ was quenched by oxygen, and the relative emission intensity followed the Stern-Volmer equation. In a methanol solution the quenching rate constant (kq) was the same order of magnitude as the monomeric complex, and independent of the complex content in the polymer. In a film, kq was higher for the polymer with higher complex content.  相似文献   

4.
Two substituted 2,2′-bipyridine lead(II) complexes, [Pb(5,5′-dm-2,2′-bpy)(tfac)2] n (1) (5,5′-dm-2,2′-bpy?=?5,5′-dimethyl-2,2′-bipyridine and tfac?=?trifluoroacetate) and [Pb2(4,4′-dmo-2,2′-bpy)2(ftfa)4] (2) (4,4′-dmo-2,2′-bpy?=?4,4′-dimethoxy-2,2′-bipyridine and ftfa?=?furoyltrifluoroacetonate), have been synthesized and characterized by elemental analysis, IR, 1H NMR, and 13C NMR spectroscopies, thermal behavior, and X-ray crystallography. Complexes 1 and 2 are 1D coordination polymer and dinuclear complex, respectively. The supramolecular features in these complexes are guided by weak directional intermolecular interactions.  相似文献   

5.
Some metal containing block copolymers derived from polystyrene-block-poly(4-vinylpyridine) were synthesized. The complexation was achieved by reacting the poly(4-vinylpyridine) block with 2,2′-bipyridyl(tricarbonyl)rhenium(I) chloride in the presence of silver perchlorate. The resulting polymer metal complexes are able to form micelles in different solvent systems. The morphologies were studied by transmission electron microscopy, and the structure of the micelles was found to be solvent dependent.  相似文献   

6.
A chiral metal-organic coordination polymer, [Cu(Tar)(2,2′-bipy) · 5H2O] (1) (Tar = L-tartrate dianion, 2,2′-bipy = 2,2′-bipyridine), has been synthesized by hydrothermal reaction of Cu(OAc)2, Na2T (H2T = 2,3-O-isopropylidene-L-tartaric acid) and 2,2′-bipyridine, and characterized by IR, UV–vis spectra, elemental analyses, TG-DTA, and single crystal X-ray diffraction. In the hydrothermal reaction, the protection group isopropylidene for tartaric acid was hydrolyzed. The crystal structure of the coordination polymer 1 shows that each tartrate chelates two Cu(II) ions at opposite ends using one carboxylate oxygen and one hydroxyl oxygen and each Cu(II) ion is chelated by two halves of tartrate dianions, forming coordination polymer chains. Distorted octahedral geometry around copper is completed by a chelating 2,2′-bipyridine molecule. The 2,2′-bipyridine groups in two of parallel 1-D chains are interwoven, constituting ladder-shaped double chains. Strong offset π–π stacking interactions with a face-to-face distance of 3.33 Å for pyridine rings are observed. All the lattice water molecules hydrogen-bond to each other or to the carbonyl oxygen of tartrate, forming a 3-D supramolecular structure.  相似文献   

7.
Heteroleptic ruthenium(II) bioflavonoid complexes of quercetin, morin, chrysin, and 3-hydroxyflavone were prepared and their interaction with CT DNA and BSA along with antioxidant and in vitro anticancer and antimicrobial activities was investigated. The formulation and characterization of complexes were achieved through elemental and thermal analysis, mass spectrometry, 1H NMR spectroscopy along with infrared, electronic absorption, and emission spectroscopy as well as square-wave voltammetry, and magnetic and conductivity measurements. Ruthenium(II) is octahedrally coordinated in cationic complex species to two bidentate diimine ligands (2,2′-bipyridine or 1,10-phenanthroline) and one bidentate monobasic flavonoid ligand through 3,4-site of quercetin, morin, and 3-hydroxyflavone or 4,5-site of chrysin. Complexes bind CT DNA by intercalation and binding constants comparable to ethidium bromide or 10 times higher. Binding constants of complexes to BSA were several times higher compared to ibuprofen and diazepam, and suggest that the complexes have a strong affinity to BSA. Antioxidant activity tests showed that the complexes are more potent in terms of radical inhibition compared to the parent flavonoids. Cytotoxic testing revealed that the Ru(II) complex of quercetin with 2,2′-bipyridine co-ligand has good selectivity to breast adenocarcinoma, while the complex of 3-hydroxyflavone with 2,2′-bipyridine co-ligand showed strong cytotoxicity toward all tested cell lines with IC50 ~ 1 μM. All complexes showed moderate activity toward Acinetobacter baumannii, while the Ru(II) complex of 3-hydroxyflavone with 2,2′-bipyridine showed excellent activity toward MRSA and Candida albicans.  相似文献   

8.
LiMnPO4 (olivine) was surface-modified by two different complexes: Ru-bis(4,4′-diethoxycarbonyl-2,2′-bipyridine)(4,4′-dicarboxylate-2,2′-bipyridine) and Ru-bis(4-carboxylic acid-4′-carboxylate-2,2′-bipyridine)(4,4′-dinonyl-2,2′bipyridine). These complexes have redox potentials of 4.45 and 4.25 V vs. Li/Li+, respectively, and are both active for molecular wiring of LiMnPO4. The surface-confined Ru(II)/Ru(III) redox reaction propagates across the monolayer via hole-hopping, allowing a subsequent chemical delithiation of the underneath olivine towards MnPO4. The activity of LiMnPO4 is about half of that of LiFePO4 (olivine) at similar experimental conditions.  相似文献   

9.
A secondary building unit (SBU), [Ni(2,2′-bipy)(5-npa)(H2O)] n [where 2,2′-bipy = 2,2′-bipyridine, 5-npa = 5-nitroisophthalic dianion], was synthesized as starting material of a polystep reaction. A ladderlike complex (LLC) Ni(II) coordination polymer, {[Ni(2,2′-bipy)(5-npa)(4,4′bipy)0.5]·(H2O)} n , was constructed by polystep reaction using this SBU. In LLC, two SBUs were cross-linked by 4,4′-bipy [where 4,4′-bipy = 4,4′-bipyridine] forming a 1-D ladderlike structure. The magnetic properties of the LLC and SBU are discussed.  相似文献   

10.
Novel polymer macroligands—copolyamides containing different quantity of bipyridyl groups—were obtained from 4,4′-diamino-2,2′-bipyridine, 4,4′-diaminodiphenyl ether and terephthaloyl-bis(3-methoxy-4-hydroxybenzoic) acid dichloroanhydride by low-temperature polycondensation. Metal–polymer complexes with different content of Ir(ppy)2 were obtained by the interaction between polymer ligand and [Ir(ppy)2Cl]2 (ppy–2-phenylpyridine). The properties of films and coatings based on these materials were studied.  相似文献   

11.
Polymeric pendant Ru(bpy)_3~(2+) complexes were prepared from homopolymer and copolymers of 4-methyl-4'-vinyl-2,2'-bipyridine (Vbpy). Vbpy was prepared from 4-methylpyridine. The comonomers were styrene (St), acrylic acid (AA), N-vinylpyrrolidone (Pyr), 4-vinylpyridine (Vpy), methyl methacrylate (MMA), 2-hydroxyethyl methacrylate (HEMA), acrylonitrile (AN) and N-ethyl-4-vinylpyridium bromide (EQ-Vpy). The fraction of the pendant Ru(bpy)_3~(2+) repeating unit in the polymeric complex was 0.022 to 0.052. Absorption maximum, molar extinction coefficient, emission maximum and relative emission intensity of the polymeric complexes were studied.  相似文献   

12.
The lowest electronic excited state of the complexes [Ru(2,2′-bipyridine)3]2+, fac-[ClRe (CO)3(2,2′-bipyridine)], and fac-[(pyridine) Re (CO)3(2,2′-bipyridine)]+ can be quenched by methyl viologen, MV2+, N,N′-dimethyl-4,4′-bipyridinium, in fluid solutions. The quenching obeys Stern—Volmer kinetics as deduced from plots of relative luminescence quantum yield vs [MV2+], and the data are consistent with a quenching process that is essentially diffusion controlled. Pulsed laser excitation (18 ns, 354.7 nm frequency tripled Nd: YAG) of the metal complexes in the presence of MV2+ shows that a detectable fraction of the quenching results in net electron transfer to form MV+. The MV+ is detectable by resonance Raman scattering from the trailing portion of the excitation pulse. Excited state electron transfer to MV2+ from a photo-excited complex anchored to SiO2 has also been detected by transient Raman spectroscopy. High surface area SiO2 was functionalized by reaction with 4-[2-(trimethoxysilyl)ethyl]pyridine to give [SiO2]-SiEtpyr. Reaction of [SiO2]-SiEtpyr with [(CH3CN)Re(CO)3(2,2′-bipyridine)]+ then yields [SiO2]-[(SiEtpyr) Re (CO)3 (2,2′-bipyridine)]+. Electron transfer quenching of the photo-excited immobilized Re complex occurs when suspended in CH3CN solutions of MV2+ to yield MV+ as detected by resonance Raman scattering and by lifetime attenuation in the presence of MV2+.  相似文献   

13.
Five substituted-2,2′-bipyridine ligands L, (4-(p-methylphenyl)-6-phenyl-2,2′-bipyridine (L1), 4-(p-bromophenyl)-6-(p-bromophenyl)-2,2′-bipyridine (L2), 4-(p-bromophenyl)-6-phenyl-2,2′-bipyridine (L3), 4-phenyl-6-(p-bromophenyl)-2,2′-bipyridine (L4), and 4-(p-fluorophenyl)-6-(p-fluorophenyl)-2,2′-bipyridine (L5) were synthesized by stepwise formation. Reaction of cis-[RuCl2(bpy)2]?2H2O or cis-[RuCl2(phen)2]?2H2O and the substituted-2,2′-bipyridine ligands in the presence of KPF6 afforded the corresponding cationic polypyridine-ruthenium complexes of the type [(bpy)2Ru(L)](PF6)2 (bpy = 2,2′-bipyridine, 15) or [(phen)2Ru(L)](PF6)2 (phen = 1,10-phenanthroline, 610), respectively. All complexes have been spectroscopically characterized by UV–vis, luminescence, and electrogenerated chemiluminescence. The structures of 1?CH3COCH3, 3?CH3COCH3, 5?2CH3COCH3, 6, 8, 9, and 10 have been determined by single-crystal X-ray diffraction.  相似文献   

14.
Zhai S  Chen Y  Wang S  Jiang J  Dong S  Li J 《Talanta》2004,63(4):927-931
Photoluminescent multilayers were fabricated by layer-by-layer deposition between europium-substituted heteropolytungstate K13[Eu(SiW11O39)2]·28H2O (denoted ESW) and a cationic polymer of quaternized poly(4-vinylpyridine) partially complexed with osmium bis(2,2′-bipyridine) (denoted as QPVP-Os) on glassy carbon and quartz substrates. The resulting photoluminescent organic-inorganic hybrid multilayers were characterized by electrochemical impedance spectroscopy, UV-Vis absorption spectrometry, cyclic voltammetry and photoluminescence spectra. Electrochemical impedance spectroscopy, UV-Vis absorption spectrometry and cyclic voltammetry results demonstrated that the multilayers were regular growth each layer adsorption. The photoluminescent properties of the films at room temperature were investigated to show the characteristic Eu3+ emission pattern of 5D07Fj.  相似文献   

15.
A series of CN-bridged trinuclear Ru complexes of the general structure [RuL2(μ-(CN)Ru(CN)L2′)2] where L is 2,2′-bipyridine-4,4′-dicarboxylic acid and L′ is 2,2′-bipyridine ( 1 )2,2′-bipyridine-4,4′-dicarboxylic acid ( 2 ), 4,4′-dimethyl-2,2′-bipyridine ( 3 ), 4,4′-diphenyl-2,2′-bipyridine ( 4 ), 1,10-phenanthroline ( 5 ), and bathophenanthrolinedisulfonic acid ( 6 ) have been synthesized, and their spectral and electrochemical properties investigated. The two carboxylic functions on the 2,2′-bipyridine ligand L serve as interlocking groups through which the dye is attached at the surface of TiO2 films having a specific surface texture. The role of these interlocking groups is to provide strong electronic coupling between the π* orbital of the 2,2′-bipyridine and the 3d-wave-function manifold of the conduction band of the TiO2, allowing the charge injection to proceed at quantum yields close to 100 %. The charge injection and recombination dynamics have been studied with colloidal TiO2, using laser photolysis technique in conjunction with time-resolved optical spectroscopy. Photocurrent action spectra obtained from photo-electrochemical experiments with these trinuclear complexes cover a very broad range in the visible, making them attractive candidates for solar light harvesting. Monochromatic incident photon-to-current conversion efficiencies are strikingly high exceeding 80% in some cases. Performance characteristics of regenerative cells operating with these trinuclear complexes and ethanolic triiodide/iodide redox electrolyte have been investigated. Optimal results were obtained with complex 1 which gave a fill factor of 75 % and a power conversion efficiency of 11.3% at 520 nm.  相似文献   

16.
The synthesis of a polymerizable methacrylate bearing the dibenzo-18-crown-6 moiety ( 5 ) was effected in three steps by condensation of 3,4-dihydroxybenzaldehyde with 1,2-bis[2′-(2″-chloroethoxy)ethoxy]benzene ( 2 ), reduction of the resulting 4-formyldibenzo-18-crown-6 ( 3 ) to 4 and subsequent reaction of 4 with methacryloyl chloride. The methacrylate ( 5 ) was polymerized in solution using 2,2′-azobis(isobutyronitrile) (AIBN) as the initiator and gave a solid polymer ( 6 ) having a midrange Tg of 67°C. The polymethacrylate bearing pendant crown-ether structure ( 6 ) was found to effectively complex potassium ions and was clearly more effective in its complexing ability than dibenzo-18-crown-6 ( 7 ) or a polyurethane bearing the dibenzo-18-crown-6 moiety ( 8 ) in the polymer backbone. The subject polymer ( 6 ) was also found to complex and consequently solubilize several metal 7,7,8,8-tetracyanoquinodimethane (TCNQ) salts to give semiconducting compositions. Evidence is given that in the presence of the neutral acceptor molecule, 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ), the polymethacrylate ( 6 ) acting as a donor, gives rise to a charge-transfer (C/T) complex.  相似文献   

17.
A novel phosphate monomer, Op‐(methacryloyloxymethyl)benzyl O,O‐diethyl phosphate (MDP) was synthesized by the reaction of diethyl phosphorochloridate with 1,4‐benzenedimethanol, followed by the reaction with methacryloyl chloride in the presence of triethylamine. The radical polymerization of MDP and copolymerization with methyl methacrylate were carried out in the presence of 2,2′‐azobisisobutyronitrile (3 mol %) in dimethylacetamide at 60 °C for 20 h to afford phosphate‐pendant polymers. The polymerization of glycidyl phenyl ether (GPE) was carried out with the phosphate‐pendant polymer as an initiator in the presence of ZnCl2. The polymerization did not proceed below 90 °C but rapidly proceeded above 90 °C to afford polyGPE. The phosphate‐pendant polymer served as a good thermally latent polymeric initiator. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 39: 3365–3370, 2001  相似文献   

18.
Methods for the synthesis of (pyridinyl)-1,2,4-triazolo[4,3-a]pyridines were developed. The principal route to the required intermediate 2-chloropyridines was based on rearrangements of mono N-oxides of 2,2′-bipyridine, 2,3′-bipyridine, 3,3′-bipyridine, 2,4′-bipyridine and 4,4′-bipyridine with phosphorus oxychloride. Reaction of 3,3′-bipyridine 1-oxide or 2,2′-bipyridine 1-oxide with phosphorus oxychloride gave mixtures of chloro isomers. Reaction with acetic anhydride, 3,3′-bipyridine 1-oxide and 2,2′-bipyridine 1-oxide gave exclusively [3,3′-bipyridine]-2(1H)-one and [2,2′-bipyridine]-6(1H)-one, respectively. 1,2,4-Triazolo[4,3-a]pyridines with pyridinyl groups at the 5,6,7 and 8 positions were synthesized.  相似文献   

19.
We report the electrodeposition, under physiological conditions, of an electrochemiluminescent (ECL) Ru2+/3+ complex-containing redox hydrogel. The ECL-hydrogels were formed by potential cycling of a solution of [poly(4-vinylpyridine)Ru(2,2′-bipyridine)2Cl?]+/2+, its un-coordinated backbone pyridines partially quaternized with bromoethylamine for solubility in water and for swelling to a hydrogel after crosslinking. The polymer was electrosorbed on plasma-oxidized graphite in the anodic half of the cycle and irreversibly crosslinked, to form a swelling but insoluble film, in the cathodic half cycle. The ECL resulted of the chemical reaction of electro-oxidatively produced tri-n-propylamine-radical with the hydrogel’s Ru2+ centers. The emission spectra of the photo-excited films and their ECL spectra were identical. The ECL-emission increased one thousand-fold, linearly with the tri-n-propylamine (TPrA) concentration, between 100 nM and 0.1 mM.  相似文献   

20.
A new ligand based on a p-tert-butylcalix[4]arene functionalised with three 6-carboxylato-2,2′-bipyridine arms reacted with lanthanide(III) cations to form 2:2 complexes in the solid state, whereas in solution, a concentration-dependent equilibrium is observed between 2:2 and 1:1 species, as evidenced by ES-MS and metal luminescence measurements. In the X-ray molecular structure of the terbium complex two branched calixarene platforms share one pendant arm in order to provide a neutral dimeric structure which is held together by a strong hydrogen bonded network together with efficient π–π stacking of two pyridine rings belonging to each ligand.  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