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 polyaddition of bis(oxetane)s 1,4‐bis[(3‐ethyl‐3‐oxetanylmethoxymethyl)]benzene (BEOB), 4,4′‐bis[(3‐ethyl‐3‐oxetanyl)methoxy]benzene (4,4′‐BEOBP), 1,4‐bis[(3‐ethy‐3‐oxetanyl)methoxy] ‐benzene (1,4‐BEOMB), 1,2‐bis[(3‐ethyl‐3‐oxetanyl)methoxy]benzene (1,2‐BEOMB), 4,4‐bis[(3‐ethyl‐3‐oxetanyl)methoxy]biphenyl (4,4′‐BEOMB), 3,3′,5,5′‐tetramethyl‐[4,4′‐bis(3‐ethyl‐3‐oxetanyl)methoxy]biphenyl (TM‐BEOBP) with active diesters di‐s‐phenylthioterephthalate (PTTP), di‐s‐phenylthioisoterephthalate (PTIP), 4,4′‐di(p‐nitrophenyl)terephthalate (NPTP), 4,4′‐di(p‐nitrophenyl)isoterephthalate (NPIP) were carried out in the presence of tetraphenylphosphonium chloride (TPPC) as a catalyst in NMP for 24 h, affording corresponding polyesters with Mn's in the range 2200–18,200 in 41–98% yields. The obtained polymers would soluble in common organic solvents and had high thermal stabilities. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 1528–1536, 2004  相似文献   

2.
Chemical investigation of the red herbal resin of Dracaena cochinchinensis resulted in the isolation of three new configurationally isomeric flavonoids: 6,4′‐dihydroxy‐7‐methoxy‐8‐methylflavane (=3,4‐dihydro‐2‐(4‐hydroxyphenyl)‐7‐methoxy‐8‐methyl‐2H‐[1]benzopyran‐6‐ol; 1 ), 5,4′‐dihydroxy‐7‐methoxy‐6‐methylflavane (=3,4‐dihydro‐2‐(4‐hydroxyphenyl)‐7‐methoxy‐6‐methyl‐2H‐[1]benzopyran‐5‐ol; 2 ), and 7,4′‐dihydroxy‐5‐ methoxyhomoisoflavane (=3,4‐dihydro‐3‐[(4‐hydroxyphenyl)methyl]‐5‐methoxy‐2H‐[1]benzopyran‐7‐ol; 3 ). Their structures were identified by means of detailed spectral analysis. In addition, thirteen known compounds were isolated from D. cochinchinensis: 7‐hydroxy‐4′‐methoxyflavane ( 4 ), 2,4,6‐trimethoxy‐4′‐hydroxydihydrochalcone ( 5 ), 2,4‐dimethoxy‐4′‐hydroxydihydrochalcone ( 6 ), 7,8‐(methylenedioxy)‐4′‐hydroxyhomoisoflavane ( 7 ), 4′,7‐dihydroxy‐8‐methylflavane ( 8 ), 2,6‐dimethoxy‐4,4′‐dihydroxydihydrochalcone ( 9 ), 2‐methoxy‐4,4′‐dihydroxydihydrochalcone ( 10 ), 7‐methoxy‐6,4′‐dihydroxyhomoisoflavane ( 11 ), 2‐methoxy‐4,4′‐dihydroxychalcone ( 12 ), 4′,7‐dihydroxyflavane ( 13 ), 7,4′‐dihydroxyhomoisoflavane ( 14 ), 7,4′‐dihydroxyhomoisoflavone ( 15 ), and 7,4′‐dihydroxyflavone ( 16 ). Compounds 7, 8, 9, 14 , and 15 have been isolated for the first time from this type of herbal source.  相似文献   

3.
Two pairs of enantiomeric compounds with formulas (S)‐ or (R)‐Co3(ppap)2(4,4′‐bpy)2(H2O)2 ? 4 H2O [(S)‐ 1 or (R)‐ 1 ], (S)‐ or (R)‐Co3(ppap)2(4,4′‐bpy)2(H2O)2 ? 3 H2O [(S)‐ or (R)‐ 2 ), and related racemic compound Co3(ppap)2(4,4′‐bpy)2(H2O)2 ? 4 H2O (rac‐ 3 ; 4,4′‐bpy=4,4′‐bipyridine, H3ppap=3‐phenyl‐2‐[(phosphonomethyl)amino]propanoic acid) are reported. Compounds 1 and rac‐ 3 show identical three‐dimensional framework structures, whereas compounds 2 have two‐dimensional layer structures. Compounds 1 and 2 are catenation isomers, formation of which is controlled solely by the pH of the reaction mixtures, whereas the formation of isomeric compounds 1 and rac‐ 3 is controlled purely by the chirality of the phosphonate ligand. The magnetic properties of fully dehydrated (S)‐ 1 , (S)‐ 2 , and rac‐ 3 are highly dependent on both structure and chirality.  相似文献   

4.
The reaction of Cu(Ac)2·4H2O with 2-[(3,5-dichloro-2-hydroxy-benzylidene)-amino]-2hydroxymethyl-propane-1,3-diol (H4L) and 4,4′-bipyridyl-N,N′-dioxide (4,4′-bipy-NO) in DMF under solvothermal conditions leads to the formation of a di-radical dinuclear copper complex [Cu2(4,4′-bipy-NO)(dcdmap)4] (1) (dcdmap is the anion of 2,4-dichloro-6dimethylaminomethyl-phenol, synthesized by an in situ reaction). Compound 1 was characterized by elemental analysis, IR spectroscopy, and X-ray single-crystal diffraction. Complex 1 displays dominant anti-ferromagnetic interaction between oxyradicals and copper ion and oxyradical.  相似文献   

5.
The organotin(IV) chlorides RnSnCl4−n (n = 3, R = Ph, PhCH2, n−Bu; and n =2, R = n−Bu, Ph, PhCH2) react with 4,4′‐bipyridine (4′4‐bpy) to give [(Ph3SnCl)2(4,4′‐bpy)1.5(C6H6)0.5] ( 1 ), [(PhCH2)3‐ SnCl]2 (4,4′‐bpy) ( 2 ), [(n−Bu)3SnCl]2(4,4′‐bpy) ( 3 ), [(n−Bu)2SnCl2(4,4′‐bpy)] ( 4 ), [Ph2SnCl2(4,4′‐bpy)] ( 5 ), and [(PhCH2)2SnCl2(4,4′‐bpy)] ( 6 ). The new complexes have been characterized by elemental analyses, IR, 1H, 13C, 119Sn NMR spectroscopy. The structures of ( 1 ), ( 2 ), ( 4 ), and ( 6 ) have been determined by X‐ray crystallography. Crystal structures of ( 1 ) and ( 2 ) show that the coordination number of tin is five. In complex ( 1 ), two different molecules exist: one is a binuclear molecule bridged by 4,4′‐bpy and another is a mononuclear one, only one N of 4,4′‐bpy coordinate to tin. Complex ( 2 ) contains an infinite 1‐D polymeric binuclear chain by weak Sn…Cl intermolecular interactions with neighboring molecules. In the complexes ( 4 ) and ( 6 ), the tin is six‐coordinate, and the 4,4′‐bpy moieties bridge adjacent dialkyltin(IV)dichloride molecules to form a linear chain. © 2004 Wiley Periodicals, Inc. Heteroatom Chem 15:338–346, 2004; Published online in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/hc.20016  相似文献   

6.
1,1‐Bis[4‐(4‐aminophenoxy)phenyl]‐1‐phenylethane (BAPPE) was prepared through nucleophilic substitution reaction of 1,1‐bis(4‐hydroxyphenyl)‐1‐phenylethane and p‐chloronitrobenzene in the presence of K2CO3 in N,N‐dimethylformamide, followed by catalytic reduction with hydrazine and Pd/C. Novel organosoluble polyimides and copolyimides were synthesized from BAPPE and six kinds of commercial dianhydrides, including pyromellitic dianhydride (PMDA, Ia ), 3,3′,4,4′‐benzophenonetetracarboxylic dianhydride (BTDA, Ib ), 3,3′,4,4′‐ biphenyltetracarboxylic dianhydride (BPDA, Ic ), 4,4′‐oxydiphthalic anhydride (ODPA, Id ), 3,3′,4,4′‐diphenylsulfonetetracarboxylic dianhydride (DSDA, Ie ) and 4,4′‐hexafluoroisopropylidenediphthalic anhydride (6FDA, If ). Differing with the conventional polyimide process by thermal cyclodehydration of poly(amic acid), when polyimides were prepared by chemical cyclodehydration with N‐methyl‐2‐pyrrolidone as used solvent, resulted polymers showed good solubility. Additional, Ia,b were mixed respectively with the rest of dianhydrides (Ic–f) and BAPPE at certain molar ratios to prepare copolyimides with arbitrary solubilities. These polyimides and copolyimides were characterized by good mechanical properties together with good thermal stability. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 2082–2090, 2000  相似文献   

7.
Two new 2 : 1 co‐crystals based on [4,4′‐bithiazole]‐2,2′‐diamine (=2,2′‐diamino‐4,4′‐bithiazole (DABTZ)) with 2,2′‐bipyridine (bipy) and benzo‐18‐crown‐6 (bk) were synthesized by slow‐evaporation method in MeOH. These co‐crystals were characterized by means of elemental analysis, and IR, and 1H‐ and 13C‐NMR spectroscopy. Also, thermal analyses under air atmosphere and X‐ray crystallography have been performed on these structures. X‐Ray single‐crystal analyses revealed that these networks contain large vacant voids. These structures, [(DABTZ)2(bipy)] and [(DABTZ)2(bk)(MeOH)], crystallized in monoclinic and triclinic forms with space groups of P21/c and P , respectively. The self‐assembly of these compounds in the solid state is likely caused by both H‐bonding and π? π stacking.  相似文献   

8.
The amino functions of the common 2′‐deoxyribo‐ and ribonucleosides were blocked by the (2‐cyanoethoxy)carbonyl group on treatment with 2‐cyanoethyl carbonochloridate ( 5 ) or 1‐[(2‐cyanoethoxy)carbonyl]‐3‐methyl‐1H‐imidazolium chloride ( 6 ) leading to 7 , 18 , 8 , 19 , 9 , and 20 . In 2′‐deoxyguanosine, the amide group was additionally blocked at the O6 position by the 2‐cyanoethyl (→ 27 ) and 2‐(4‐nitrophenyl)ethyl group (→ 31 , 32 ). Comparative kinetic studies regarding the cleavage of the ce/ceoc and npe/npeoc group by β‐elimination revealed valuable information about the ease and sequential deprotection of the various blocking groups at different sites of the nucleobases. Besides the 5′‐O‐(dimethoxytrityl)‐protected 3′‐(2‐cyanoethyl diisopropylphosphoramidites) 38 and 39 of N4‐[(2‐cyanoethoxy)carbonyl]‐2′‐deoxycytidine and N6‐[(2‐cyanoethoxy)carbonyl]‐2′‐deoxyadenosine, respectively, the N2‐[(2‐cyanoethoxy)carbonyl]‐2′‐deoxy‐O6‐[2‐(4‐nitrophenyl)ethyl]guanosine analog 40 is recommended as building block for oligo‐2′‐deoxyribonucleotide synthesis.  相似文献   

9.
The reactions of butadienylketene with variety of 1,4‐diazabuta‐1,3‐dienes are studied. The reactions resulted in the formation of previously unknown functionalized cis butadienyl‐4‐iminomethyl‐azetidin‐2‐ones and butenylidene‐butadienyl‐[2,2′‐biazetidine]‐4,4′‐ diones. Butadienyl ketene reacts in [2+2] cycloaddition fashion with both iminic portion of 1,4‐ diazabuta‐1,3‐dienes and competitive [4+2] cycloaddition reaction of 1,4‐diazabuta‐1,3‐dienes as 4π component with butadienyl ketene as 2π component are not observed.  相似文献   

10.
A total of 13 of the 16 possible methylol derivatives of 2,4′‐ and 4,4′‐dihydroxydiphenylmethane have been synthesized, isolated, and identified. These compounds are found as intermediates in the cure process of resol phenol–formaldehyde (PF) resins. Analysis of the 13C NMR spectra (in acetone‐d6) of these compounds provided a way to evaluate the seven methylolphenol ring types (methylol derivatives of 2‐hydroxyphenyl and 4‐hydroxyphenyl rings) found in typical resol PF resins using the ipso carbon region from 150 to 160 ppm. A simple diagnostic test was developed using the chemical shift values of the methylol methylene carbon atoms to identify the presence of intermediates containing either a 2‐hydroxyphenyl or a 4‐hydroxyphenyl ring. Using these data it is now possible to analyze the major components in extracted prepreg PF resins. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

11.
Novel optically active substituted acetylenes HC? CCH2CR1(CO2CH3)NHR2 [(S)‐/(R)‐ 1 : R1 = H, R2 = Boc, (S)‐ 2 : R1 = CH3, R2 = Boc, (S)‐ 3 : R1 = H, R2 = Fmoc, (S)‐ 4 : R1 = CH3, R2 = Fmoc (Boc = tert‐butoxycarbonyl, Fmoc = 9‐fluorenylmethoxycarbonyl)] were synthesized from α‐propargylglycine and α‐propargylalanine, and polymerized with a rhodium catalyst to provide the polymers with number‐average molecular weights of 2400–38,900 in good yields. Polarimetric, circular dichroism (CD), and UV–vis spectroscopic analyses indicated that poly[(S)‐ 1 ], poly[(R)‐ 1 ], and poly[(S)‐ 4 ] formed predominantly one‐handed helical structures both in polar and nonpolar solvents. Poly[(S)‐ 1a ] carrying unprotected carboxy groups was obtained by alkaline hydrolysis of poly[(S)‐ 1 ], and poly[(S)‐ 4b ] carrying unprotected amino groups was obtained by removal of Fmoc groups of poly[(S)‐ 4 ] using piperidine. Poly[(S)‐ 1a ] and poly[(S)‐ 4b ] also exhibited clear CD signals, which were different from those of the precursors, poly[(S)‐ 1 ] and poly[(S)‐ 4 ]. The solution‐state IR measurement revealed the presence of intramolecular hydrogen bonding between the carbamate groups of poly[(S)‐ 1 ] and poly[(S)‐ 1a ]. The plus CD signal of poly[(S)‐ 1a ] turned into minus one on addition of alkali hydroxides and tetrabutylammonium fluoride, accompanying the red‐shift of λmax. The degree of λmax shift became large as the size of cation of the additive. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

12.
Three‐ and five‐membered rings that bear the (Si‐C‐S ) and (Si‐C‐C‐C‐S ) unit have been synthesized by the reactions of L SiCl ( 1 ; L =PhC(NtBu)2) and L′ Si ( 2 ; L′ =CH{(C?CH2)(CMe)(2,6‐iPr2C6H3N)2}) with the thioketone 4,4′‐bis(dimethylamino)thiobenzophenone. Treatment of 4,4′‐bis(dimethylamino)thiobenzophenone with L SiCl at room temperature furnished the [1+2]‐cycloaddition product silathiacyclopropane 3 . However, reaction of 4,4′‐bis(dimethylamino)thiobenzophenone with L′ Si at low temperature afforded a [1+4]‐cycloaddition to yield the five‐membered ring product 4 . Compounds 3 and 4 were characterized by NMR spectroscopy, EIMS, and elemental analysis. The molecular structures of 3 and 4 were unambiguously established by single‐crystal X‐ray structural analysis. The room‐temperature reaction of 4,4′‐bis(dimethylamino)thiobenzophenone with L′ Si resulted in products 4 and 5 , in which 4 is the dearomatized product and 5 is formed under the 1,3‐migration of a hydrogen atom from the aromatic phenyl ring to the carbon atom of the C? S unit. Furthermore, the optimized structures of probable products were investigated by using DFT calculations.  相似文献   

13.
Two new dithioether ligands, 1,4‐bis[(phenylsulfanyl)methyl]naphthalene ( L1 ), and 4,4′‐bis[(tert‐butylsulfanyl)methyl]biphenyl ( L2 ) were synthesized and their silver(I) complexes were studied. Both AgI complexes, [Ag L1 (NO3)]n ( 1 ) and [Ag L2 (NO3)]2 ( 2 ), were synthesized at ambient temperature and characterized by elemental analysis, IR spectroscopy, and single‐crystal X‐ray diffraction analysis. Single‐crystal X‐ray analysis shows that complex 1 has a one‐dimensional helical chain structure with the neutral repeating unit [Ag(μ2‐ L1 )(NO3)], whereas complex 2 has a centrosymmetrical neutral dinuclear structure. Moreover, complexes 1 and 2 are further extended into three‐dimensional supramolecular frameworks by hydrogen bonding and π–π stacking interactions, respectively. In addition, complexes 1 and 2 display strong blue emission in the solid state at room temperature.  相似文献   

14.
For the efficient synthesis of oligoribonucleotides by the 5′‐O‐(4,4′‐dimethoxytrityl) phosphoramidite approach, the 2′‐O‐[1‐(benzyloxy)ethyl]acetals 56 – 67 were investigated. Studies with the 2′‐O‐[1‐(benzyloxy)ethyl]‐5′‐O‐(dimethoxytrityl)ribonucleoside 3′‐phosphoramidites 56 – 59 gave, however, only reasonable results. The oligoribonucleotides obtained showed some impurities since the acid stabilities of the acetal and dimethoxytrityl functions are too close to guarantee a high selectivity. A combination of new acid‐labile protected 2′‐O‐protecting groups with the 2‐(4‐nitrophenyl)ethyl/[2‐(4‐nitrophenyl)ethoxy]carbonyl (npe/npeoc) strategy for base protection was more successful. The synthesis and physical properties of the monomeric building units and their intermediates 8 – 67 and the conditions for the automated generation of homo‐ and mixed oligoribonucleotides is described. The new 2′‐acetal protecting group could be cleaved off in a two step procedure and was designed for levelling their stability with regard to the attached nucleobase as well. Therefore, we used the 1‐{{3‐fluoro‐4‐{{[2‐(4‐nitrophenyl)ethoxy]carbonyl}oxy}benzyl}oxy}ethyl (fnebe) moiety for the protection of 2′‐OH of uridine, and for that of 2′‐OH of A, C, and G, the 1‐{{4‐{{[2‐(4‐nitrophenyl)ethoxy]carbonyl}oxy}benzyl}oxy}ethyl (nebe) residue. After selective deprotection by β‐elimination induced by a strong organic base like DBU, the remaining activated acetal was hydrolyzed under very mild acidic protic conditions, which reduced 2′‐3′ isomerization and chain cleavage. Also storage, handling, and purification of the chemically and enzymatically sensitive oligomers was simplified by this approach.  相似文献   

15.
(Z)‐1,2‐Diaryl‐1,2‐bis(pinacolatoboryl)ethenes underwent double‐cross‐coupling reactions with 1‐bromo‐2‐[(Z)‐2‐bromoethenyl]arenes in the presence of [Pd(PPh3)4] as a catalyst and 3 M aqueous Cs2CO3 as a base in THF at 80 °C. The double‐coupling reaction gave multisubstituted naphthalenes in good to high yields. Annulation of 1,2‐bis(pinacolatoboryl)arenes with bromo(bromoethenyl)arenes in the presence of a catalyst system that consisted of [Pd2(dba)3] (dba=dibenzylideneacetone) and 2‐dicyclohexylphosphino‐2′,6′‐dimethoxybiphenyl (SPhos) under the same conditions produced fused phenanthrenes in good to high yields. The first annulation coupling occurred regiospecifically at the bromoethenyl moiety. This procedure is applicable to the facile synthesis of polysubstituted anthracenes, benzothiophenes, and dibenzoanthracenes through a double annulation pathway by using the corresponding dibromobis[(Z)‐2‐bromoethenyl]benzenes as diboryl coupling partners.  相似文献   

16.
Four new 3,4‐dihydro‐1‐benzoxepin‐5(2H )‐one derivatives, namely (E )‐4‐(5‐bromo‐2‐hydroxybenzylidene)‐6,8‐dimethoxy‐3,4‐dihydrobenzo[b ]oxepin‐5(2H )‐one, ( 7 ), (E )‐4‐[(E )‐3‐(5‐bromo‐2‐hydroxyphenyl)allylidene]‐6,8‐dimethoxy‐3,4‐dihydrobenzo[b ]oxepin‐5(2H )‐one, ( 8 ), (E )‐4‐(5‐bromo‐2‐hydroxybenzylidene)‐6‐hydroxy‐8‐methoxy‐3,4‐dihydrobenzo[b ]oxepin‐5(2H )‐one, C18H15BrO5, ( 9 ), and (E )‐4‐[(E )‐3‐(5‐bromo‐2‐hydroxyphenyl)allylidene]‐6‐hydroxy‐8‐methoxy‐3,4‐dihydrobenzo[b ]oxepin‐5(2H )‐one, ( 10 ), have been synthesized and characterized by FT–IR, NMR and MS. The structure of ( 9 ) was confirmed by single‐crystal X‐ray diffraction. Crystal structure analysis shows that molecules of ( 9 ) are connected into a one‐dimensional chain in the [010] direction through classical hydrogen bonds and these chains are further extended into a three‐dimensional network via C—H…O interactions. The inhibitory activities of these compounds against protein–tyrosine kinases (PTKs) show that 6‐hydroxy‐substituted compounds ( 9 ) and ( 10 ) are more effective for inhibiting ErbB1 and ErbB2 than are 6‐methoxy‐substituted compounds ( 7 ) and ( 8 ). This may be because ( 9 ) and ( 10 ) could effectively bind to the active pockets of the protein through intermolecular interactions.  相似文献   

17.
The coordination chemistry of multinuclear metal compounds is important because of their relevance to the multi‐metal active sites of various metalloproteins and metalloenzymes. Multinuclear CuII and MnIII compounds are of interest due to their various properties in the fields of coordination chemistry, inorganic biochemistry, catalysis, and optical and magnetic materials. Oxygen‐bridged binuclear MnIII complexes generally exhibit antiferromagnetic interactions and a few examples of ferromagnetic interactions have also been reported. Binuclear CuII complexes are important due to the fact that they provide examples of the simplest case of magnetic interaction involving only two unpaired electrons. Two novel dioxygen‐bridged copper(II) and manganese(III) Schiff base complexes, namely bis(μ‐4‐bromo‐2‐{[(3‐oxidopropyl)imino]methyl}phenolato)dicopper(II), [Cu2(C10H10BrNO2)2], (1), and bis(diaqua{4,4′‐dichloro‐2,2′‐[(1,1‐dimethylethane‐1,2‐diyl)bis(nitrilomethanylylidene)]diphenolato}manganese(III)) bis{μ‐4,4′‐dichloro‐2,2′‐[(1,1‐dimethylethane‐1,2‐diyl)bis(nitrilomethanylylidene)]diphenolato}bis[aquamanganese(III)] tetrakis(perchlorate) ethanol disolvate, [Mn(C18H16Cl2N2O2)(H2O)2]2[Mn2(C18H16Cl2N2O2)2(H2O)2](ClO4)4·2C2H5OH, (2), have been synthesized and single‐crystal X‐ray diffraction has been used to analyze their crystal structures. The structure analyses of (1) and (2) show that each CuII atom is four‐coordinated, with long weak Cu…O interactions of 2.8631 (13) Å linking the dinuclear halves of the centrosymmetric tetranucelar molecules, while each MnIII atom is six‐coordinated. The shortest intra‐ and intermolecular nonbonding Mn…Mn separations are 3.3277 (16) and 5.1763 (19) Å for (2), while the Cu…Cu separations are 3.0237 (3) and 3.4846 (3) Å for (1). The magnetic susceptibilities of (1) and (2) in the solid state were measured in the temperature range 2–300 K and reveal the presence of antiferromagnetic spin‐exchange interactions between the transition metal ions.  相似文献   

18.
Three new heteroscorpionate ligands, (2‐hydroxyphenyl)bis(imidazol‐1‐yl)methane (HL1), (4‐diethylamino‐2‐hydroxyphenyl)bis(imidazol‐1‐yl)methane (HL2) and (5‐bromo‐2‐hydroxyphenyl)bis(imidazol‐1‐yl)methane (HL3), and their heteroleptic copper(II) complexes of the type [Cu(L1–3)diimine]ClO4 ( 1 – 6 ; where diimine =2,2′‐bipyridyl or 1,10‐phenanthroline) have been synthesized and characterized using spectroscopic methods. The molecular structure of ligand HL1 was determined by single‐crystal X‐ray diffraction. UV–visible, electron paramagnetic resonance and theoretical studies suggest a distorted square pyramidal geometry around copper(II) ion. Analyses of highest occupied and lowest unoccupied molecular orbitals have been used to explain the charge transfer taking place within the complexes. The antioxidant activities of the heteroscorpionate ligands and their heteroleptic copper(II) complexes were determined using ABTS, DPPH and H2O2 free radical scavenging assays with respect to standard antioxidant ascorbic acid. In molecular docking studies, the complexes showed π–π, hydrogen bonding, van der Waals and electrostatic interactions with fibroblast growth factor receptor kinase. In vitro cytotoxicity activities of ligands and copper(II) complexes were examined on human breast adenocarcinoma (MCF‐7), cervical (HeLa) and lung (A549) cancer cell lines and normal human dermal fibroblast cell line using MTT assay. Complex 4 exhibited higher anticancer activity than the other complexes against all three cancer lines, being more potent than the standard drug cisplatin.  相似文献   

19.
First examples of transition metal complexes with HpicOH [Cu(picOH)2(H2O)2] ( 1 ), [Cu(picO)(2,2′‐bpy)]·2H2O ( 2 ), [Cu(picO)(4,4′‐bpy)0.5(H2O)]n ( 3 ), and [Cu(picO)(bpe)0.5(H2O)]n ( 4 ) (HpicOH = 6‐hydroxy‐picolinic acid; 2,2′‐bpy = 2,2′‐bipyridine; 4,4′‐bpy = 4,4′‐bipyridine; bpe = 1,2‐bis(4‐pyridyl)ethane) have been synthesized and characterized by single‐crystal X‐ray diffraction. The results show that HpicOH ligand can be in the enol or ketonic form, and adopts different coordination modes under different pH value of the reaction mixture. In complex 1 , HpicOH ligand is in the enol form and adopts a bidentate mode. While in complexes 2 – 4 , as the pH rises, HpicOH ligand becomes in the ketonic form and adopts a tridentate mode. The coordination modes in complexes 1 – 4 have not been reported before. Because of the introduction of the terminal ligands 2,2′‐bpy, complex 2 is of binuclear species; whereas in complexes 3 and 4 , picO ligands together with bridging ligands 4,4′‐bpy and bpe connect CuII ions to form 2D nets with (123)2(12)3 topology.  相似文献   

20.
Abstract

The effects of two metal complexes of 2,2′‐dipyridylamine (bpya) ligand, [(bpya)Cu]Cl2 and [(bpya)Zn]Cl2, in promoting the hydrolysis of bis(4‐nitrophenyl) phosphate (BNPP) have been kinetically investigated in Brij35 micellar solution and at 298 K, pH ranging from 6.41 to 8.6. In neutral micellar solution at 298 K, pH 7.02, the rate constants for the catalytic hydrolysis of BNPP by [(bpya)Cu]Cl2 and [(bpya)Zn]Cl2 are 1.2 × 106 times and 1.5 × 105 times higher than those for the spontaneous hydrolysis, respectively. Kinetic studies show that the active species in the catalytic hydrolysis of BNPP is the aquo‐hydroxy form, and the relative kinetic and thermodynamic parameters indicate that the mechanism of the reaction involves intramolecular nucleophilic attack on the metal center‐bound diester.  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