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 Rose Bengal‐sensitized photooxidations of the dipeptides l ‐tryptophyl‐l ‐phenylalanine (Trp‐Phe), l ‐tryptophyl‐l ‐tyrosine (Trp‐Tyr) and l ‐tryptophyl‐l ‐tryptophan (Trp‐Trp) have been studied in pH 7 water solution using static photolysis and time‐resolved methods. Kinetic results indicate that the tryptophan (Trp) moiety interacts with singlet molecular oxygen (O2(1Δg)) both through chemical reaction and through physical quenching, and that the photooxidations can be compared with those of equimolecular mixtures of the corresponding free amino acids, with minimum, if any, influence of the peptide bond on the chemical reaction. This is not a common behavior in other di‐ and polypeptides of photooxidizable amino acids. The ratio between chemical (kr) and overall (kt) rate constants for the interaction O2(1Δg)‐dipeptide indicates that Trp‐Phe and Trp‐Trp are good candidates to suffer photodynamic action, with krlkt values of 0.72 and 0.60, respectively (0.65 for free Trp). In the case of Trp‐Tyr, a lower krlkt value (0.18) has been found, likely as a result of the high component of physical deactivation of O2(1Δg) by the tyrosine moiety. The analysis of the photooxidation products shows that the main target for O2(1Δg) attack is the Trp group and suggests a much lower accumulation of kynurenine‐type products, as compared with free Trp. This is possibly because of the occurrence of another accepted alternative pathway of oxidation that gives rise to 3a‐oxidized hydrogenated pyrrolo[2,3‐b]indoles.  相似文献   

2.
Three novel copper(II) complexes, [Cu(Gly‐l ‐Val)(HPBM)(H2O)]·ClO4·H2O ( 1 ), [Cu(Gly‐l ‐Val)(TBZ)(H2O)]·ClO4 ( 2 ) and [Cu(Gly‐l ‐Val)(PBO)(H2O)]·ClO4 ( 3 ) (Gly‐l ‐Val = glycyl‐l ‐valine anion, HPBM = 5‐methyl‐2‐(2′‐pyridyl)benzimidazole, TBZ = 2‐(4′‐thiazolyl)benzimidazole, PBO = 2‐(2′‐pyridyl)benzoxazole), have been prepared and characterized with elemental analyses, conductivity measurements as well as various spectroscopic techniques. The interactions of these copper complexes with calf thymus DNA were explored using UV–visible, fluorescence, circular dichroism, thermal denaturation, viscosity and docking analyses methods. The experimental results showed that all three complexes could bind to DNA via an intercalative mode. Moreover, the cytotoxic effects were evaluated using the MTT method, and the antimicrobial activity of these complexes was tested against Bacillus subtilis, Staphylococcus aureus, Escherichia coli and Pseudomonas aeruginosa. The results showed that the activities are consistent with their DNA binding abilities, following the order of 1 > 2 > 3 .  相似文献   

3.
α‐l ‐Fucosidases (EC 3.2.1.51) are exo‐glycosidases. On the basis of the multi‐alignment of amino acid sequence, α‐l ‐fucosidases were classified into two families of glycoside hydrolases, GH‐29 and GH‐95. They are responsible for the removal of l ‐fucosyl residues from the non‐reducing end of glycoconjugates. Deficiency of α‐l ‐fucosidase results in Fucosidosis due to the accumulation of fucose‐containing glycolipids, glycoproteins and oligosaccharides in various tissues. Recent studies discovered that the fucosylation levels are increased on the membrane surfaces of many carcinomas, indicating the biological function of α‐l ‐fucosidases may relate to this abnormal cell physiology. Although the gene of human α‐l ‐fucosidase (h‐fuc) was cloned, the recombinant enzyme has rarely been overexpressed as a soluble and active from. We report herein that, with carefully control on the growing condition, an active human α‐l ‐fucosidases (h‐Fuc) was successfully expressed in Escherichia coli for the first time. After a series steps of ion‐exchange and gel‐filtration chromatographic purification, the recombinant h‐Fuc with 95% homogeneity was obtained. The molecular weight of the enzyme was analyzed by SDS‐PAGE (~50 kDa) and confirmed by ESI mass (50895 Da). The recombinant h‐Fuc was stable up to 55 °C with incubation at pH 6.8 for 2 h; the optimum temperature for h‐Fuc is approximately 55 °C. The enzyme was stable at pH 2.5–7.0 for 2 h; the enzyme activity decreased greatly for pH greater than 8.0 or less than 2.0. The Km and kcat values of the recombinant h‐Fuc (at pH 6.8) were determined to be 0.28 mM and 17.1 s?1, respectively. The study of pH‐dependent activity showed that the recombinant enzyme exhibited optimum activity at two regions near at pH 4.5 and pH 6.5. These features of the recombinant h‐Fuc are comparable to the native enzyme purified directly from human liver. Studies on the transfucosylation and common intermediate of the enzymatic reaction by NMR support that h‐Fuc functions as a retaining enzyme catalyzing the hydrolysis of substrate via a two‐step, double displacement mechanism.  相似文献   

4.
Macroinitiator‐amino terminated poly(ethylene glycol) (PEG) (NH2‐PEO‐NH2) was prepared by converting both terminal hydroxyl groups of PEG to more reactive primary amino groups. The synthetic route involved reactions of chloridize, phthalimide and finally hydrazinolysis. Furthermore, poly(γ‐benzyl‐L ‐glutamate)‐poly(ethylene oxide)‐poly(γ‐benzyl‐L ‐glutamate) (PBLG‐PEO‐PBLG) triblock copolymer was synthesized by polymerization of γ‐benzyl‐L ‐glutamate N‐carboxyanhydride (Bz‐L‐GluNCA) using NH2‐PEO‐NH2 as macroinitiator. The resultant NH2‐PEO‐NH2 and triblock copolymer were characterized by FT‐IR, 1H‐NMR and gel permeation chromatography (GPC) techniques. The results demonstrated that the degree of amination of the NH2‐PEO‐NH2 could be up to 1.95. The molecular weight of the PBLG‐PEO‐PBLG triblock copolymer could be adjusted easily by controlling the molar ratio of Bz‐L ‐Glu NCA to the macroinitiator NH2‐PEO‐NH2. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

5.
A novel chromogenic method to measure the peroxidase activity using para‐phenylenediamine dihydrochloride (=benzene‐1,4‐diamine hydrochloride; PPDD) and N‐(1‐naphthyl)ethylenediamine dihydrochloride (=N‐(2‐aminoethyl)naphthalen‐1‐amine; NEDA) is presented. The PPDD entraps the free radical and gets oxidized to electrophilic diimine, which couples with NEDA to give an intense red‐colored chromogenic species with maximum absorbance at 490 nm. This assay was adopted for the quantification of H2O2 between 20 and 160 μM . Catalytic efficiency and catalytic power of the commercial peroxidase were found to be 4.47×104 M ?1 min?1 and 3.38×10?4 min?1, respectively. The catalytic constant (kcat) and specificity constant (kcat/Km) at saturated concentration of the co‐substrates were 0.0245×103 min?1 and 0.0445 μM ?1 min?1, respectively. The chromogenic coupling reaction has a minimum interference from the reducing substances such as ascorbic acid, L ‐cystein, citric acid, and oxalic acid. The method being simple, rapid, precise, and sensitive, its applicability has been tested in the crude vegetable extracts that showed peroxidase activity.  相似文献   

6.
Copper(I) halides with triphenyl phosphine and imidaozlidine‐2‐thiones (L ‐NMe, L ‐NEt, and L ‐NPh) in acetonitrile/methanol (or dichloromethane) yielded copper(I) mixed‐ligand complexes: mononuclear, namely, [CuCl(κ1‐S‐L ‐NMe)(PPh3)2] ( 1 ), [CuBr(κ1‐S‐L ‐NMe)(PPh3)2] ( 2 ), [CuBr(κ1‐S‐L ‐NEt)(PPh3)2] ( 5 ), [CuI(κ1‐S‐L ‐NEt)(PPh3)2] ( 6 ), [CuCl(κ1‐S‐L ‐NPh)(PPh3)2] ( 7 ), and [CuBr(κ1‐S‐L ‐NPh)(PPh3)2] ( 8 ), and dinuclear, [Cu21‐I)2(μ‐S‐L ‐NMe)2(PPh3)2] ( 3 ) and [Cu2(μ‐Cl)21‐S‐L ‐NEt)2(PPh3)2] ( 4 ). All complexes were characterized with analytical data, IR and NMR spectroscopy, and X‐ray crystallography. Complexes 2 – 4 , 7 , and 8 each formed crystals in the triclinic system with P$\bar{1}$ space group, whereas complexes 1 , 5 , and 6 crystallized in the monoclinic crystal system with space groups P21/c, C2/c, and P21/n, respectively. Complex 2 has shown two independent molecules, [(CuBr(κ1‐S‐L ‐NMe)(PPh3)2] and [CuBr(PPh3)2] in the unit cell. For X = Cl, the thio‐ligand bonded to metal as terminal in complex 4 , whereas for X = I it is sulfur‐bridged in complex 3 .  相似文献   

7.
A stereospecific synthesis of (2S)3‐(2,4,5‐trifluorophenyl)propane‐1,2‐diol from D ‐mannitol has been developed. The reaction of 2,3‐O‐isopropylidene‐D ‐glyceraldehyde with 2,4,5‐trifluorophenylmagnesium bromide gave [(4R)‐2,2‐dimethyl‐1,3‐dioxolan‐4‐yl](2,4,5‐trifluorophenyl)methanol in 65% yield as a mixture of diastereoisomers (1 : 1). The Ph3P catalyzed reaction of the latter with C2Cl6 followed by reduction with Pd/C‐catalyzed hydrogenation gave (2S)‐3‐(2,4,5‐trifluorophenyl)propane‐1,2‐diol with >99% ee and 65% yield.  相似文献   

8.
Enantioresolution of the calcimimetic drug (R,S)‐Cinacalcet was achieved using both indirect and direct approaches. Six chiral variants of Marfey's reagent having l ‐Ala‐NH2, l ‐Phe‐NH2, l ‐Val‐NH2, l ‐Leu‐NH2, l ‐Met‐NH2 and d ‐Phg‐NH2 as chiral auxiliaries were used as derivatizing reagents under microwave irradiation. Derivatization conditions were optimized. Reversed‐phase high‐performance liquid chromatography was successful using binary mixtures of aqueous trifluoroacetic acid and acetonitrile for separation of diastereomeric pairs with detection at 340 nm. Thin silica gel layers impregnated with optically pure l ‐histidine and l ‐arginine were used for direct resolution of enantiomers. The limit of detection was found to be 60 pmol in HPLC while in TLC it was found to be in the range of 0.26–0.28 µg for each enantiomers. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

9.
In the structure of the 1:1 proton‐transfer compound from the reaction of l ‐tartaric acid with the azo‐dye precursor aniline yellow [4‐(phenyldiazenyl)aniline], namely 4‐(phenyldiazenyl)anilinium (2R,3R)‐3‐carboxy‐2,3‐dihydroxypropanoate, C12H12N3+·C4H5O6, the asymmetric unit contains two independent 4‐(phenyldiazenyl)anilinium cations and two hydrogen l ‐tartrate anions. The structure is unusual in that all four phenyl rings of the two cations have identical rotational disorder with equal occupancy of the conformations. The two hydrogen l ‐tartrate anions form independent but similar chains through head‐to‐tail carboxyl–carboxylate O—H...O hydrogen bonds [graph set C(7)], which are then extended into a two‐dimensional hydrogen‐bonded sheet structure through hydroxy O—H...O hydrogen‐bonded links. The anilinium groups of the 4‐(phenyldiazenyl)anilinium cations are incorporated into the sheets and also provide internal hydrogen‐bonded extensions, while their aromatic tails are layered in the structure without significant association except for weak π–π interactions [minimum ring centroid separation = 3.844 (3) Å]. The hydrogen l ‐tartrate residues of both anions exhibit the common short intramolecular hydroxy–carboxylate O—H...O hydogen bonds. This work provides a solution to the unusual disorder problem inherent in the structure of this salt, as well as giving another example of the utility of the hydrogen tartrate anion in the generation of sheet substructures in molecular assembly processes.  相似文献   

10.
Substituted phenyl‐N‐butyl carbamates ( 1 ) and p‐nitrophenyl‐N‐substituted carbamates ( 2 ) are characterized as “pseudo‐pseudo‐substrate” inhibitors of acetylcholinesterase. Since the inhibitors protonate in pH 7.0 buffer solution, the virtual inhibition constants (Ki's) of the protonated inhibitors can be calculated from the equation, ‐logKi' = ‐logKi ‐ pKa + 14. The ‐logKi' and logkc values for acetylcholinesterase inhibitions by carbamates 1 correlate with the Hammett equation (log(k/k0) = ρσ); moreover, those by carbamates 2 correlate with the Taft equation (log(k/k0) = ρ* σ*). With modified Hammett‐Taft cross‐interaction variations, multiple linear regressions of the ‐logKi' and logkc values of carbamates 1 and 2 give good correlations, and the cross‐interaction constants (ρXR) are 0.5 and 0.0, respectively. The ρXR value of 0.5 indicates that the carbamate O‐C(O)‐N‐R geometries for the transition states that lead to enzyme‐carbamate tetrahedral intermediates are all pseudo‐trans conformations. Therefore, the carbamate moiety of the inhibitors stretches along the active site gorge of the enzyme but does not bind in the acyl binding site pocket of the enzyme. Overall, the carbamate O‐C(O)‐N‐R geometries for carbamates 1 and 2 , protonated carbamates 1 and 2 , and the tetrahedral intermediate are all retained in pseudo‐trans conformations. The ρXR value of 0.0 suggests that the transition states that lead to the carbamyl enzymes are breaking C‐O bonds and are excluding the leaving groups, substituted phenols.  相似文献   

11.
Two water‐soluble 6‐(pyrazin‐2‐yl)‐1,3,5‐triazine‐2,4‐diamino (pzta)‐based Cu(II) complexes, namely [Cu(l ‐Val)(pzta)(H2O)]ClO4 ( 1 ) and [Cu(l ‐Thr)(pzta)(H2O)]ClO4 ( 2 ) (l ‐Val: l ‐valinate; l ‐Thr: l ‐threoninate), were synthesized and characterized using elemental analyses, molar conductance measurements, spectroscopic methods and single‐crystal X‐ray diffraction. The results indicated that the molecular structures of the complexes are five‐coordinated and show a distorted square‐pyramidal geometry, in which the central copper ions are coordinated to N,N atoms of pzta and N,O atoms of amino acids. The interactions of the complexes with DNA were investigated using electronic absorption, competitive fluorescence titration, circular dichroism and viscosity measurements. These studies confirmed that the complexes bind to DNA through a groove binding mode with certain affinities (Kb = 4.71 × 103 and 1.98 × 103 M?1 for 1 and 2 , respectively). The human serum albumin (HSA) binding properties of the complexes were also evaluated using fluorescence and synchronous fluorescence spectroscopies, indicating that the complexes could quench the intrinsic fluorescence of HSA in a static quenching process. The relevant thermodynamic parameters revealed the involvement of van der Waals forces and hydrogen bonds in the formation of complex–HSA systems. Finally, molecular docking technology was also used to further verify the interactions of the complexes with DNA/HSA.  相似文献   

12.
Comprehensive mechanistic studies on the enantioselective aldol reaction between isatin ( 1 a ) and acetone, catalyzed by L ‐leucinol ( 3 a ), unraveled that isatin, apart from being a substrate, also plays an active catalytic role. Conversion of the intermediate oxazolidine 4 into the reactive syn‐enamine 6 , catalyzed by isatin, was identified as the rate‐determining step by both the calculations (ΔG=26.1 kcal mol?1 for the analogous L ‐alaninol, 3 b ) and the kinetic isotope effect (kH/kD=2.7 observed for the reaction using [D6]acetone). The subsequent reaction of the syn‐enamine 6 with isatin produces (S)‐ 2 a (calculated ΔG=11.6 kcal mol?1). The calculations suggest that the overall stereochemistry is controlled by two key events: 1) the isatin‐catalyzed formation of the syn‐enamine 6 , which is thermodynamically favored over its anti‐rotamer 7 by 2.3 kcal mol?1; and 2) the high preference of the syn‐enamine 6 to produce (S)‐ 2 a on reaction with isatin ( 1 a ) rather than its enantiomer (ΔΔG=2.6 kcal mol?1).  相似文献   

13.
The polymers with functionalized alkoxy groups and with narrow molecular weight distribution (Mw/Mn < 1.12) are obtained from the living polymerization of 2‐alkoxy‐1‐methylenecyclopropanes using π‐allylpalladium complex, [(PhC3H4)Pd(μ‐Cl)]2, as the initiator. The polymers with oligoethylene glycol groups in the alkoxy substituent are soluble in water, and hydroboration of the C?C double bond and ensuing addition of the OH groups to C?N bond of alkyl isocyanate produce the polymers with urethane pendant groups. The reaction decreases solubility of the polymer in water significantly. Di‐ and triblock copolymers of the 2‐alkoxy‐1‐methylenecyclopropanes are prepared by consecutive addition of the two or three 2‐alkoxy‐1‐methylenecyclopropane monomers to the Pd initiator. The polymers which contain both hydrophobic butoxy or tert‐butoxy group and hydrophilic oligoethylene glycol group dissolve in water and/or organic solvents, depending on the substituents. The 1H NMR spectrum of poly( 1a ‐b‐ 1h ) (? (CH2C(?CH2)CHOBu)n? (CH2C(?CH2)CH(OCH2CH2)3OMe)m? ) in D2O solution exhibits peaks because of the butoxy and ?CH2 hydrogen in decreased intensity, indicating that the polymer forms micelle particles containing the hydrophilic segments in their external parts. Aqueous solution of the polymer with a small amount of DPH (DPH = 1,6‐diphenyl‐1,3,5‐hexatriene) shows the absorbance due to DPH at concentration of the polymer higher than 5.82 × 10?5 g mL?1. Other block copolymers such as poly( 1b ‐b‐ 1h ) and poly( 1a ‐b‐ 1g ) also form the micelles that contain DPH in their core. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 959–972, 2009  相似文献   

14.
The radical S‐adenosyl‐l ‐methionine (SAM) enzyme NosL catalyzes the transformation of l ‐tryptophan into 3‐methyl‐2‐indolic acid (MIA), which is a key intermediate in the biosynthesis of a clinically interesting antibiotic nosiheptide. NosL catalysis was investigated by using the substrate analogue 2‐methyl‐3‐(indol‐3‐yl)propanoic acid (MIPA), which can be converted into MIA by NosL. Biochemical assays with different MIPA isotopomers in D2O and H2O unambiguously indicated that the 5′‐deoxyadenosyl (dAdo)‐radical‐mediated hydrogen abstraction is from the amino group of l ‐tryptophan and not a protein residue. Surprisingly, the dAdo‐radical‐mediated hydrogen abstraction occurs at two different sites of MIPA, thereby partitioning the substrate into different reaction pathways. Together with identification of an α,β‐unsaturated ketone shunt product, our study provides valuable mechanistic insight into NosL catalysis and highlights the remarkable catalytic flexibility of radical SAM enzymes.  相似文献   

15.
1,1‐Diethyl‐1‐germa‐2,3,4,5‐tetra‐ tert ‐butyl‐2,3,4,5‐tetraphospholane (C2H5)2Ge( t BuP)4, Molecular and Crystal Structure The reaction of the diphosphide K2[(tBuP)4] · THF ( 1 ) with the germanium(IV) compound (C2H5)2GeCl2 leads via a [4 + 1]‐cyclo‐condensation reaction to 1,1‐diethyl‐1‐germa‐2,3,4,5‐tetra‐tert‐butyl‐2,3,4,5‐tetraphospholane (C2H5)2Ge(tBuP)4 ( 2 ) with the 5‐membered GeP4 ring system. 2 could be characterized 31P NMR spectroscopically, mass spectrometrically and by a single crystal structure analysis.  相似文献   

16.
In this work, rare earth tris(borohydride) complexes, Ln(BH4)3(THF)3 (Ln = Sc, Y, La, and Dy), have been used to catalyze the ring‐opening polymerization of γ‐benzyl‐L ‐glutamate N‐carboxyanhydride (BLG NCA). All the catalysts show high activities and the resulting poly(γ‐benzyl‐L ‐glutamate)s (PBLGs) are recovered with high yields (≥90%). The molecular weights (MWs) of PBLG can be controlled by the molar ratios of monomer to catalyst, and the MW distributions (MWDs) are relatively narrow (as low as 1.16) depending on the rare earth metals and reaction temperatures. Block copolypeptides can be easily synthesized by the sequential addition of two monomers. The obtained P(γ‐benzyl‐L ‐glutamate‐b‐ε‐carbobenzoxy‐L ‐lysine) [P(BLG‐b‐BLL)] and P(γ‐benzyl‐L ‐glutamate‐b‐alanine) [P(BLG‐b‐ALA)] have been well characterized by NMR, gel permeation chromatography, and differential scanning calorimetry measurements. A random copolymer P(BLG‐co‐BLL) with a narrow MWD of 1.07 has also been synthesized. The polymerization mechanisms have been investigated in detail. The results show that both nucleophilic attack at the 5‐CO of NCA and deprotonation of 3‐NH of NCA in the initiation process take place simultaneously, resulting in two active centers, that is, an yttrium ALA carbamate derivative [H2BOCH2(CH)NHC(O)OLn? ] and a N‐yttriumlated ALA NCA. Propagation then proceeds on these centers via both normal monomer insertion and polycondensation. After termination, two kinds of telechelic polypeptide chains, that is, α‐hydroxyl‐ω‐aminotelechelic chains and α‐carboxylic‐ω‐aminotelechelic ones, are formed as characterized by MALDI‐TOF MS, 1H NMR, 13C NMR, 1H–1H COSY, and 1H–13C HMQC measurements. By decreasing the reaction temperature, the normal monomer insertion pathway can be exclusively selected, forming an unprecedented α‐hydroxyl‐ω‐aminotelechelic polypeptide. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

17.
A series of l ‐amino acid‐modified benzimidazoles have been synthesized and their application in Ru‐catalysed asymmetric transfer hydrogenation (ATH) of ketones has been evaluated using i‐propanol as hydrogen donor as well as solvent. The enantioselectivity of the product was sensitive to the reaction conditions and showed reversed temperature effect. Better reaction activity and enantioselectivity were obtained at higher reaction temperature and higher molar ratio of substrate to catalyst (S/C). At an S/C of 2000:1 and 5000:1, the turnover frequency (TOF) and the enantiomeric excess (ee) value of the ATH of acetophenone promoted by the combination of l ‐proline benzimidazole 7c and [RuCl2(p‐cymene)]2 were up to 6880, 13 000 h?1 and 67%, 51% respectively. The ee value dropped off after the reaction reached equilibrium. The high activity and enantioselectivity of [RuCl2(p‐cymene)]2/ 7c was primarily influenced by the (S,R) configuration, whereas the (S,S) configuration had low activity and enantioselectivity. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

18.
In the presence or absence of water, a Schiff‐base compound, 4‐amino‐3‐(2‐(2‐hydroxybenzylidene)hydrazinyl)‐1H‐1,2,4‐triazole‐5(4H)‐thione ( HATT ), forms different crystalline states ( HATT , HATT ? 2 H2O, and a lamellar structure, m‐HATT ? n H2O), which show different luminescence emission properties. Herein, we investigate the emission of HATT and the role of water molecules. A water molecule, which acts as both a hydrogen‐bond acceptor and ‐donor, enlarges the distance between adjacent HATT molecules and hinders non‐radiative decay pathways.  相似文献   

19.
Di(tert‐butyl)diazomethane ( 4 ) is a nucleophilic 1,3‐dipole with strong steric hindrance at one terminus. In its reaction with 2,3‐bis(trifluoromethyl)fumaronitrile ((E)‐ BTE ), a highly electrophilic tetra‐acceptor‐substituted ethene, an imino‐substituted cyclopentene 9 is formed as a 1 : 2 product. The open‐chain zwitterion 10 , assumed as intermediate, adds the second molecule of (E)‐ BTE . The 19F‐ and 13C‐NMR spectra allow the structural assignment of two diastereoisomers, 9A and 9B . The zwitterion 10 can also be intercepted by dimethyl 2,3‐dicyanofumarate ( 11 ) and furnishes diastereoisomeric cyclopentenes 12A and 12B ; an X‐ray‐analysis of 12B confirms the ‘mixed’ 1 : 1 : 1 product. Competing is an (E)‐ BTE ‐catalyzed decomposition of 4 to give 2,3,4,4‐tetramethylpent‐1‐ene ( 7 )+N2; the reaction of (E)‐ BTE with a trace of water appears to be responsible for the chain initiation. The H2SO4‐catalyzed decomposition of diazoalkane 4 , indeed, produced the alkene 7 in high yield. The attack on the hindered diazoalkane 4 by 11 is slower than that by (E)‐ BTE ; the zwitterionic intermediate 21 undergoes cyclization and furnishes the tetrasubstituted furan 22 . In fumaronitrile, electrophilicity and steric demand are diminished, and a 1,3‐cycloaddition produces the 4,5‐dihydro‐1H‐pyrazole derivative 25 . The reaction of 4 with dimethyl acetylenedicarboxylate leads to pyrazole 29 +isobutene.  相似文献   

20.
In situ functionalization of a 4‐aminothiophenol (4ATP) self‐assembled monolayer (SAM) on a Au electrode (4ATP/Au SAM) by the Michael addition reaction is considered. Under optimized conditions, the nucleophilic attack of the amino group of 4ATP/Au SAM to give an electrogenerated ortho‐quinone produced a novel electroactive SAM (ESAM). The ESAM could be oxidized to quinone‐imine SAM (QI SAM) for the covalent immobilization of L ‐arginine monolayers. Cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS) and Fourier transform infrared (FTIR) spectroscopy are employed to characterize these systems.. The apparent heterogeneous rate constant (ksapp) for ESAM/Au and the rate constant (k′) of the pseudo‐first order Michael addition reaction of L ‐arginine and ESAM/Au are calculated.  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