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 593 毫秒
1.
Among catalytic antibodies, the well-characterized antibody 43C9 is unique in its ability to catalyze the difficult, but desirable, reaction of selective amide hydrolysis. The crystallographic structures that we present here for the single-chain variable fragment of the 43C9 antibody, both with and without the bound product p -nitrophenol, strongly support and extend the structural and mechanistic information previously provided by a three-dimensional computational model, together with extensive biochemical, kinetics, and mutagenesis results. The structures reveal an unexpected extended beta-sheet conformation of the third complementarity determining region of the heavy chain, which may be coupled to the novel indole ring orientation of the adjacent Trp H103. This unusual conformation creates an antigen-binding site that is significantly deeper than predicted in the computational model, with a hydrophobic pocket that encloses the p -nitrophenol product. Despite these differences, the previously proposed roles for Arg L96 in transition-state stabilization and for His L91 as the nucleophile that forms a covalent acyl-antibody intermediate are fully supported by the crystallographic structures. His L91 is now centered at the bottom of the antigen-binding site with the imidazole ring poised for nucleophilic attack. His L91, Arg L96, and the bound p -nitrophenol are linked into a hydrogen-bonding network by two well-ordered water molecules. These water molecules may mimic the positions of the phosphonamidate oxygen atoms of the antigen, which in turn mimic the transition state of the reaction. This network also contains His H35, suggesting that this residue may also stabilize the transition-states. A possible proton-transfer pathway from His L91 through two tyrosine residues may assist nucleophilic attack. Although transition-state stabilization is commonly observed in esterolytic antibodies, nucleophilic attack appears to be unique to 43C9 and accounts for the unusually high catalytic activity of this antibody.  相似文献   

2.
Tang W  Ng SC 《Nature protocols》2007,2(12):3195-3200
We describe a protocol for the synthesis of mono-6(A)-(1-butyl-3-imidazolium)-6(A)-deoxy-beta-cyclodextrin chloride (BIMCD), a cationic, water-soluble cyclodextrin used in the chiral separation of amino acids and anionic pharmaceuticals by capillary electrophoresis. Starting from commercially available chemicals, BIMCD is synthesized in five steps. The first step involves a nucleophilic substitution between p-toluenesulfonyl chloride and imidazole to afford 1-(p-toluenesulfonyl)imidazole (A). In the second step, a nucleophilic substitution between beta-cyclodextrin and A affords mono-6(A)-(p-toluenesulfonyl)-6(A)-deoxy-beta-cyclodextrin (B). In the third step, a nucleophilic substitution between 1-bromobutane and imidazole affords 1-butylimidazole (C). In the fourth step, a nucleophilic addition between A and C affords BIMCD tosylate. In the final step, anion exchange using an ion-exchange resin yields BIMCD as a highly water-soluble solid. Each step takes up to 2 d, including the time required for product purification. The overall protocol requires approximately 6 d.  相似文献   

3.
The nucleophilic addition of GSH to quinonoid compounds, characterized as a 1,4-reductive addition of the Michael type, was studied with p-benzoquinone- and 1,4-naphthoquinone epoxides with different degree of methyl substitution. Identification and evaluation of molecular products from the above reaction were assessed by h.p.l.c. with either reductive or oxidative electrochemical detection, based on the redox properties retained in the molecular products formed. It was found that the degree of methyl substitution of the quinone epoxide, from either the 1,4-naphthoquinone- or p-benzoquinone epoxide series, determined their rate of reaction with GSH. The reductive addition implied the rearrangement of the quinone structure with opening of the epoxide ring yielding as the primary product a hydroxy-glutathionyl substituted adduct of either p-benzohydroquinone or 1,4-naphthohydroquinone. The primary product undergoes elimination reactions and redox transitions which bring about a number of secondary molecular products. The distribution pattern of the latter depends on the degree of methyl substitution of the quinone epoxide studied and on the concentration of O2 in the solution. The occurrence of the hydroxy-substituent in position alpha, adjacent to the carbonyl group, enhances the autoxidation properties of the compound resulting in an augmented O2 consumption and H2O2 production. Therefore, it could be expected that the chemical reactivity of the products originating from the thiol-mediated nucleophilic addition to quinone epoxides would be of toxicological interest.  相似文献   

4.
6,7-Dichloroquinoline-5,8-dione (1) was reacted with a number of 2-aminopyridine derivatives. Of the several possible products of this reaction, 4a,10,11-triazabenzo[3,2-a]fluorene-5,6-dione (6), produced by condensation and rearrangement, was obtained as the major product, and its structure was subsequently unambigously determined by X-ray crystallographic study. Ortho-quinones were produced via nucleophilic substitution at position C7, which was unexpected, considering that para-quinones were produced via C6 substitution in the reaction between compound 1 and ethyl acetoacetate in our previous work. Such unexpected nucleophilic substitution at C7 provides an effective, yet simple route, to the preparation of biologically active ortho-quinone derivatives.  相似文献   

5.
Understanding of the atomic movements involved in an enzymatic reaction needs structural information on the active and inactive native enzyme molecules and on the enzyme-substrate, enzyme-intermediate, and enzyme-product(s) complexes. By using the X-ray crystallographic method, four crystal structures of Aspergillus fumigatus phytase were obtained at resolution higher than 1.7 A. The pH-dependent catalytic activity of A. fumigatus phytase was linked to three water molecules that may prevent the substrate from binding and thus block nucleophilic attack of the catalytic imidazole nitrogen. Comparison of various structures also identified the water molecule that attacks the phosphamide bond during the hydrolysis process, and established the hydrolysis pathway of the intermediate. Additionally, two reaction product phosphates were observed at the active site, suggesting a possible product release pathway after hydrolysis of the intermediate. These results can help explain the catalytic mechanism throughout the whole acid phosphatase family, as all key residues are conserved.  相似文献   

6.
Shen K. Yang 《Chirality》1994,6(3):175-184
Enantiomers of 3-O-methyloxazepam (MeOX) and 3-O-ethyloxazepam (EtOX) were resolved by chiral stationary phase high-performance liquid chromatography (CSP-HPLC). Reaction kinetics and deuterium isotope effects of acid-catalyzed racemization of enantiomeric MeOX in ethanol and enantiomeric EtOX in methanol were studied by spectropolarimetry. The acid-catalyzed heteronucleophilic substitution reactions of racemic MeOX in ethanol and racemic EtOX in methanol were studied by reversed-phase HPLC. Thermodynamic parameters involved in the reactions were obtained by temperature-dependent reaction rates. The effects of solvent's dielectric constant on the heteronucleophilic substitution reactions were also determined. A nucleophilically solvated and transient C3 carbocation intermediate resulting from an N4-protonated enantiomer, derived from a 1,4-benzodiazcpine either in M (minus) or P (plus) conformation, is proposed to be an intermediate and responsible for the acid-catalyzed stereoselective nucleophilic substitution and the resulting racemization. © 1994 Wiley-Liss, Inc.  相似文献   

7.
Wong FF  Chen CY  Chen TH  Huang JJ  Fang HP  Yeh MY 《Steroids》2006,71(1):77-82
The synthesis of biologically active 3alpha-hydroxyl-21-(1'-imidazolyl)-3beta-methoxymethyl-5alpha-pregnan-20-one was accomplished in six steps. The key steps were the improvement of stereoselectivity for acetyl isomers in C-17 and the introduction of imidazole into the core structure by use of lithium imidazole. This latter key step provided the desired product in 82% yield without the formation of 1,3-disubstituted imidazolium salt as impurity, which is generally observed in traditional method.  相似文献   

8.
In contrast to the one-pot, two step syntheses we recently reported for a large number of substituted bidentate alkoxy-carbene ligands derived from epoxides, the reaction of imidazole with 2-adamantyl epoxide readily affords the bis(ethoxyadamantyl) substituted imidazolium salt [1-C{(NR)CHCH(NRH)}] (RH = CH2(2-adamantyl)OH), which has been isolated and structurally characterised as its iodide salt, [HC{(NRH)CH}2]I. Treatment with group 1 bases results in the loss of one ethoxy arm, to afford the structurally characterised monosubstituted alcohol imidazole, [HC{(NR)CHCHN}], or the lithium carbene complex [LiC{(NR)CHCHN}], a carbene complex containing a singly-N-functionalised alkoxy carbene. Alternatively, the monosubstituted alcohol imidazole may ben requaternised at the nitrogen atom with iso-propyl iodide to form [HC{(NRH)CHCH(NPri)}]I, from which a more standard lithium alkoxycarbene complex [LiC{(NR)CHCHNPri}] may be generated.  相似文献   

9.
Clark ME  Berti PJ 《Biochemistry》2007,46(7):1933-1940
Enolpyruvylshikimate-3-phosphate synthase (AroA, also called EPSP synthase) is a carboxyvinyl transferase involved in aromatic amino acid biosynthesis, forming EPSP from shikimate 3-phosphate and phosphoenolpyruvate. Upon extended incubation, EPSP ketal, a side product, forms by intramolecular nucleophilic addition of O4 to C2' of the enolpyruvyl group. The catalytic significance of this reaction was unclear, as it was initially proposed to arise from nonenzymatic breakdown of tetrahedral intermediate that had dissociated from AroA. This study shows that EPSP ketal formed in AroA's active site, not nonenzymatically, by demonstrating its formation in the presence of excess AroA. It formed both in the normal reaction and during AroA-catalyzed EPSP hydrolysis. In addition, nonenzymatic EPSP hydrolysis was studied to elucidate the catalytic imperative for enolpyruvyl reactions. Hydrolysis was acid-catalyzed, with a rate enhancement of >5 x 10(8)-fold. There was no detectable EPSP breakdown after 16 days at 90 degrees C in 1 M KOH, a solution that is 1000-fold more nucleophilic than neutral aqueous solutions. Thus, an unactivated enolpyruvyl group is not susceptible to nucleophilic attack. Enzymatic EPSP ketal formation therefore requires enolpyruvyl activation through protonation of C3' to form either a cationic intermediate or a highly cation-like transition state. Forming an EPSP cation requires the investment of considerable catalytic power by AroA. Such an intermediate is a potential target motif for inhibitor design.  相似文献   

10.
Four ring A steroidal epoxyenones as probable intermediate in the formation of catechol estrogens were synthesized. The isomeric 1 alpha,2 alpha-epoxy-17 beta-hydroxyestr-4-en-3-one (9) and 1 beta,2 beta-epoxy-17 beta-hydroxyestr-4-en-3-one (8) were synthesized from 17 beta-hydroxy-5 alpha-estra-3-one. The isomeric 4 alpha,5 alpha-epoxy-17 beta-hydroxyestr-1-en-3-one (11) and 4 beta,5 beta-epoxy-17 beta-hydroxyestr-1-en-3-one (10) were prepared from 19-nortestosterone. The reaction of 9 and 10 with sodium/ethanethiol resulted in the formation of three types of reactions leading to multiple products: 1,4-addition, opening of epoxide, and epoxide opening followed by dehydration. Reaction of 8 with ethanethiol gave only one compound identified as 2-ethanethio-1,4-estradien-17 beta-ol-3-one, while reaction of 9 with ethanethiol gave an unusual product identified as 4-estren-1 alpha,17 beta-diol-3-one. Unlike reaction of ethanethiol with 9 and 10, reaction with N-acetylecysteine or glutathione results in epoxide opening followed by dehydration leading to the formation of estradiol-4-thioethers.  相似文献   

11.
The glyoxalase system catalyzes the conversion of toxic methylglyoxal to nontoxic d-lactic acid using glutathione (GSH) as a coenzyme. Glyoxalase II (GlxII) is a binuclear Zn enzyme that catalyzes the second step of this conversion, namely the hydrolysis of S-d-lactoylglutathione, which is the product of the Glyoxalase I (GlxI) reaction. In this paper we use density functional theory method to investigate the reaction mechanism of GlxII. A model of the active site is constructed on the basis of the X-ray crystal structure of the native enzyme. Stationary points along the reaction pathway are optimized and the potential energy surface for the reaction is calculated. The calculations give strong support to the previously proposed mechanism. It is found that the bridging hydroxide is capable of performing nucleophilic attack at the substrate carbonyl to form a tetrahedral intermediate. This step is followed by a proton transfer from the bridging oxygen to Asp58 and finally C-S bond cleavage. The roles of the two zinc ions in the reaction mechanism are analyzed. Zn2 is found to stabilize the charge of tetrahedral intermediate thereby lowering the barrier for the nucleophilic attack, while Zn1 stabilizes the charge of the thiolate product, thereby facilitating the C-S bond cleavage. Finally, the energies involved in the product release and active-site regeneration are estimated and a new possible mechanism is suggested.  相似文献   

12.
3-Acyloxyl-2-oxopropyl ethers of umbelliferone were investigated as new fluorogenic substrates for lipases and esterases. The aliphatic primary alcohol-leaving group released the fluorescent product umbelliferone by an enolization/beta-elimination reaction similar to the triose phosphate isomerase (TIM) reaction. A similarly designed phenylacetamide provided a fluorescent probe for penicillin G acylase, whereby the enolization/beta-elimination sequence from the intermediate aminoketone was very fast and spontaneous even under acidic conditions. The corresponding epoxyketone was not fluorogenic with epoxide hydrolases (EH). These substrates represent periodate-free Clips-otrade mark substrates.  相似文献   

13.
Tetrachlorohydroquinone dehalogenase catalyzes the reductive dehalogenation of tetrachlorohydroquinone and trichlorohydroquinone during the biodegradation of the xenobiotic compound pentachlorophenol by Sphingobium chlorophenolicum. The mechanism of this transformation is of interest because it is unusual and difficult, and because aerobic microorganisms rarely catalyze reductive dehalogenation reactions. Tetrachlorohydroquinone dehalogenase is a member of the glutathione S-transferase superfamily. Many enzymes in this superfamily are capable of catalyzing nucleophilic aromatic substitution reactions. On the basis of this precedent, we have considered a mechanism for tetrachlorohydroquinone dehalogenase that involves a nucleophilic aromatic substitution reaction, either via an S(N)Ar mechanism or an S(RN)1-like mechanism, in the initial part of the reaction. Mechanistic studies were carried out with the wild type enzyme and with the C13S mutant enzyme, which catalyzes only the initial steps in the reaction. Three findings eliminate the possibility of a nucleophilic aromatic substitution reaction. First, the product of such a reaction, 2,3,5-trichloro-6-S-glutathionylhydroquinone, is not a kinetically competent intermediate. Second, the enzyme can carry out the reaction when the substrate is deprotonated at the active site. Nucleophilic aromatic substitution should not be possible when the substrate is negatively charged. Third, substantial normal solvent kinetic isotope effects on k(cat) and k(cat)/K(M,TriCHQ) are observed. Nonenzymatic and enzymatic nucleophilic S(N)Ar reactions typically show inverse solvent kinetic isotope effects.  相似文献   

14.
Rabeh WM  Alguindigue SS  Cook PF 《Biochemistry》2005,44(14):5541-5550
O-Acetylserine sulfhydrylase (OASS) catalyzes the last step in the cysteine biosynthetic pathway in enteric bacteria and plants, substitution of the beta-acetoxy group of O-acetyl-l-serine (OAS) with inorganic bisulfide. The first half of the sulfhydrylase reaction, formation of the alpha-aminoacrylate intermediate, limits the overall reaction rate, while in the second half-reaction, with bisulfide as the substrate, chemistry is thought to be diffusion-limited. In order to characterize the second half-reaction, the pH dependence of the pseudo-first-order rate constant for disappearance of the alpha-aminoacrylate intermediate was measured over the pH range 6.0-9.5 using the natural substrate bisulfide, and a number of nucleophilic analogues. The rate is pH-dependent for substrates with a pK(a) > 7, while the rate constant is pH-independent for substrates with a pK(a) < 7 suggesting that the pK(a)s of the substrate and an enzyme group are important in this half of the reaction. In D(2)O, at low pD values, the amino acid external Schiff base is trapped, while in H(2)O the reaction proceeds through release of the amino acid product, which is likely rate-limiting for all nucleophilic reactants. A number of new beta-substituted amino acids were produced and characterized by (1)H NMR spectroscopy.  相似文献   

15.
The unprotected methyl L-arabinofuranosides, D-ribofuranosides and D-xylofuranosides are transformed into the corresponding S-acetyl-5-thio derivatives by the thio-Mitsunobu reaction. Mesylation and subsequent reaction with sodium hydrogen carbonate led, depending on the configuration of the intermediate, to 2,5-anhydro-2-thio- or 3,5-anhydro-3-thiopentofuranosides. Due to inversion at C-3 or C-2 during the intramolecular nucleophilic displacement the products exhibit L-lyxo-, D-arabino- or D-lyxo-configuration. Analogously, the methyl 2,3-anhydro-D-ribofuranosides yielded 5-thio-S-acetates with intact 2,3-oxirane groups, which were cyclised with sodium hydrogen carbonate by epoxide ring opening and concomitant ring closure to form exclusively 3,5-anhydro-3-thio-D-xylofuranosides. A related 3,5-anhydro-3-seleno-D-lyxofuranoside was obtained by reaction of a 3,5-di-O-mesyl-D-arabinofuranoside with sodium hydrogen selenide. Several X-ray diffraction analyses proved the structures of the products.  相似文献   

16.
17.
Selective tosylation of 6,1′,6′-tri-O-tritylsucrose afforded the 2-O-tosyl derivative and not the 3-O-tosyl derivative as previously claimed. Treatment of the 2-tosylate with base afforded mainly (40%) the 2,3-manno-epoxide together with the 3,4-altro-epoxide which arose by migration of the epoxide ring. Ring-opening of the 2,3-epoxide with a variety of nucleophilic anions took place exclusively at C-3 to give altropyranosyl derivatives, whereas reaction of the epoxide with ammonium thiocyanate afforded the 2,3-allo-episulphide. Ring-opening of the 2,3-manno-epoxide with lithium iodide in ether gave 37% of the 3-deoxy-3-iodomannopyranosyl isomer, which arose by prior rearrangement of the 2,3-epoxide to the 3,4-epoxide.  相似文献   

18.
P J Lodi  J R Knowles 《Biochemistry》1991,30(28):6948-6956
To illuminate the role of histidine-95 in the catalytic reaction mediated by triosephosphate isomerase, 13C and 15N NMR titration studies have been carried out both on the wild-type enzyme and on a mutant isomerase in which the single remaining histidine (that at the active site) has been isotopically enriched in the imidazole ring. 15N NMR has proved especially useful in the unambiguous demonstration that the imidazole ring of histidine-95 is uncharged over the entire pH range of isomerase activity, between pH 5 and pH 9.9. The results require that the first pKa of histidine-95 is below 4.5. This abnormally low pKa rules out the traditional view that the positively charged imidazolium cation of histidine-95 donates a proton to the developing charge on the substrate's carbonyl oxygen. 15N NMR experiments on the enzyme in the presence of the reaction intermediate analogue phosphoglycolohydroxamate show the presence of a strong hydrogen bond between N epsilon 2 of histidine-95 and the bound inhibitor. These findings indicate that, in the catalyzed reaction, proton abstraction from C-1 of dihydroxyacetone phosphate first yields an enediolate intermediate that is strongly hydrogen bonded to the neutral imidazole side chain of histidine-95. The imidazole proton involved in this hydrogen bond then protonates the enediolate, with the transient formation of the enediol-imidazolate ion pair. Abstraction of the hydroxyl proton on O-1 now produces the other enediolate intermediate, which collapses to give the product glyceraldehyde 3-phosphate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Oximic 1,2,4-triazole ligands 2a-e were prepared from the reaction of 3-acetyl-4,5-dihydro-1H-1,2,4-triazoles 1a-e with hydroxylamine hydrochloride at room temperature. At higher temperatures, the reaction afforded, however, the novel ring transformation product 4-amino-2-(4-chlorophenyl)-5-methyl-2H-1,2,3,6-oxatriazine 3. The reaction of the ligands 2a-e with nickel (II) and palladium (II) acetates in ethanol at room temperature yielded the respective square planar complexes 5a-e, 6a,e. X-ray structure determination of one of these complexes (5a) revealed that metallation led to unexpected ring transformation of the triazole ligand. It is probable that such ring transformation generated the imidazole-N-oxide intermediate 4a which coordinated to Ni(II) ion, and the 4N-donor set comprises both imidazole nitrogen and arylhydrazone nitrogen. The whole process is associated with loss of one hydrogen molecule and formation of one new π-bond. The new compounds were characterized by elemental analysis, IR, 1H NMR, 13C NMR and HRMS.  相似文献   

20.
Dihydroflavins are facile reducing agents and potent nucleophiles. The dihydroflavin nucleophilic reactivity, as measured by the rate of covalent flavin adduct formation with tetrahydronaphthalene epoxides, is comparable to that of the thiolate anion (Y. T. Lee and J. F. Fisher (1993) J. Org. Chem. 58, 3712). In these reactions there appears subsequent to the nucleophilic cleavage of the epoxide by the dihydroflavin the product corresponding to formal hydride reduction product (at the benzylic carbon) of these epoxides. Thus the reaction of (+/-)-1a,2,3, 7b-tetrahydro-(1aalpha,2alpha,3beta,7balpha)-naphth[1,2-b]oxirene-2,3-diol (1), (+/-)-1a,2,3,7b-tetrahydro-(1aalpha,2beta,3alpha,7balpha)-naphth[1,2-b]oxirene-2,3-diol (2), and (+/-)-1a,2,3,7b-tetrahydro-(1aalpha,7balpha)-naphth[1,2-b]oxirene (3) in 9:1 (v/v) aqueous Tris buffer-dioxane, at both acidic and neutral pH, with FMNH(2) and 1,5-dihydrolumiflavin (LFH(2)) gave (following covalent flavin-epoxide adduct formation) the products having a methylene group at the benzylic position. The reduction product yield was proportional to the yield of the N(5) flavin-epoxide adduct intermediate, and the rate of the reaction was proportional to the dihydroflavin concentration. These observations are consistent with these reduction products resulting from bimolecular reaction between the dihydroflavin-epoxide adduct and a second molecule of dihydroflavin. Copyright 2000 Academic Press.  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