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 role of methionine and α-chymotrypsin-catalysed reactions   总被引:2,自引:2,他引:0  
1. The reaction of α-chymotrypsin with sodium periodate at pH5·0 has been investigated. The enzyme consumes 2 moles of periodate/mole, and there is a concomitant fall in enzymic activity (with respect to l-tyrosine ethyl ester) to 55% of that of the native enzyme. After 3hr. no further change is observed in periodate uptake or in catalytic activity. 2. The oxidized enzyme is a homogeneous preparation of partially active chymotrypsin. 3. In the oxidized enzyme, one of the two methionine residues in the molecule has been converted into its sulphoxide. It is this reaction only that is responsible for the loss of activity. 4. The rate constants for the enzyme-catalysed acylation and deacylation reactions are unaltered by oxidation of the enzyme, both for a non-specific substrate (p-nitrophenyl acetate), and for three specific substrates: N-acetyl-l-tryptophan ethyl ester, N-acetyl-l-tryptophanamide and N-acetyl-l-valine ethyl ester. 5. The Km values for the aromatic substrates with the oxidized enzyme are twice those with the native enzyme. No change in Michaelis constant is seen for the non-aromatic substrate N-acetyl-l-valine ethyl ester. 6. The evidence points to the oxidized methionine residue in the modified enzyme being situated in the locus of the active site at which aromatic (or bulky) side chains of the substrates are bound.  相似文献   

2.
《Carbohydrate research》1987,162(2):199-207
The 2,1′-O-isopropylidene derivative (1) of 3-O-acetyl-4,6-O-isopropylidene-α-d-glucopyranosyl 6-O-acetyl-3,4-anhydro-β-d-lyxo-hexulofuranoside and 2,3,4-tri-O-acetyl-6-O-trityl-α-d-glucopyranosyl 3,4-anhydro-1,6-di-O-trityl-β-d-lyxo-hexulofuranoside have been synthesised and 1 has been converted into 2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl 1,6-di-O-acetyl-3,4-anhydro-β-d-lyxo-hexulofuranoside (2). The SN2 reactions of 2 with azide and chloride nucleophiles gave the corresponding 2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl 1,3,6-tri-O-acetyl-4-azido-4-deoxy-β-d-fructofuranoside (6) and 2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl 1,3,6-tri-O-acetyl-4-chloro-4-deoxy-β-d-fructofuranoside (8), respectively. The azide 6 was catalytically hydrogenated and the resulting amine was isolated as 2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl 4-acetamido-1,3,6-tri-O-acetyl-4-deoxy-β-d-fructofuranoside. Treatment of 5 with hydrogen bromide in glacial acetic acid followed by conventional acetylation gave 2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl 1,3,6-tri-O-acetyl-4-bromo-4-deoxy-β-d-fructofuranoside. Similar SN2 reactions with 2,3,4,6-tetra-O-acetyl-α-d-glucopyranosyl 1,6-di-O-acetyl-3,4-anhydro-β-d-ribo-hexulofuranoside (12) resulted in a number of 4′-derivatives of α-d-glucopyranosyl β-d-sorbofuranoside. The regiospecific nucleophilic substitution at position 4′ in 2 and 12 has been explained on the basis of steric and polar factors.  相似文献   

3.
Summary Nucleoside-5-phosphorimidazolides react readily with acylating agents to give N-substituted products that are highly activated. In most cases these acylated derivatives undergo rapid hydrolysis to give nucleoside 5-phosphates, whether or not a complementary template is present. However, guanosine 5-phosphorimidazolide reacts with diethyl pyrocarbonate to give a derivative that oligomerizes rapidly and efficiently in the presence of polycytidylic acid and Pb2+. The reaction is complete in about 1 h, whereas the corresponding reaction in the absence of an acylating agent takes several days. However, the final yield of long oligomers is lower when diethyl pyrocarbonate is present.  相似文献   

4.
Summary The -galactosidase from Streptococcus thermophilus formed transferase products (including up to six disaccharides and two trisaccharides) during the hydrolysis of lactose to glucose and galactose. The extent of transferase products formed was dependent on the initial lactose concentration, reaching up to 40% of the total carbohydrate at 70% w/v lactose. At high lactose concentrations (40% w/v) trisaccharide transferase products were formed initially, followed by the appearance of disaccharide transferase products. In contrast, at low lactose concentrations (7.5 w/v), only traces of the trisaccharides were detected with disaccharides being the predominant transferase products. The disaccharide products accumulated to relatively high concentrations late in the overall hydrolysis of lactose, at both high and low initial lactose concentrations, while the trisaccharides peaked much earlier and were themselves subsequently hydrolysed prior to the complete disappearance of lactose. It was possible to study the hydrolysis of galactosyl lactose by the S. thermophilus -galactosidase using a semi-pure galactosyl lactose preparation containing 5% lactose. The hydrolysis of this trisaccharide occurred via at least four disaccharide intermidiates, which appeared chromatographically identical to the disaccharide transferase products formed during lactose hydrolysis. This suggests that the enzymic formation and subsequent hydrolysis of galactosyl lactose occurs via coincident reaction pathways. The initial rate of galactose over glucose formation during galactosyl lactose hydrolysis changed from a ratio of 3:1 at low (2–3% w/v) substrate concentrations to 1.5:1 at high (>20% w/v) concentrations. This indicates a shift in the preferred initial cleavage site from the galactose-galactose bond to the galactose-glucose bond.  相似文献   

5.
6.
Summary The diastereofacial selective imine-ene reactions with-imino esters, prepared from (–)-8-phenylmenthyl glyoxylate, are shown to provide an efficient entry to the asymmetric synthesis of-amino acids. The feasibility study of the asymmetric catalysis is also reported on the enantiofacial selective ene reactions with prochiral-imino esters.  相似文献   

7.
The synthesis of the fully benzylated α- and β-d-glucopyranosyluronic esters of 1-benzyl N-benzyloxycarbonyl-l-aspartic and -glutamic acids and N-(tert-butoxycarbonyl)-l-phenylalanine, followed by hydrogenolysis, afforded the respective anomers of the 1-O-acyl-d-glucopyranuronic acids 2, 7, and 12. Esterification of both anomers of the N-acetylated derivatives of 2 and 7 by diazomethane was accompanied by glycosyl-bond cleavage, and, in the case of the α anomers, with concomitant 1→2 acyl migration to give, after O-acetylation, the 2-O-acyl O-acetyl methyl ester derivatives 5 and 10, respectively. Similarly, 12α yielded methyl 1,3,4-tri-O-acetyl-2-O-[N-(tert-butoxycarbonyl)-l-phenylalanyl]-d-glucopyranuronate and an analogue having a furanurono-6,3-lactone structure. Esterification of the C-5 carboxyl group, in 1-O-acyl-α-d-glucopyranuronic acids by methanol in the presence of the BF3?-MeOH reagent (1–1.5 equiv.) proceeded without acyl migration. By using this procedure, followed by acetylation, the N-acetylated derivative of afforded methyl 2,3,4-tri-O-acetyl-1-O-(1-methyl N-acetyl-l-glutam-5-oyl)-α-d-glucopyranuronate, and 12α gave methyl 2,3,4-tri-O-acetyl-1-O-(N-acetyl-l-phenylalanyl)-α-d-glucopyranuronate; the formation of the latter involved cleavage of the tert-butoxycarbonyl group by BF3, followed by N-acetylation in the next step.  相似文献   

8.
The stereochemistry of the replacement of the SH-group of cysteine by CN catalyzed by β-cyanoalanine synthetase was studied using cysteine stereospecifically tritiated at C-3. Analysis of the resulting β-cyanoalanine by conversion into fumarate via aspartate and malate showed that the reaction had occurred with retention of configuration at C-3. Using cystine stereospecifically labeled at C-3 with tritium or with tritium and deuterium, it was found that the α,β-elimination reaction catalyzed by S-alkylcysteine lyase involves stereo-specific replacement of the β-substituent of the substrate by a hydrogen derived from the solvent, D2O or H2O, with retention of configuration to give pyruvate containing a chiral methyl group. The results are discussed, particularly in the light of mechanistic proposals by Braunstem and co-workers.  相似文献   

9.
β-Glucosidases (EC 3.2.1.21), abundant enzymes distributed in animals, plants and microorganism, has been generating lots of attentions for bioethanol production from cellulosic biomass. In this study, using three different origins of β-glucosidases, glucose productivity of β-glucosidase-catalyzed hydrolysis reactions in the presence of synthetic betaine-type metabolite analog (2-N,N,N-tri-n-butylammonium) acetate, was investigated. By the addition of the analog, the hydrolysis yields for all β-glucosidases was highly improved from 4–13 to 64–100 %. To understand the factors affecting on the yield enhancements, the kinetic parameters, inhibition constants of end-product and temporal stability of β-glucosidases were compared. As a result, enhancement of the yields is mainly related to the increase in the temporal stability of β-glucosidases in the presence of the analog. The present findings lead to not only improve the glucose productivity of β-glucosidase-catalyzed hydrolysis reaction toward bioethanol production but also apply to a new stabilization method for various unstable enzymes.  相似文献   

10.
Treatment of 6,6′-dichloro-6,6′-dideoxysucrose with a combination of 2,2-dimethoxypropane, N,N-dimethylformamide, and toluene-p-sulphonic acid (reagent A), followed by acetylation, gave the 1′,2:3,4-diacetal 1 (39%) and the 1′,2-acetal 2 (37%). A similar reaction of methyl 6-chloro-6-deoxy-α-D-glucopyranoside with reagent A yielded the corresponding 2,3- and 3,4-acetal derivatives in yields of 29% and 9%, respectively. The structures of 1 and 2 have been confirmed by 1H-n.m.r. spectroscopy and by chemical transformations.  相似文献   

11.
A novel class of inhibitors of the enzyme γ-glutamyl transpeptidase (GGT) were evaluated. The analog OU749 was shown previously to be an uncompetitive inhibitor of the GGT transpeptidation reaction. The data in this study show that it is an equally potent uncompetitive inhibitor of the hydrolysis reaction, the primary reaction catalyzed by GGT in vivo. A series of structural analogs of OU749 were evaluated. For many of the analogs, the potency of the inhibition differed between the hydrolysis and transpeptidation reactions, providing insight into the malleability of the active site of the enzyme. Analogs with electron withdrawing groups on the benzosulfonamide ring, accelerated the hydrolysis reaction, but inhibited the transpeptidation reaction by competing with a dipeptide acceptor. Several of the OU749 analogs inhibited the transpeptidation reaction by slow onset kinetics, similar to acivicin. Further development of inhibitors of the GGT hydrolysis reaction is necessary to provide new therapeutic compounds.  相似文献   

12.
Summary Various lipases have been evaluated as biocatalysts for the enrichment of -linolenic acid from a commercial fungal oil derived from Mucor sp. by selective esterification of the fungal oil fatty acids with n-butanol or by selective hydrolysis of the oil. Lipase from M. miehei (Lipozyme), as compared to lipases from Candida cylindracea, Penicillium cyclopium, and Rhizopus arrhizus, was found to be most effective in the enrichment of -linolenic acid in unesterified fatty acids upon esterification of the fungal oil fatty acids with n-butanol. Thus, the -linolenic acid content could be raised from 10.4% in the starting material to 68.8% in the unesterified fatty acids. Selective hydrolysis of the fungal oil triacyglycerols using Lipozyme resulted in about 1.5-fold enrichment of -linolenic acid in the unhydrolysed acylglycerols. Other lipases tested, such as those from P. cyclopium, C. cylindracea, R. arrhizus, Penicillium sp. (Lipase G), porcine pancreas and Chromobacterium viscosum, were also rather ineffective in the enrichment of -linolenic acid by selective hydrolysis of the fungal oil triacylglycerols. Offprint requests to: K. D. Mukherjee  相似文献   

13.
14.
15.
Reaction of mono-, di-, and trisaccharide derivatives of methyl beta-D- and octyl beta-D-mannopyranosides bearing ester groups at isolated and non-isolated positions on the same molecule, under Zemplén conditions (catalytic amount of sodium methoxide in methanol) gave partially deacylated compounds, in which the O-acyl groups were retained at isolated sites. In the case of one disaccharide, all the benzoyl groups remained intact at the reducing end, while all the acetyl functions were removable from the nonreducing end. In another case, both isolated ester groups at positions 2 and 4 were retained at the reducing end. The isolated 2-O-acyl groups on methyl alpha-D-mannopyranoside compounds were more labile than on the corresponding beta-mannosides under the same conditions. The mechanism of the reaction may be different for ester groups at isolated or non-isolated positions. In the latter case, acyl migration may take place and carry acyl groups into a less hindered position.  相似文献   

16.
In the presence of the modifier protein -lactalbumin, bovine milk galactosyltransferase transfers galactose to glucose forming lactose instead of transferring toN-acetylglucosamine formingN-acetyllactosamine. At low concentrations of -lact-albumin, the lactosamine synthase activity is stimulated by -lactalbumin and decreases when the lactose synthase activity develops along a sigmoidal curve. The observation suggests that different interactions between -lactalbumin and enzyme were responsible for the modulating effect of the -lactalbumin in the lactose and lactosamine synthase reactions.To study the nature of the protein-protein interactions, -lactalbumin was both modified and cleaved chemically. Reduction and alkylation with iodoacetic acid, iodoacetamide or 4-vinylpyridine abolished the ability of the -lactalbumin to induce lactose synthase activity but stimulated lactosamine synthase activity 7-to 12-fold.A peptide fragment corresponding to residues 26–60 of -lactalbumin isolated from a 2-(2-nitrophenylsulphenyl)-3-methyl-3-bromo-indolene (BNPS-skatole) fragmentation of the molecule was active in the lactosamine but not lactose synthase reaction. We concluded that, whereas lactose synthase required -lactalbumin, in the native conformation, lactosamine synthase activity was stimulated by a linear sequence of amino acids in peptide 26–60.Abbreviations MES 4-N-morpholinoethanesulfonic acid - TRIS 2-amino-2-(hydroxymethyl)-1,3-propanediol - UDP-Gal uridinediphosphogalactose - BNPS-skatole 2-(2-nitrophenylsulphenyl)-3-methyl-3-bromo-indolene - EDTA ethylene diamine tetra acetic acid  相似文献   

17.
The reaction of sucrose with a combination of 2,2-dimethoxypropane, N,N-dimethylformamide, and toluene-p-sulphonic acid (reagent A) gave, after acetylation followed by chromatography, 1′,2:4,6-di-O-isopropylidenesucrose tetra-acetate (1) in 15% yield. The structure of 1 was determined on the basis of p.m.r. and mass spectrometry, and by chemical transformations. Treatment of 1 with aqueous acetic acid afforded sucrose 3,3′,4′,6′-tetra-acetate 2. Reacetalation of 2 using reagent A gave 1 in 80% yield. The p.m.r. spectrum of 2 confirmed the presence of hydroxyl groups at C-2 and C-4. The following sequence of reactions showed that the remaining two hydroxyl groups were located at C-6 and C-1′. Selective tritylation of 2 gave 1′,6-di-O-tritylsucrose 3,3′,4′,6′-tetra-acetate (3) as the minor, and 6-O-tritylsucrose 3,3′,4′,6′-tetra-acetate (4) as the major, product. When tritylation was carried out under forcing conditions, 2 gave 3 as the major product. Acetylation of 4 afforded 6-O-tritylsucrose hepta-acetate. Mesylation of 2 gave the tetramethanesulphonate 5, which afforded the 6-dcoxy-6-iodo derivative 6 on treatment with a refluxing solution of sodium iodide in butanone. Treatment of 3 with methanesulphonyl chloride in pyridine gave the disulphonate 7, which on detritylation followed by acetylation gave 2,4-di-O-methanesulphonylsucrose hexa-acetate (9). Treatment of 9 with sodium benzoate in hexamethylphosphoric triamide displaced the 4-sulphonate, with inversion of configuration, to give the galacto derivative 10.  相似文献   

18.
The preponderant pathway of octene-1 degradation by octane- and octene-1-grownPseudomonas aeruginosa cells (strain 473) starts with oxidation of the methyl group.In addition, with both types of cells minor reactions occur that involve the double bond. The formation of a 1,2-epoxide was reported earlier. In addition, the identification of the saturated C8 fatty acid is a strong indication that the terminal methylene group is partially converted into an aldehydic group. The aldehyde seems to be formed beside the epoxide and the latter is not an intermediate. Enzymatic dihydroxylation of the double bond, if at all occurring, is masked by non-enzymatic hydrolysis of the epoxide.The formation of a saturated methyl ketone could not be detected. Nor is the olefinic group converted into a primary or secondary alcohol group under conditions which result in accumulation of octanol-1 from octane and of 7-octenol-1 from octene-1.The absence of a saturated alcohol among the intermediates in octene-1 degradation excludes hydration of the double bond as well as other mechanisms leading to saturated alcohols.Accumulation of C8 fatty acids was effected by inhibition of -oxidation with acrylate, whereas addition of a competing alcohol substrate (octanediol-1,8) yielded detectable amounts of the alcoholic intermediates.  相似文献   

19.
Enantiomerically pure α-hydroxy amides have been prepared from the corresponding α-oxo esters by the use of a double sequence reaction involving in a first step the highly enantioselective Saccharomyces cerevisiae bioreduction and then in a second step, the resulting α-hydroxy esters followed a non-enantiospecific lipase catalyzed aminolysis with n-butylamine reaction. In the first non-organic solvent process, the moistened baker’s yeast reduced seven α-oxo esters with high conversions degree (93% for one substrate and >99% for the others) and high enantioselectivities [>99% for all the substrates except for ketopantoyl lactone, which gave 88% of enantiomeric excess (ee)]. At the same way, the isolated resulting chiral α-hydroxy esters were subjected to the second Candida antarctica lipase fraction B (CAL-B) catalyzed aminolysis in dioxane conducting to the corresponding chiral α-hydroxy amides with high conversions degree, between 88 and 99%. Both processes were carried out at 28–30°C.  相似文献   

20.
Prey species gain protection by imitating signals of unpalatable models in defensive mimicry. Mimics have been traditionally classified as Batesian (palatable mimic resembling an unpalatable model) or Müllerian (unpalatable mimic resembling a similarly unpalatable model). However, recent studies suggest that rather than discrete categories, the phenomenon of mimicry can be better understood as a continuum. The level of unpalatability of defended prey is a key factor in determining the type of mimetic relationship. Herein, we used insects (ladybugs and true bugs) from a putative European “red–black” mimetic complex as experimental models of defended species and crickets as a control prey. We offered the prey to two species of sympatric invertebrate predators (praying mantis and spider) and video recorded the interactions. We tested three alternative hypotheses, namely (i) the three red–black species tested are similarly defended against both predators; (ii) some red–black species are better defended than others against both predator species, and (iii) the effectiveness of the red–black species defenses is predator dependent. Both predators attacked all prey types with a similar frequency. But while all three red–black species similarly elicited aversive behaviors in spiders, the mantises' aversive reactions varied depending on the prey species. Our results provide support to the third hypothesis, suggesting that the same prey species can fall into different parts of the spectrum of palatability–unpalatability depending on the type of predator.  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