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 15 毫秒
1.
Summary Water-sobuble trypsin specific macroligands were prepared to separate the trypsin -chymotrypsin mixture with affinity-ultrafiltration technique. Soya bean trypsin inhibitor (STI) attached to cyanogen bromideactivated Dextran showed a good selectivity and low non-specific adsorption properties. The experiments performed with STI-Dextran polymer gave a 81% purified trypsin from 50%-50% mixture.  相似文献   

2.
Does trypsin cut before proline?   总被引:1,自引:0,他引:1  
Trypsin is the most commonly used enzyme in mass spectrometry for protein digestion with high substrate specificity. Many peptide identification algorithms incorporate these specificity rules as filtering criteria. A generally accepted "Keil rule" is that trypsin cleaves next to arginine or lysine, but not before proline. Since this rule was derived two decades ago based on a small number of experimentally confirmed cleavages, we decided to re-examine it using 14.5 million tandem spectra (2 orders of magnitude increase in the number of observed tryptic cleavages). Our analysis revealed a surprisingly large number of cleavages before proline. We examine several hypotheses to explain these cleavages and argue that trypsin specificity rules used in peptide identification algorithms should be modified to "legitimatize" cleavages before proline. Our approach can be applied to analyze any protease, and we further argue that specificity rules for other enzymes should also be re-evaluated based on statistical evidence derived from large MS/MS data sets.  相似文献   

3.
Vukoti KM  Kadiyala CS  Miyagi M 《FEBS letters》2011,585(24):3898-3902
Streptomyces erythraeus trypsin (SET) is a serine protease that is secreted extracellularly by S. erythraeus. We investigated the inhibitory effect of α(1)-antitrypsin on the catalytic activity of SET. Intriguingly, we found that SET is not inhibited by α(1)-antitrypsin. Our investigations into the molecular mechanism underlying this observation revealed that SET hydrolyzes the Met-Ser bond in the reaction center loop of α(1)-antitrypsin. However, SET somehow avoids entrapment by α(1)-antitrypsin. We also confirmed that α(1)-antitrypsin loses its inhibitory activity after incubation with SET. Thus, our study demonstrates that SET is not only resistant to α(1)-antitrypsin but also inactivates α(1)-antitrypsin.  相似文献   

4.
5.
Wheat albumins were extracted from whole wheat flour with 150 mM sodium chloride solution and precipitated between 0·4 and 1·8 M ammonium sulphate. The albumin precipitate was separated by gel filtration on Sephadex G100 into five peaks. Three peaks (II, III, and IV), whose MWs were 60 000, 24 000 and 12 500 daltons respectively, were active toward several insect α-amylases, whereas only peak III inhibited human saliva and pancreatic α-amylases. Peaks III and IV also inhibited trypsin. In each active peak, we found several α-amylase inhibitors slightly different in their electrophoretic mobilities in a Tris—glycine buffer system (pH 8·5), whereas only one major trypsin inhibitor was present in peaks III and IV. In contrast to α-amylase inhibitors that were all anodic, trypsin inhibitors migrated to the cathode under our experimental conditions. From a quantitative standpoint, wheat albumins that inhibit trypsin are negligible, whereas about 2/3 of the total albumin inhibits amylases from different origins. All inhibitor components of peak III were active toward both insect and mammalian α-amylases. Moreover, they reversibly dissociated in the presence of 6 M guanidine hydrochloride giving two similar subunits.  相似文献   

6.
A trypsin inhibitor, isolated from whole-wheat grain (Triticum aestivum L.) by the method of biospecific chromatography on trypsin-Sepharose, was potent in inhibiting human salivary α-amylase. The bifunctional α-amylase/trypsin inhibitor was characterized by a narrow specificity for other α-amylases and proteinases. The high thermostability of the inhibitor was lost in the presence of SH group-reducing agents. The inhibitor-trypsin complex retained its activity against α-amylase. The inhibitor—α-amylase complex was active against trypsin. Studies of the enzyme kinetics demonstrated that the inhibition of α-amylase and trypsin was noncompetitive. Our results suggest the existence of two independent active sites responsible for the interaction with the enzymes.  相似文献   

7.
The monoaldehyde derivative of -cyclodextrin was attached to trypsin via reductive alkylation with NaBH4. The thermostability was enhanced from 49.5 °C to 60 °C for modified trypsin. The activation free energy of thermal inactivation at 50 °C was increased by 3.2 kJ mol–1. The conjugated enzyme retained 100% of its initial activity after 3 h incubation at pH 9.  相似文献   

8.
9.
Bovine pancreatic trypsin was chemically modified by several β-cyclodextrin (β-CD) derivatives using 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide as coupling agent. The modifying agents used were mono-6-amino-6-deoxy-β-cyclodextrin (CDNH2), mono-6-ethylenediamino-6-deoxy-β-cyclodextrin (CDEN), mono-6-propylenediamino-6-deoxy-β-cyclodextrin (CDPN) and mono-6-butylenediamino-6-deoxy-β-cyclodextrin (CDBN). The enzyme–cyclodextrin conjugates contained about 2 mol of oligosaccharide per mol of trypsin. The catalytic and thermal stability properties of trypsin were improved by the attachment of cyclodextrin residues, and these effects were markedly noticeable for cyclodextrin (CD) derivatives having an even number of carbon atoms in the spacer arms. The thermostability of the enzyme was increased by about 2.4–14.5 °C after modification. The conjugates prepared were also more stable against thermal incubation at different temperatures ranging from 45 to 60 °C. In comparison with native trypsin, the enzyme–cyclodextrin complexes were markedly more resistant to autolytic degradation at pH 9.0. Attending to the results here reported, we suggest that conjugation of enzymes with β-CD derivatives might be an useful method for improving the stability and the catalytic properties of these biocatalysts.  相似文献   

10.
Summary Crystalline trypsin was irradiated in oxygen-free suspension media of methanol, ethanol and n-heptane with60Co--rays at 77 K or 273 K. Measurements with ESR and activity determinations revealed no influence of ethanol and n-heptane on the formation of free radicals and inactivation of trypsin. Especially, the results are independent on the polarity of the suspension media and correspond to an irradiation of trypsin in vacuum. On the other hand, methanol leads to a decay of radiation induced radicals and to an increased inactivation. The results are discussed in comparison to analogous experiments carried out with ultra-violet light.  相似文献   

11.
Clear G-bands were revealed by applying the TUG method on chromosomes of 5 species of higher plants(Lilium davidii,Viciafaba,Hordeum vulgare,Ginkgo biloba and Triticum monococcum).Some details of the TUG method which consisted of treating chromosomes with both trypsin and urea were also studied.The mechanisms that might account for the presence of G-bands in plants were discussed.  相似文献   

12.
13.
考察了H^ 和Na^ 变化对紫外吸收单波长法、双波长法和考马斯亮蓝法的影响。结果表明双波长法比考马斯亮蓝法和单波长法稳定性要好。选择双波长法用于磺酸型离子交换树脂吸附胰蛋白酶(trypsin)和牛血清白蛋白(BSA)的检测,结果表明:交换容量为4-88mmol/g的树脂对BSA和trypsin的吸附速率和饱和吸附量均高于交换容量为3.17mmol/g的树脂。  相似文献   

14.
Summary To distinguish ligand-induced structural states of the (Na+–K+)-ATPase, the purified membrane-bound enzyme isolated from rat kidneys was digested with trypsin in the presence of various combinations of Na+, K+, Mg++ and ATP. It was found that first the large and then the small polypeptide chain of the (Na+–K+)-ATPase was degraded, indicating that the lysine and arginine residues of the large chain are more exposed than are those of the small one. The (Na+–K+)-ATPase activity was inactivated in parallel with the degradation of the large polypeptide chain. After the degradation of the large polypeptide chain, about 75% of the (Na+–K+)-ATPase protein remained bound to the membrane, demonstrating that the split protein segments were only partially released.It was found that the combinations of ATP, Mg++, Na+ and K+ present during trypsin digestion influenced the time course and degree of degradation of the (Na+–K+)-ATPase protein. The degradations of the large and the small polypeptide chain were affected in parallel. Thus, certain ATP and ligand combinations influenced neither the degradation of the large nor the degradation of the small polypeptide chain, whereas by other combinations of ATP and ligands the degree of susceptibility of both polypeptide chains to trypsin was equally increased or reduced.In the absence of ATP the time course of trypsin digestion of the (Na+–K+)-ATPase was the same, whether Na+ or K+ was present. With low ATP concentrations (e.g., 0.1mm), however, binding of Na+ or K+ led to different degradation patterns of the enzyme. If a high concentration of ATP (e.g., 10mm) was present, Na+ and K+ also influenced the degradation pattern of the (Na+–K+)-ATPase, but differentially compared to that at low ATP concentrations, since the effects of Na+ and K+ were reversed. Furthermore, it was found that the degradation of the small chain was only influenced by certain combinations of ATP, Mg++, Na+ and K+ if the large chain was intact when the ligands were added to the enzyme.The described results demonstrate structural alterations of the (Na+–K+)-ATPase complex which are supposed to include a synchronous protrusion or retraction of both (Na+–K+)-ATPase subunits. The data further suggest that ATP and other ligands primarily alter the structure of the large (Na+–K+)-ATPase subunit. This structural alteration is presumed to lead to a synchronous movement of the small subunit of the enzyme. The structural state of the (Na+–K+)-ATPase is regulated by binding of Na+ or K+ to the enzyme-ATP complex. The effects of Na+ and K+ on the (Na+–K+)-ATPase structure are modulated by the ATP binding to high affinity and to low affinity ATP binding sites.  相似文献   

15.
When human α2 macroglobulin (α2M) or its asialo-[3H]galactose derivative reacts with trypsin, a glycopeptide of molecular weight 3500–4000 is released from the α2M. The glycopeptide was purified on Biogel P-4 columns and its amino acid and carbohydrate composition were determined. The oligosaccharide contains sialic acid, galactose, mannose and GlcNAc in a ratio of 1.0:0.73:3.85:2.85 and is apparently attached to protein in a GlcNAc→asparagine linkage.  相似文献   

16.
The recently identified fungal protease inhibitors cnispin, from Clitocybe nebularis, and cospin, from Coprinopsis cinerea, are both β-trefoil proteins highly specific for trypsin. The reactive site residue of cospin, Arg27, is located on the β2–β3 loop. We show here, that the reactive site residue in cnispin is Lys127, located on the β11–β12 loop. Cnispin is a substrate-like inhibitor and the β11–β12 loop is yet another β-trefoil fold loop recruited for serine protease inhibition. By site-directed mutagenesis of the P1 residues in the β2–β3 and β11–β12 loops in cospin and cnispin, protease inhibitors with different specificities for trypsin and chymotrypsin inhibition have been engineered. Double headed inhibitors of trypsin or trypsin and chymotrypsin were prepared by introducing a second specific site residue into the β2–β3 loop in cnispin and into the β11–β12 loop in cospin. These results show that β-trefoil protease inhibitors from mushrooms exhibit broad plasticity of loop utilization in protease inhibition.  相似文献   

17.
Several trypsin inhibitors with different mobilities on polyacrylamide gel electrophoresis occur in the tubers of taro (Colocasia antiquorum), and they each have a dimeric molecular weight of 40,000. Of all the constituent subunits, molecular weight 20,000, of the taro trypsin inhibitor (TTI), three major subunit components were separated by chromatography on SP-Sephadex C-25 in 8 M urea, and they were named protomers alpha, beta, and gamma in the order of their elution from the SP-Sephadex column. After removal or dilution of the urea, the three protomers could be either reassociated individually or hybridized with each other to form dimeric inhibitors. All of the reassociated dimers were powerful inhibitors of trypsin. Among them, each dimer derived from protomers alpha and gamma was a weak inhibitor of chymotrypsin, whereas the dimer of protomer beta did not inhibit the enzyme. Therefore TTI is presumed to be a mixture of heterogeneous and homogenous dimers whose properties reflect those of their constituent protomers. It was also proved that the major three trypsin inhibitors (TTI-I, TTI-II, and TTI-III) previously isolated from taro tubers are composed of protomers alpha and gamma, i.e., TTI-II is a heterogeneous dimer of protomers alpha and gamma, and TTI-I and TTI-III are homogeneous dimers of protomers alpha and gamma, respectively. The molecular weight of a trypsin-TTI complex saturated with trypsin was found to be 79,000, suggesting the formation of a tetrameric complex.  相似文献   

18.
Extracts ofAscophyllum nodosum, Fucus serratus, F. vesiculosus andPelvetia canaliculata contain inhibitors of α-amylase, lipase and trypsin. The inhibitors were isolated and identified by1H NMR spectroscopy as polyphenols which have apparent molecular weights in the range from 30 000 to 100 000 daltons, as determined by ultra-filtration with Amicon membranes. These polyphenols account for the whole of the inhibitory activity in crude algal extracts. The compounds inhibit α-amylase and trypsin in an apparently non-competitive manner, when preincubated with the enzymes, and the inhibition is directly proportional to the concentration of the inhibitor. Starch protects α-amylase when added to the enzyme together with the inhibitors. Under this condition the effectiveness of the inhibitors is reduced ten-fold.  相似文献   

19.
Affinity adsorbents for bovine trypsin were prepared by covalently coupling p-(p′-amino-phenoxypropoxy)benzamidine to cellulose and to agarose. Trypsin binds to both adsorbents at pH6–8 and is released at low pH values or in the presence of n-butylamine hydrochloride. Pure β-trypsin may be eluted from crude trypsin bound at pH8.0 to the cellulose adsorbent by stepwise elution with an acetate buffer, pH5.0. Both α- and β-trypsin may be isolated by chromatography of crude trypsin on the agarose derivative in an acetate buffer, pH4.0. These two methods for purifying the trypsin are specific to the particular adsorbents. They are rapid and convenient in use. Both methods leave a mixture of the two enzymes bound to the adsorbent and release occurs only at low pH values. The effects of pH, composition and ionic strength of buffer and other variables on both purification methods are described. Affinity adsorbents of soya-bean trypsin inhibitor and of N-α-(N′-methyl-N′-sulphanilyl) sulphanilylagmatine bound to agarose were prepared, but were found to be of limited usefulness in the purification of trypsin.  相似文献   

20.
《Bioorganic chemistry》1987,15(1):50-58
Trypsin-specific substrate analogs, “inverse substrates,” carrying the fluorescent dimethylaminonaphthalene group were synthesized. Preparation of acyl trypsins in which the fluorescent group is attached to the catalytic residue through a spacer group of various chain lengths was successfully carried out by use of these inverse substrates. The topographical structure of the trypsin active site vicinity was estimated on the basis of the fluorescence spectra of these acyl trypsins.  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