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.
Abstract

Hydrolysis studies of photo-switchable N-maleyl-amino acid-(4-phenylazophenyl) esters of Ala, Gly, Met, Phe, and Pro were performed using the proteases trypsin and chymotrypsin. It has been found out that the cis-isomers were hydrolyzed faster than the trans-isomers. In dependence of the amino acid in the P1 position the velocity graduation is Phe?>?Met?>?Ala?>?Gly?>?Pro for trypsin and Phe???Met?>?Ala?>?Gly?>?Pro for chymotrypsin for both isomers.  相似文献   

2.
Conformations of two pairs of dehydropeptides with the opposite configuration of the ΔPhe residue, Boc‐Gly‐ΔZPhe‐Gly‐Phe‐OMe ( Z‐ OMe ), Boc‐Gly‐ΔEPhe‐Gly‐Phe‐OMe ( E‐ OMe ), Boc‐Gly‐ΔZPhe‐Gly‐Phe‐p‐NA ( Z‐p‐ NA ), and Boc‐Gly‐ΔEPhe‐Gly‐Phe‐p‐NA ( E‐p‐ NA ) were compared on the basis of CD and NMR studies in MeOH, trifluoroethanol (TFE), MeCN, chloroform, and dimethylsulfoxide (DMSO). The CD results were used as the additional input data for the NMR‐based determination of the detailed solution conformations of the peptides. It was found that E‐ OMe is unordered and Z‐ OMe , Z‐p‐ NA , and E‐p‐ NA adopt the β‐turn conformation. There are two overlapping β‐turns in each of those peptides: type II and type III′ in Z‐ OMe and Z‐p‐ NA , and two type III in E‐p‐ NA . The ordered structure‐inducing properties of ΔZPhe and ΔEPhe in the peptides studied depend on the C‐terminal blocking group. In methyl esters, the ΔZPhe residue is a strong inducer of ordered conformations whereas the ΔEPhe one has no such properties. In p‐nitroanilides, both isomers of ΔPhe cause the peptides to adopt ordered structures to a similar extent. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 1055–1064, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

3.
The activities of novel Cbz-N-protected α-aminophosphonic phenyl esters, analogs of leucine (1–15) and phenylalanine (17–29), which are substituted at the phenyl ester rings, as well as of their peptidic derivatives (31–43), were investigated for their inhibitory effects on chymotrypsin and subtilisin. The chemical nature and position of the examined substituents clearly demonstrated a strong structure–activity relationship. Among all synthesized compounds the most potent phosphonic-type inhibitors of subtilisin and chymotrypsin were identified, with k2/Ki values 114,380?M?1s?1 and 307,380?M?1s?1, respectively.  相似文献   

4.
Tosylphenylalanine chloromethyl ketone (TPCK) and tosyllysine chloromethyl ketone (TLCK) are irreversible modifiers of histidine which is located in the catalytic triad of chymotrypsin and trypsin, respectively. The effects of TPCK and TLCK on the histidine in the catalytic triad of the desensitized butyrylcholinesterase (BChE), prepared from human serum by heating at 45°C for 24 h, were investigated in detail. It is found that these reagents do not modify, but reversibly inhibit the desensitized enzyme as a function of time. Just as it is for the native enzyme, TPCK is a hyperbolic mixed-type inhibitor of the desensitized BChE with Ki, a and ß values of 0.017 ± 0.003 mM, 3.942 ± 1.125 and 0.524 ± 0.070, respectively. However, TLCK is the pure competitive inhibitor of the desensitized BChE with a Ki value of 0.008 ± 0.000 mM, while it is hyperbolic mixed-type inhibitor of the native form. These findings show that the conformation of the active site cavity of desensitized BChE is different from that of the native enzyme.  相似文献   

5.
Two chymotrypsin‐like serine proteinases are produced by B. bassiana 278 when grown on different carbon and nitrogen sources. By employing acetone precipitation, gel filtration and ion‐exchange chromatographies, the enzymes were separated from the culture filtrate after propagation of the fungus on medium enriched either with ground larvae of Apis mellifera (Proteinase I) or porcine blood plasma (Proteinase II). The purified enzymes with a molecular mass of approximately 32 kDa hydrolyzed natural protein substrates: casein, hide powder azure (HPA), azocoll and much less elastin Congo Red and collagen. They differ from each other in the optimum pH value, amino acid composition, Michaelis constant and susceptibility to natural chymotrypsin inhibitors. Both proteinases hydrolyze suc‐Ala‐Ala‐Pro‐Phe‐p‐NA with an apparent Km of 2.03 × 10—3 M and 1.04 × 10—4 M, respectively. The turkey ovomukoid (OMTKY) and cathepsin G/chymotrypsin inhibitor inhibit only Proteinase II from the larvae hemolymph of Apis mellifera (AMCI). The association constant of the interaction of this enzyme with AMCI was estimated to be very high (4.11 × 109 M—1).  相似文献   

6.
The nearly 50,000 known Nudix proteins have a diverse array of functions, of which the most extensively studied is the catalyzed hydrolysis of aberrant nucleotide triphosphates. The functions of 171 Nudix proteins have been characterized to some degree, although physiological relevance of the assayed activities has not always been conclusively demonstrated. We investigated substrate specificity for eight structurally characterized Nudix proteins, whose functions were unknown. These proteins were screened for hydrolase activity against a 74‐compound library of known Nudix enzyme substrates. We found substrates for four enzymes with kcat/Km values >10,000 M?1 s?1: Q92EH0_LISIN of Listeria innocua serovar 6a against ADP‐ribose, Q5LBB1_BACFN of Bacillus fragilis against 5‐Me‐CTP, and Q0TTC5_CLOP1 and Q0TS82_CLOP1 of Clostridium perfringens against 8‐oxo‐dATP and 3'‐dGTP, respectively. To ascertain whether these identified substrates were physiologically relevant, we surveyed all reported Nudix hydrolytic activities against NTPs. Twenty‐two Nudix enzymes are reported to have activity against canonical NTPs. With a single exception, we find that the reported kcat/Km values exhibited against these canonical substrates are well under 105 M?1 s?1. By contrast, several Nudix enzymes show much larger kcat/Km values (in the range of 105 to >107 M?1 s?1) against noncanonical NTPs. We therefore conclude that hydrolytic activities exhibited by these enzymes against canonical NTPs are not likely their physiological function, but rather the result of unavoidable collateral damage occasioned by the enzymes' inability to distinguish completely between similar substrate structures. Proteins 2016; 84:1810–1822. © 2016 The Authors Proteins: Structure, Function, and Bioinformatics Published by Wiley Periodicals, Inc.  相似文献   

7.
Probable participation of sperm protease in the acrosome reaction was investigated using several inhibitors and substrates. Among those examined, L-l-tosylamide-2-phenylethyl chloromethyl ketone (TPCK) and chymostatin, chymotrypsin inhibitors, p-nitrophenyl-p′-guanidinobenzoate (NPGB), a serine protease inhibitor, and N-benzoyl-L-tyrosine ethyl ester (BTEE), a chymotrypsin substrate, inhibited the egg jelly-induced acrosome reaction of Strongylocentrotus intermedius. TPCK and BTEE, however, did not inhibit the reaction caused by ionophores, A23187, or nigericin. To know the mechanism of inhibition by chymotrypsin inhibitors and substrates of the egg jelly-induced acrosome reaction, intraccllular Ca2+ concentration ([Ca2+]i) and pH (pHi) were measured with fura-2 and 2′,7′-bis (carboxy-ethyl)carboxyfluorescein (BCECF), respectively. Egg jelly caused increase of [Ca2+]i which was depressed by BTEE. Egg jelly also caused a transient rise of pHi, which was not depressed by BTEE. In the presence of verapamil, the acrosome reaction by egg jelly was significantly inhibited concomitant with depressed increase of [Ca2+]i. The rise of pHj was not depressed by verapamil. Thus, modes of action of BTEE and of verapamil are similar to each other. Bringing these findings together, the authors present a view that a chymotrypsin-like protease of sea urchin sperm activates verapamil-sensitive Ca2+ channels, which take part in the acrosome reaction.  相似文献   

8.
9.
A serine protease was isolated from midguts of the bumblebee male Bombus terrestris by a combination of precipitation procedures with column chromatography. The purified enzyme exhibited two bands with molecular masses of 25 and 26 kDa as determined by sodium dodecyl sulfate polyacrylamide gel electrophoresis. These bands showed a proteolytic activity in zymography assay. Midgut enzymes showed optimum proteolytic activity at pH 9 and 35°C using N‐succinyl‐L‐alanyl‐L‐alanyl‐L‐prolyl‐L‐phenyl‐alanine 4‐nitroanilide as a substrate. The Michaelis constant (Km) and maximum reaction rate (Vmax) were 0.55 ± 0.042 mM and 0.714 ± 0.056 μmol p‐nitroalanine produced min?1 mg protein?1, respectively. Inhibition was affected by trypsin inhibitor, but not by phenylmethylsulfonyl fluoride and N‐tosyl‐L‐phenylalanine chloromethyl ketone, which indicated the trypsin‐like but not chymotrypsin‐like specificity. The identity of the serine protease was confirmed by nanoliquid‐tandem mass spectrometry. Eleven unique peptides of the B. terrestris serine protease were found. It shows high homology to a previously reported B. ignitus serine protease covering more than 65% of the protein amino acid sequence.  相似文献   

10.
Aims: Haloarchaeal proteases function optimally in high salt (low water activity); thus, they offer an advantage over the nonhalophilic counterparts as biocatalysts for protease‐catalysed peptide synthesis. The haloalkaliphilic archaeon Natrialba magadii secretes a solvent‐tolerant protease, Nep (Natrialba magadii extracellular protease). In this work, the ability of Nep to catalyse peptide synthesis was examined. Methods and Results: The tripeptide Ac‐Phe‐Gly‐Phe‐NH2 was synthesized using Ac‐Phe‐OEt and Gly‐Phe‐NH2 substrates as building blocks in the presence of Nep, 30% (v/v) dimethyl sulfoxide (DMSO) and 1·5 or 0·5 mol l?1 NaCl. Purification and identification of the peptide product was achieved by RP‐HPLC and ESI‐MS, respectively. The native as well as the recombinant enzyme produced in Haloferax volcanii (HvNep) was similarly effective as catalysts for the synthesis of this model tripeptide with yields of up to 60% and without secondary hydrolysis of the product. HvNep catalysed the synthesis of various tripeptides with preference for those having aromatic amino acids in the P1 site. Conclusion: Nep is able to catalyse peptide synthesis under different salt concentrations in the presence of DMSO. Significance and Impact of Study: The catalytic property of Nep in peptide synthesis combined with overproduction of this protease in Hfx. volcanii anticipates the potential applicability of this haloarchaeal protease in biotechnology.  相似文献   

11.
1. Two chymotrypsins with isoelectric points pI 6.2 and 5.8 were purified from the pyloric caeca of Atlantic cod using a phenyl-Sepharose column and chromatofocusing chromatography. The apparent molecular weight was 26,000 as judged by SDS-polyacrylamide gel electrophoresis and gel filtration. 2. The cod enzymes differed from bovine chymotrypsin in having a slightly higher molecular weight and more acidic pI points. N-terminal amino acid sequence analysis of cod chymotrypsin B showed considerable similarity with bovine chymotrypsin. 3. Heat stability and stability towards acidic pH were reduced in the cod enzymes. Generally, the cod and bovine chymotrypsins responded similarly to various protease inhibitors. However, the cod chymotrypsins were less sensitive to aprotinin inhibition but more sensitive towards soybean trypsin inhibitor and cysteine. 4. Kinetic properties were examined and the cod enzymes found to be more active towards both ester (N-benzoyl-tyrosine ethyl ester) and amide (N-benzoyl-tyrosine-p-nitroanilide) substrates. The observed differences in kinetic properties are indicative of an adaptive response towards the low temperature environment in which the cod lives.  相似文献   

12.
The emergence of Gram‐negative “superbugs” exhibiting resistance to known antibacterials poses a major public health concern. Low molecular weight Gram‐negative antibacterials are believed to penetrate the outer bacterial membrane (OM) through porin channels. Therefore, intracellular exposure needed to drive antibacterial target occupancy should depend critically on the translocation rates through these proteins and avoidance of efflux pumps. We used electrophysiology to study the structure‐translocation kinetics relationships of a set of carbapenem antibacterials through purified porin OmpC reconstituted in phospholipid bilayers. We also studied the relative susceptibility of OmpC+ and OmpC‐ E. coli to these compounds as an orthogonal test of translocation. Carbapenems exhibit good efficacy in OmpC‐expressing E. coli cells compared with other known antibacterials. Ertapenem, which contains an additional acidic group compared to other analogs, exhibits the fastest entry into OmpC (kon ≈ 2 × 104 M?1 s?1). Zwitterionic compounds with highly polar groups attached to the penem‐2 ring, including panipenem, imipenem and doripenem exhibit faster kon (>104 M?1 s?1), while meropenem and biapenem with fewer exposed polar groups exhibit slower kon (~5 × 103 M?1 s?1). Tebipenem pivoxil and razupenem exhibit ~13‐fold slower kon (~1.5 × 103 M?1 s?1) than ertapenem. Overall, our results suggest that (a) OmpC serves as an important route of entry of these antibacterials into E. coli cells; and (b) that the structure‐kinetic relationships of carbapenem translocation are governed by H‐bond acceptor/donor composition (in accordance with our previous findings that the enthalpic cost of transferring water from the constriction zone to bulk solvent increases in the presence of exposed nonpolar groups). Proteins 2014; 82:2998–3012. © 2014 Wiley Periodicals, Inc.  相似文献   

13.
The interactions of chymotrypsin, subtilisin and trypsin with a low MW proteinase inhibitor from potatoes were investigated. The Ki value calculated for the binding of inhibitor to chymotrypsin was 1.6 ± 0.9 × 10?10M, while the second-order rate constant for association was 6 × 105 M?1/sec. Although binding was not observed to chymotrypsin which had been treated with diisopropyl fluorophosphate or with l-tosylamide-2-phenylethyl chloromethyl ketone, the 3-methylhistidine-57 derivative bound inhibitor with a Ki value of 9.6 × 10?9 M. The inhibitor also exhibited a tight association with subtilisin (Ki < 4 × 10?9 M). In contrast, little inhibition of trypsin was observed, and this was believed to be due to low levels of a contaminant in our preparations. No evidence for reactive site cleavage was observed after incubation of the inhibitor with catalytic amounts of chymotrypsin or subtilisin at acid pH.  相似文献   

14.
Protease inhibitors affecting the activity of the proteasome were reported to induce programmed cell death (apoptosis) in some mammalian cell lines. Proteasome activity can be suppressed by specific peptide derivatives and by N‐tosyl‐lysine‐chloromethyl‐ketone (TLCK) and N‐tosyl‐phenylalanine‐chloromethyl‐ketone (TPCK), which affect the trypsine‐ and chymotrypsine‐like activities of the proteasome, respectively. Particularly TLCK and TPCK caused necrotic cell death in the unicellular green alga Chlamydomonas reinhardtii. As a control, the effects of these protease inhibitors on the survival of human WISH cells were also studied. Bleaching of the Chlamydomonas cells after addition of TLCK or TPCK indicated that reactive oxygen species (ROS) were involved in this process. Indeed, increased levels of ROS were detected in Chlamydomonas cells treated with TLCK or TPCK. Furthermore, cell death induced by these protease inhibitors was accelerated by illumination and prevented or slowed down by scavengers of ROS.  相似文献   

15.
The use of 6-(N-acetyl-L -phenylalanyl)-aminoluciferin as a novel substrate for α-chymotrypsin has been demonstrated. The kinetic parameters determined are KM = 0.38mmol/L, kcat = 6.5 s?1 and kcat/kM = 17,100 (L/mols). The test principle of the coupled assay is the release of aminoluciferin by enzymatic cleavage of 6-(N-acetyl-L -phenylalanyl)-aminoluciferin. Aminoluciferin is oxidized, with light emission, by firefly luciferase (Photinus pyralis) and can be quantified in a luminometric assay. The detection limit for chymotrypsin was found to be 0.3 ng per assay. 6-(N-acetyl-L -phenylalanyl)-aminoluciferin has been synthesized as an example for a new class of highly sensitive substrates. By modification of the peptide residue these new substrates may be suitable for ultrasensitive detection of different proteinases.  相似文献   

16.
An aspartic protease that is significantly produced by baculovirus-infected Spodoptera frugiperda Sf9 insect cells was purified to homogeneity from a growth medium. To monitor aspartic protease activity, an internally quenched fluoresce (IQF) substrate specific to cathepsin D was used. The purified aspartic protease showed a single protein band on SDS–PAGE with an apparent molecular mass of 40 kDa. The N-terminal amino acid sequence of the enzyme had a high homology to a Bombyx mori aspartic protease. The enzyme showed greatest affinity for the IQF substrate at pH 3.0 with a K m of 0.85 μM. The k cat and k cat?K m values were 13 s?1 and 15 s?1 μM?1 respectively. Pepstatin A proved to be a potent competitive inhibitor with inhibitor constant, K i, of 25 pM.  相似文献   

17.
It is demonstrated that cyanobacteria (both azotrophic and non‐azotrophic) contain heme b oxidoreductases that can convert chlorite to chloride and molecular oxygen (incorrectly denominated chlorite ‘dismutase’, Cld). Beside the water‐splitting manganese complex of photosystem II, this metalloenzyme is the second known enzyme that catalyses the formation of a covalent oxygen–oxygen bond. All cyanobacterial Clds have a truncated N‐terminus and are dimeric (i.e. clade 2) proteins. As model protein, Cld from Cyanothece sp. PCC7425 (CCld) was recombinantly produced in Escherichia coli and shown to efficiently degrade chlorite with an activity optimum at pH 5.0 [kcat 1144 ± 23.8 s?1, KM 162 ± 10.0 μM, catalytic efficiency (7.1 ± 0.6) × 106 M?1 s?1]. The resting ferric high‐spin axially symmetric heme enzyme has a standard reduction potential of the Fe(III)/Fe(II) couple of ?126 ± 1.9 mV at pH 7.0. Cyanide mediates the formation of a low‐spin complex with kon = (1.6 ± 0.1) × 105 M?1 s?1 and koff = 1.4 ± 2.9 s?1 (KD ~ 8.6 μM). Both, thermal and chemical unfolding follows a non‐two‐state unfolding pathway with the first transition being related to the release of the prosthetic group. The obtained data are discussed with respect to known structure–function relationships of Clds. We ask for the physiological substrate and putative function of these O2‐producing proteins in (nitrogen‐fixing) cyanobacteria.  相似文献   

18.
Caspases are a family of cysteine‐dependent proteases known to be involved in the process of programmed cell death in metazoans. Recently, cyanobacteria were also found to contain caspase‐like proteins, but their existence has only been identified in silico up to now. Here, we present the first experimental characterisation of a prokaryotic caspase homologue. We have expressed the putative caspase‐like gene MaOC1 from the toxic bloom‐forming cyanobacterium Microcystis aeruginosa PCC 7806 in Escherichia coli. Kinetic characterisation showed that MaOC1 is an endopeptidase with a preference for arginine in the P1 position and a pH optimum of 7.5. MaOC1 exhibited high catalytic rates with the kcat/KM value for Z‐RR‐AMC substrate of the order 106 M?1 s?1. In contrast to plant or metazoan caspase‐like proteins, whose activity is calcium‐dependent or requires dimerisation for activation, MaOC1 was activated by autocatalytic processing after residue Arg219, which separated the catalytic domain and the remaining 55 kDa subunit. The Arg219Ala mutant was resistant to autoprocessing and exhibited no proteolytic activity, confirming that processing of MaOC1 is a prerequisite for its activity. Due to their structural and functional differences to other known caspase‐like proteins, we suggest to name these evolutionary primitive proteins orthocaspases.  相似文献   

19.
The bacterial serine protease, SGPB, was inhibited by two specific tripeptide chloromethyl ketones, N-t-butyloxycarbonyl-l-alanylglycyl-l-phenylalanine chloromethyl ketone (BocAGFCK) and N-t-butyloxycarbonyl-glycyl-l-leucyl-l-phenylalanine chloromethyl ketone (BocGLFCK). Crystals of the inhibited complexes were grown and examined by X-ray crystallographic methods. The peptide backbone of each inhibitor is bound by three hydrogen bonds to the main chain of residues Ser214 to Gly216. There are two well-characterized hydrophobic pockets, S1 and S2, on the surface of SGPB which accommodate the P1 and P2 side-chains of the BocGLFCK inhibitor. A conformational change of Tyr171 is induced by the binding of this inhibitor. Both inhibitors make two covalent bonds to the SGPB enzyme. The imidazole ring of His57 is alkylated at the N?2 atom and Oγ of Ser195 forms a hemiketal bond with the carbonyl-carbon atom of the inhibitor. Comparison of the binding modes of the two tripeptides in conjunction with the differences in their inhibition constants (KI) allows one to estimate the binding energy of the leucyl side-chain as ?2.6 kcal mol?1. The importance of an electrophilic component in the serine protease mechanism, which involves the polarization of the susceptible carbonyl bond of a substrate or inhibitor by the peptide NH groups of Gly193 and Ser195 is discussed.  相似文献   

20.
Lu Z  Dunaway-Mariano D  Allen KN 《Proteins》2011,79(11):3099-3107
Analysis of the haloalkanoate dehalogenase superfamily (HADSF) has uncovered homologues occurring within the same organism that are found to possess broad, overlapping substrate specificities, and low catalytic efficiencies. Here we compare the HADSF phosphatase BT1666 from Bacteroides thetaiotaomicron VPI‐5482 to a homologue with high sequence identity (40%) from the same organism BT4131, a known hexose‐phosphate phosphatase. The goal is to find whether these enzymes represent duplicated versus paralogous activities. The X‐ray crystal structure of BT1666 was determined to 1.82 Å resolution. Superposition of the BT1666 and BT4131 structures revealed a conserved fold and identical active sites suggestive of a common physiological substrate. The steady‐state kinetic constants for BT1666 were determined for a diverse panel of phosphorylated metabolites to define its substrate specificity profile and overall level of catalytic efficiency. Whereas BT1666 and BT4131 are both promiscuous, their substrate specificity profiles are distinct. The catalytic efficiency of BT1666 (kcat/Km = 4.4 × 102M?1 s?1 for the best substrate fructose 1,6‐(bis)phosphate) is an order of magnitude less than that of BT4131 (kcat/Km = 6.7 × 103M?1 s?1 for 2‐deoxyglucose 6‐phosphate). The seemingly identical active‐site structures point to sequence variation outside the active site causing differences in conformational dynamics or subtle catalytic positioning effects that drive the divergence in catalytic efficiency and selectivity. The overlapping substrate profiles may be understood in terms of differential regulation of expression of the two enzymes or a conferred advantage in metabolic housekeeping functions by having a larger range of possible metabolites as substrates. Proteins 2011;. © 2011 Wiley‐Liss, Inc.  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