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 exchange of 18O between H2O and long-chain free fatty acids is catalyzed by pancreatic carboxylester lipase (EC 1.1.1.13). For palmitic, oleic, and arachidonic acid in aqueous suspension and for 13,16-cis,cis-docosadienoic acid (DA) in monomolecular films, carboxyl oxygens were completely exchanged with water oxygens of the bulk aqueous phase. With enzyme at either substrate or catalytic concentrations in the argon-buffer interface, the exchange of DA oxygens obeyed a random sequential mechanism, i.e., 18O,18O-DA in equilibrium with 18O,16O-DA in equilibrium with 16O,16O-DA. This indicates that the dissociation of the enzyme-DA complex is much faster than the rate-limiting step in the overall exchange reaction. Kinetic analysis of 18O exchange showed a first-order dependence on surface enzyme and DA concentrations, i.e., the reaction was limited by the acylation rate. The values of kcat/Km, 0.118 cm2 pmol-1 s-1, for the exchange reaction was comparable to that for methyl oleate hydrolysis and 5-fold higher than that for cholesteryl oleate hydrolysis in monolayers [Bhat, S., & Brockman, H. L. (1982) Biochemistry 21, 1547]. Thus, fatty acids are good "substrates" for carboxylester lipase. With substrate levels of carboxylester lipase in the interfacial phase, the acylation rate constant kcat/Km was 200-fold lower than that obtained with catalytic levels of enzyme. This suggests a possible restriction of substrate diffusion in the protein-covered substrate monolayer.  相似文献   

2.
This paper describes a heavy isotope coding strategy for the analysis of all types of tryptic peptides, including those that are N-terminally blocked and from the C-terminus of proteins. The method exploits differential derivatization of amine and carboxyl groups generated during proteolysis as a means of coding. Carboxyl groups produced during proteolysis incorporate 18O from H218O. Peptides from the C-terminus of proteins were not labeled with 18O unless they contained a basic C-terminal amino acid. Primary amines from control and experimental samples were differentially acylated after proteolysis with either 1H3- or 2H3-N-acetoxysuccinamide. When these two types of labeling were combined, unique coding patterns were achieved for peptides arising from the C-termini and blocked N-termini of proteins. This method was used to (1) distinguish C-terminal peptides in model proteins, (2) recognize N-terminal peptides from proteins in which the amino terminus is acylated, and (3) identify primary structure variations between proteins from different sources.  相似文献   

3.
Phosphoenolpyruvate when heated in acidic solution exchanges its phosphoryl and carboxyl oxygens rapidly and its enolic oxygen much more slowly with oxygens from water. The incorporation of 18O into phosphoenolpyruvate was measured by gas chromatography-mass spectrometry and phosphorus-31 nuclear magnetic resonance after heating in H218O at 98 degrees C. The rates of exchange of all six oxygens of phosphoenolpyruvate with water increase with increasing acidity, and the phosphoryl oxygens exchange more rapidly than the carboxyl oxygens. The rate of exchange of each oxygen of the phosphoryl group is 16-fold greater than the hydrolysis rate at 1 N HCl. This provides a simple and useful method for the synthesis of [18O]phosphoenolpyruvate highly enriched in its phosphoryl-group oxygens. An enrichment of 89% was obtained with a 50% yield. The [18O]-phosphoenolpyruvate showed a binomial distribution of 18O in the phosphoryl-group oxygens. The exchange may be explained by the reversible formation of a transient cyclic phosphate and, for exchange of the enolic oxygen, a transient acyl phosphate. Preparation of [18O]phosphoenolypyruvate from [18O]Pi by a chemical synthesis from beta-chlorolactate was not satisfactory because of drastic loss of 18O during the procedures used. Some loss of 18O also occurred during an enzymic synthesis with KCNO, [18O]Pi, carbamate kinase, and pyruvate kinase.  相似文献   

4.
A series of 2-aroylthiophenes derived from tienilic acid by replacement of its OCH2COOH substituent with groups bearing various functions have been synthesized and studied as possible substrates of recombinant human liver cytochrome P450s 2C9 and 2C18 expressed in yeast. Whereas only compounds bearing a negative charge acted as substrates of CYP 2C9 and were hydroxylated at position 5 of their thiophene ring at a significant rate, many neutral 2-aroylthiophenes were 5-hydroxylated by CYP 2C18 with kcat values of >2 min-1. Among the various compounds that were studied, those bearing an alcohol function were the best CYP 2C18 substrates. One of them, compound 3, which bears a terminal O(CH2)3OH function, appeared to be a particularly good substrate of CYP 2C18. It was regioselectively hydroxylated by CYP 2C18 at position 5 of its thiophene ring with a KM value of 9 +/- 1 microM and a kcat value of 125 +/- 25 min-1, which are the highest described so far for a CYP 2C. A comparison of the oxidations of 3, by yeast-expressed CYP 1A1, 1A2, 2C8, 2C9, 2C18, 2C19, 2D6, 2E1, 3A4, and 3A5, showed that only CYP 2C8, 2C18, and 2C19 were able to catalyze the 5-hydroxylation of 3. However, the catalytic efficiency of CYP 2C18 for that reaction was considerably higher (kcat/KM value being 3-4 orders of magnitude larger than those found for CYP 2C8 and 2C19). Several human P450s exhibited small activities for the oxidative O-dealkylation of 3. The four recombinant CYP 2Cs were the best catalysts for that reaction (kcat between 1 and 5 min-1) when compared to all the P450s that were tested, even though it is a minor reaction in the case of CYP 2C18. All these results show that compound 3 is a new, selective, and highly efficient substrate for CYP 2C18 that should be useful for the study of this P450 in various organs and tissues. They also suggest some key differences between the active sites of CYP 2C9 and CYP 2C18 for substrate recognition.  相似文献   

5.
We have investigated the mechanism by which the complement protease, Factor D, achieves its high specificity for the cleavage of Factor B in complex with C3(H2O). Kinetic experiments showed that Factor B and C3(H2O) associate with a KD of >/=2.5 microM and that Factor D acts on this complex with a second-order rate constant of kcat/KM >/= 2 x 10(6) M-1 s-1, close to the rate of a diffusion-controlled reaction for proteins of this size. In contrast, Factor D, which is a member of the trypsin family of serine proteases, was 10(3)-10(4)-fold less active than trypsin toward both thioester and p-nitroanilide substrates containing an arginine at P1. Furthermore, peptides spanning the Factor B cleavage site were not detectably cleaved by Factor D (kcat/KM /=9 kcal/mol of binding energy to stabilize the transition state for reaction. In support of this, we demonstrate that chemical modification of Factor D at a single lysine residue that is distant from the active site abolishes the activity of the enzyme toward Factor B while not affecting activity toward small synthetic substrates. We propose that Factor D may exemplify a special case of the induced fit mechanism in which the requirement for conformational activation of the enzyme results in a substantial increase in substrate specificity.  相似文献   

6.
The potential capabilities of a new proteolytic 18O labeling method employing peptidyl-Lys metalloendopeptidase (Lys-N) have been demonstrated for use in comparative proteomics. Conditions (pH>or=9.5) have been found such that Lys-N incorporates only a single 18O atom into the carboxyl terminus of each proteolytically generated peptide. This 18O labeling method has a major advantage over current protelytic 18O labeling methods that generate a mixture of isotopic isoforms resulting from the incorporation of one or two 18O atoms into each peptide species by the proteases (trypsin, Lys-C, or Glu-C) used. We demonstrate that the single 18O atom incorporation property of Lys-N overcomes the major problem of the current proteolytic 18O labeling methods and provides accurate quantification results for isotopically labeled peptides.  相似文献   

7.
H2(18)O isotope exchange into specifically 13C-labeled substrate was used to obtain information on the rate-limiting step in the action of the phospholipase A2 from the venom of the Indian cobra (Naja naja naja). Incorporation of 18O was detected by the effect of 18O on 13C chemical shifts in 13C NMR. The enzymatic hydrolysis of a micellar phosphatidylcholine analogue of platelet-activating factor 1-alkyl-2-[1-13C]lauroyl-sn-glycero-3-phosphorylcholine proceeds by an O-acyl cleavage of the sn-2 ester bond. The reaction was examined for simultaneous 18O incorporation into the substrate. No exchange was found, suggesting that the hydrolytic step is not followed by a higher energy transition state and that it or a step before it appears to be rate-limiting. Previous experiments on phosphatidylethanolamine activation indicate that kcat is altered but that the km remains the same upon activation, suggesting that the binding steps occurring before the hydrolytic step are not affected. This strongly suggests that the hydrolytic step is in fact the rate-limiting step under these conditions. The 13C, 18O NMR technique should be generally applicable to mechanistic questions of this type.  相似文献   

8.
Stable isotopes are essential tools in biological mass spectrometry. Historically, 18O-stable isotopes have been extensively used to study the catalytic mechanisms of proteolytic enzymes1-3. With the advent of mass spectrometry-based proteomics, the enzymatically-catalyzed incorporation of 18O-atoms from stable isotopically enriched water has become a popular method to quantitatively compare protein expression levels (reviewed by Fenselau and Yao4, Miyagi and Rao5 and Ye et al.6). 18O-labeling constitutes a simple and low-cost alternative to chemical (e.g. iTRAQ, ICAT) and metabolic (e.g. SILAC) labeling techniques7. Depending on the protease utilized, 18O-labeling can result in the incorporation of up to two 18O-atoms in the C-terminal carboxyl group of the cleavage product3. The labeling reaction can be subdivided into two independent processes, the peptide bond cleavage and the carboxyl oxygen exchange reaction8. In our PALeO (protease-assisted labeling employing 18O-enriched water) adaptation of enzymatic 18O-labeling, we utilized 50% 18O-enriched water to yield distinctive isotope signatures. In combination with high-resolution matrix-assisted laser desorption ionization time-of-flight tandem mass spectrometry (MALDI-TOF/TOF MS/MS), the characteristic isotope envelopes can be used to identify cleavage products with a high level of specificity. We previously have used the PALeO-methodology to detect and characterize endogenous proteases9 and monitor proteolytic reactions10-11. Since PALeO encodes the very essence of the proteolytic cleavage reaction, the experimental setup is simple and biochemical enrichment steps of cleavage products can be circumvented. The PALeO-method can easily be extended to (i) time course experiments that monitor the dynamics of proteolytic cleavage reactions and (ii) the analysis of proteolysis in complex biological samples that represent physiological conditions. PALeO-TimeCourse experiments help identifying rate-limiting processing steps and reaction intermediates in complex proteolytic pathway reactions. Furthermore, the PALeO-reaction allows us to identify proteolytic enzymes such as the serine protease trypsin that is capable to rebind its cleavage products and catalyze the incorporation of a second 18O-atom. Such "double-labeling" enzymes can be used for postdigestion 18O-labeling, in which peptides are exclusively labeled by the carboxyl oxygen exchange reaction. Our third strategy extends labeling employing 18O-enriched water beyond enzymes and uses acidic pH conditions to introduce 18O-stable isotope signatures into peptides.  相似文献   

9.
The mechanism of the enzyme asparaginase (L-asparagine amidohydrolase, EC 3.5.1.1) from Escherichia coli was examined using 13C NMR spectroscopy. The pH-dependent oxygen exchange reactions between water and aspartic acid were followed by use of the 18O isotope-induced shift of the resonance positions of directly bonded 13C nuclei. Both L-1- and L-1,4-[13C]aspartic acid were used in experiments with previously 18O-labeled aspartic acid, or in experiments involving the use of 18O-labeled solvent water. Asparaginase catalyzes a relatively efficient exchange between the oxygens of water and those on one carboxyl group of aspartic acid. Exchange at C-4 occurs rapidly but, within experimental error, no exchange at C-1 could be detected. These and related experiments involving the position of 18O incorporation during hydrolysis of aspartic acid beta-methyl ester are all consistent with possible acyl-enzyme mechanisms involving C-4, but do not support a free aspartic acid anhydride mechanism.  相似文献   

10.
The pH dependency of the carboxyl oxygen exchange reaction catalyzed by lysyl endopeptidase (Lys-C) and trypsin has been studied. The reaction was quantitatively monitored by measuring the incorporation of 18O atom into the alpha-carboxyl group of N(alpha)-acetyl-L-lysine from H2(18)O solvent. The optimum pHs of the carboxyl oxygen exchange reaction catalyzed by Lys-C and trypsin were found to be pH 5.0 and 6.0, respectively, which were significantly shifted toward acidic pHs compared to the most favorable pHs of their amidase activities for N(alpha)-acetyl-L-lysine amide in the pHs examined. Steady-state kinetics parameters were also determined for both enzymes at two different pHs, one at the pH optimum for their carboxyl oxygen exchange activity (pH 5-6) and the other at the favorable pH for their amidase activity (pH 8-9). Significantly lower Km (2-fold lower for Lys-C, 3-fold lower for trypsin), and higher kcat values (1.5-fold higher for Lys-C, 5-fold higher for trypsin) were obtained at the acidic pHs compared to the alkaline pHs, suggesting that Lys-C and trypsin have higher substrate binding affinities and higher catalytic rates at the acidic pHs than at the alkaline pHs. The higher carboxyl oxygen exchange activities at the acidic pHs were also confirmed with peptide substrates derived from apomyoglobin. These findings are significant toward the goal of improving the efficiency of the Lys-C and trypsin catalyzed 18O labeling reactions and are thus pertinent to improving the accuracy and reliability of quantitative proteomic experiments utilizing 18O labeling.  相似文献   

11.
Recently, proteolytic 18O labeling has been demonstrated as a promising strategy for comparative proteomic studies (Yao, X.; Freas, A.; Ramirez, J.; Demirev, P. A.; Fenselau, C. Anal. Chem. 2001, 73, 2836-42). In this approach, protein mixtures are digested in parallel in H216O and H218O and the ratios of isotopically distinct peptide products are measured by mass spectrometry. In the initial report from this laboratory, trypsin was shown to catalyze incorporation of two 18O atoms into the carboxyl terminus of each new peptide formed by cleavage of the adenovirus proteome. In the present study, a second enzyme, endoprotease Glu-C, is evaluated as an agent for cleavage and labeling. Proteolytic 18O labeling by Glu-C is shown to occur readily with phosphorylated and glycosylated proteins and with cysteinealkylated and disulfide-linked proteins. A sequential double-labeling strategy is used to characterize N-linked glycopeptides. Labeled and unlabeled peptide pairs are found to coelute chromatographically, and measurements of isotope ratios by nanospray and capillary LC-MS are found to be accurate and precise.  相似文献   

12.
16O/18O labeling is one differential proteomics technology among many that promises diagnostic and prognostic biomarkers of disease. Although the incorporation of 18O in the C-terminal carboxyl group during endoproteinase digestion in the presence of H2 18O makes the process of labeling facile, the ease and effectiveness of label incorporation have in some regards been outweighed by the difficulties in interpreting the resulting spectra. Complex isotope patterns result from the composition of unlabeled (18O(0)), singly labeled (18O(1)), and doubly labeled species (18O(2)) as well as contributions from the naturally occurring isotopes (e.g. 13C and 15N). Moreover because labeling is enzymatic, the number of 18O atoms incorporated can vary from peptide to peptide. Finally it is difficult to distinguish highly up-regulated from highly down-regulated or C-terminal peptides. We have developed an algorithm entitled regression analysis applied to mass spectrometry (RAAMS) that automatically, rapidly, and confidently interprets spectra of 18O-labeled peptides without requiring chemical composition information derived from product ion spectra. The algorithm is able to measure the effective 18O incorporation rate due to variable enzyme substrate specificity of the pseudosubstrate during the isotope exchange reaction and corrects for the 18O(0) abundance that remains in the labeled sample when using a two-step digestion/labeling procedure. We have also incorporated a method for distinguishing pure 18O(0) from pure 18O(2) peptides utilizing impure H2 18O. The algorithm operates on centroided peak lists and is therefore very fast: nine chromatograms of, on average, 1,168 spectra and containing, on average, 6,761 isotopic clusters were interpreted in, on average, 45 s per chromatogram. RAAMS is fast enough (average, 38 ms/spectrum) to allow the possibility of performing information-dependent MS/MS on a chromatographic time scale on species exceeding predetermined ratio thresholds. We describe in detail the operation of the algorithm and demonstrate its use on datasets with known and unknown ratios.  相似文献   

13.
18O labeling studies on the catechol 1,2-dioxygenase-catalyzed oxidative cleavage of pyrogallol demonstrate that the enzyme functions both as a dioxygenase and a monooxygenase in this reaction. Two products are observed, 2-pyrone-6-carboxylic acid, 99% singly labeled at the carboxylate, and 2-hydroxy-cis,cis-muconic acid, 74% doubly labeled (one 18O at each carboxylate) and 24% single labeled (one 18O at either carboxylate). The labeling pattern observed shows that 2-pyrone-6-carboxylic acid cannot be derived enzymatically from the lactonization of the 2-hydroxy-cis,cis-muconic acid, thus eliminating the dioxetane as an intermediate in the dioxygenase mechanism. The observations are interpreted to indicate the intermediacy of 2-hydroxymuconic anhydride. This anhydride or the corresponding muconyl enzyme species must be sufficiently long-lived to allow the exchange of labeled hydroxide with solvent. Evidence for mechanism-based enzyme inactivation by a pyrogallol-derived intermediate is also presented.  相似文献   

14.
15.
Baer BR  Kunze KL  Rettie AE 《Biochemistry》2007,46(41):11598-11605
Cytochrome P450s in the CYP4 family covalently bind their heme prosthetic group to a conserved acidic I-helix residue via an autocatalytic oxidation. This study was designed to evaluate the source of oxygen atoms in the covalent ester link in CYP4B1 enzymes labeled with [18O]glutamate and [18O]aspartate. The fate of the heavy isotope was then traced into wild-type CYP4B1 or the E310D mutant-derived 5-hydroxyhemes. Glutamate-containing tryptic peptides of wild-type CYP4B1 were found labeled to a level of 11-13% 18O. Base hydrolysis of labeled protein released 5-hydroxyheme which contained 12.8 +/- 1.9% 18O. Aspartate-containing peptides of the E310D mutant were labeled with 6.0-6.5% 18O, but as expected, no label was transmitted to recovered 5-hydroxyheme. These data demonstrate that the oxygen atom in 5-hydroxyheme derived from wild-type CYP4B1 originates in Glu310. Stoichiometric incorporation of the heavy isotope from the wild-type enzyme supports a perferryl-initiated carbocation mechanism for covalent heme formation in CYP4B1.  相似文献   

16.
We describe a protocol for quantitative labeling of tryptic peptides with oxygen-18. Proteins are first digested in natural water with trypsin, the pH is then lowered to 4.5 and the mixture is dried. Oxygen-18 water is added and two oxygen-18 atoms are incorporated at the peptides' carboxyl termini. Trypsin is finally inactivated by cysteine alkylation under denaturing conditions, which blocks oxygen back-exchange. The general value of this labeling strategy for differential proteomics is illustrated by the analysis and identification of several couples of differently labeled amino terminal peptides isolated from a human platelet proteome by a previously described chromatographic procedure.  相似文献   

17.
Recently, there has been an increased interest in isotopical labeling of peptides. Although there are several techniques allowing for a complete labeling of all carboxyl groups in peptides, regioselective labeling would be beneficial in many situations. Such labeling requires the use of 18O‐labeled Fmoc amino acids. We have designed a method for such labeling that is an improvement on a technique proposed earlier. The new procedure is suitable for microscale synthesis and could be used in peptide and proteomics laboratories. Although for the majority of tested amino acids our method gives good labeling efficiency, it is time consuming. Therefore, we have decided to use microwave‐assisted procedure. This approach resulted in reduction of reaction time to 15 min and increased reaction efficiency. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
A direct continuous UV-Vis spectrophotometric assay has been developed for VanX, a D-alanyl-D-alanine aminodipeptidase necessary for vancomycin resistance. This method is based on the hydrolysis of the alternative substrate D-alanyl-alpha-(R)-phenylthio-glycine D-Ala-D-Gly(S-Ph)-OH (H-DAla-DPsg-OH, 5a). Spontaneous decomposition of the released phenylthioglycine generates thiophenol, which is quantified using Ellman's reagent. The dipeptide behaved as an excellent substrate of VanX, exhibiting Michaelis-Menten kinetics with a kcat of 76 +/- 5/s and a KM of 0.83 +/- 0.08 mm (kcat = 46 +/- 3/s and KM = 0.11 +/- 0.01 mm for D-Ala-D-Ala). Determination of the kinetic parameters of the previously reported mechanism-based inhibitor D-Ala-D-Gly(SPhip-CHF2)-OH (H-D-Ala-DPfg-OH, 5c) [Araoz, R., Anhalt, E., René, L., Badet-Denisot, M.-A., Courvalin, P. & Badet, B. (2000) Biochemistry 39, 15971-15979] using the substrate reported in the present study yielded values of Kirr of 22 +/- 1 microM and kinact of 9.3 +/- 0.4/min in good agreement with values previously obtained in our laboratory (Kirr = 30 +/- 1 mm; kinact = 7.3 +/- 0.3/min). In addition, inhibition by the competing substrate D-Ala-D-Ala resulted in determination of a Ki = 70 +/- 6 microM close to the previously reported KM value. These results demonstrate that the present assay is a convenient, rapid and sensitive tool in the search for VanX inhibitors.  相似文献   

19.
K K Shukla  H M Levy 《Biochemistry》1977,16(1):132-136
The gamma-phosphoryl groups of two intermediates (M-ATP and M-ADP-P1) in the pathway of MgATP hydrolysis by myosin undergo extensive oxygen exchange with water. Actin activates the overall rate of hydrolysis at a rate-limiting step which follows these exchange reactions. Thus, actin, by decreasing the turnover time of hydrolysis, would be expected to proportionately decrease the time available for oxygen exchange. Using subfragment 1 of myosin, the turnover time of hydrolysis can be varied over a wide range by changing the concentration of actin. An estimate for the rate constant of exchange can then be obtained by relating these turnover times to measured values for oxygen exchange (incorporation of 18O from H218O into the inorganic phosphate (Pi) released by hydrolysis). The results of such an experiment, with turnover times between 0.2 and 25 s, indicate that, for each gamma-phosphoryl group, one oxygen from the medium is added rapidly (to cleave the phosphoryl group or form a pentacoordinate phosphroyl complex); two more oxygens exchange with a rate constant, kc, of about 1 s-1; and a fourth oxygen exchanges slowly with ke about 0.2 s-1. The higher value is about 18 times smaller than the rate constant, 5-3, for the reverse cleavage step of the myosin pathway, which is postulated to be responsible for oxygen exchange. The data, then, indicate that the rate-limiting step for oxygen exchange is not k-3, but may be the rate of rotation of oxygens around the phosphorus atom, with one oxygen severely restricted by its binding to the active site. The finding that kc differs for the four oxygens in each phosphate group is related to past observations on myosin-catalyzed oxygen exchange.  相似文献   

20.
The kinetic parameters kcat, KM, and kcat/KM were determined at 25 degrees C and pH 4.5, 5.5, and 6.0 for the series of penicillopepsin substrates Ac-Alam-Lys-(NO2)Phe-Alan-amide, where (NO2)Phe is p-nitrophenylalanine and m and n equal 0-3. KM values at pH 6.0 were the same for all 12 peptides and averaged 0.088 +/- 0.02 mM but increased to different degrees at lower pH. In contrast, kcat values increased with increasing chain length. At pH 6 and at the pH optimum of kcat, the largest increases (about 37-fold on average) were obtained when alanine residues were added in positions P2' and P3. Only 1-2-fold increases were observed for positions P2, P3', P4, and P4'. These results show that occupation of subsites S2' and S3 is largely responsible for the rate enhancements caused by secondary substrate interactions with this series of peptides. Additional support for an important role of subsite S3 comes from the observation that the two peptides where m = 1 and n = 1 or 2, respectively, are cleaved not only between lysine and p-nitrophenylalanine but also between the latter and alanine, suggesting that occupation of subsite S3 by the N-terminal alanine overcomes the unfavorable interaction of alanine in subsite P1'. Subsite S3 is also important in the binding of pepstatin analogues and in transpeptidation reactions. It is proposed that the roles of subsites S3 and S2' are to facilitate the conversion of the first enzyme-substrate complex into a productive complex and to assist in the distortion of the scissile bond.(ABSTRACT TRUNCATED AT 250 WORDS)  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