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 218 毫秒
1.
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   

2.
Solvent exchange of 18O-labeled buried water in bovine pancreatic trypsin inhibitor (BPTI), trypsin, and trypsin-BPTI complex is measured by high-precision isotope ratio mass spectrometry. Buried water is labeled by equilibration of the protein in 18O-enriched water. Protein samples are then rapidly dialyzed against water of normal isotope composition by gel filtration and stored. The exchangeable 18O label eluting with the protein in 10-300 s is determined by an H2O-CO2 equilibration technique. Exchange of buried waters with solvent water is complete before 10-15 s in BPTI, trypsin, and BPTI-trypsin, as well as in lysozyme and carboxypeptidase measured as controls. When in-exchange dialysis and storage are carried out at pH greater than or equal to 2.5, trypsin-BPTI and trypsin, but not free BPTI, have the equivalent of one 18O atom that exchanges slowly (after 300 s and before several days). This oxygen is probably covalently bound to a specific site in trypsin. When in-exchange dialysis and storage are carried out at pH 1.1, the equivalent of three to seven 18O atoms per molecule is associated with the trypsin-BPTI complex, apparently due to nonspecific covalent 18O labeling of carboxyl groups at low pH. In addition to 18O exchange of buried waters, the hydrogen isotope exchange of buried NH groups H bonded to buried waters was also measured. Their base-catalyzed exchange rate constants are on the order of NH groups that in the crystal are exposed to solvent (static accessibility greater than 0) and hydrogen-bonded main chain O, and their pH min is similar to that for model compounds. The pH dependence of their exchange rate constants suggests that direct exchange with water may significantly contribute to their observed exchange rate.  相似文献   

3.
Soybean nodule bacteroids and Bradyrhizobium japonicum free-living cells induced for H2-uptake hydrogenase, actively catalyze the evolution of H2 in a reaction highly dependent on the pH. The optimal pHs for the evolution and uptake reactions were 4.0 and 7.5-8.0, respectively. No differences were found between free-living cells and bacteroids with respect to hydrogen acceptor specificity, although absolute rates of H2 uptake were higher for free-living cells. Both types of cells were able to evolve hydrogen from reduced methyl viologen at low pH. These intact cells also catalyzed the exchange reaction between tritium and water in the absence of oxygen. The pH profile of the exchange activity showed two peaks at values near the optimal pHs for the evolution and uptake reactions.  相似文献   

4.
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   

5.
K Alexander  I G Young 《Biochemistry》1978,17(22):4745-4750
The biosynthetic origin of the oxygen atoms of ubiquinone 8 from aerobically grown Escherichia coli was studied by 18O labeling. An apparatus was developed which allowed the growth of cells under a defined atmosphere. Mass spectral analysis of ubiquinone 8 from cells grown under highly enriched 18O2 showed that three oxygen atoms of the quinone are derived from molecular oxygen. It was established that the molecular oxygen is incorporated into the two methoxyl groups (at C-5 and C-6) and one of the carbonyl positions of the ubiquinone molecule by demonstrating that only one of the incorporated oxygens will exchange with water under acidic conditions that specifically catalyze the exchange of carbonyl, but not methoxyl, oxygens. That the C-4 carbonyl oxygen is derived from molecular oxygen was shown by the incorporation of three atoms of 18O2 into ubiquinone 8 biosynthesized from added 4-hydroxybenzoic acid. Comparison of ubiquinone 8 and menaquinone 8 from E. coli grown under 18O2 confirmed that the labeled carbonyl oxygen of the [18O2]ubiquinone 8 is incorporated biosynthetically and not by chemical exchange in the cell. It is concluded that the three hydroxylation reactions involved in the pathway for the aerobic biosynthesis of ubiquinone are all catalyzed by monooxygenases. The implications of this study for the anaerobic biosynthesis of ubiquinone 8 in E coli are discussed.  相似文献   

6.
Proteolytic labeling in H2(18)O has been recently revived as a versatile method for proteomics research. To understand the molecular basis of the labeling process, we have dissected the process into two separate events: cleavage of the peptide amide bonds and exchange of the terminal carboxyl oxygens. It was demonstrated that both carboxyl oxygens can be catalytically labeled, independent of the cleavage step. Reaction kinetics of the tryptic 16O-to-18O exchange of YGGFMR, YGGFMK, and the tryptic digest of apomyoglobin were studied by matrix-assisted laser desorption/ionization Fourier transform ion cyclotron resonance mass spectrometry. A larger KM for the Lys-peptide (4400 +/- 700 microM), when compared to that of the Arg-peptide (KM 1300 +/- 300 microM), was mainly responsible for the slower reaction with YGGFMK (kcat/KM 0.64 +/- 0.14 microM(-1)min(-1)) compared to YGGFMR (kcat/KM 2.6 +/- 0.9 microM(-1)min(-1)). Multiplexed kinetic studies showed that endoprotease-catalyzed oxygen exchange is a general phenomenon, allowing homogeneous 18O2-coding of a variety of peptides. It was demonstrated for the first time that chymotrypsin 18O2-codes peptides during proteolysis. On the basis of the analyses reported here, we propose that proteolytic 18O labeling can be advantageously decoupled from protein digestion, and endoproteases can be used in a separate step to 18O2-code peptides for comparative studies after proteolysis has taken place.  相似文献   

7.
Membrane-bound phospholipase B was purified to a homogeneous state from Torulaspora delbrueckii cell homogenate. Cell homogenate was extracted with Triton X-100, and the enzyme was precipitated with acetone. The acetone powder was washed repeatedly with Tris-HCl buffer (pH 8.0) until no phospholipae B activity was detected in the soluble fraction. The enzyme was extracted with Triton X-100 from the final residue and purified about 1,390-fold by sequential chromatofocusing, Sepharose 6B, and DEAE-Sephadex A-50 column chromatography. The final preparation showed a single broad protein band on SDS-polyacrylamide gel electrophoresis when stained with silver stain reagent and PAS-reagent. The molecular weight of phospholipase B was about 390,000 and 140,000-190,000 as estimated by gel filtration on Sepharose 6B and SDS-polyacrylamide gel electrophoresis, respectively, suggesting that phospholipase B is an oligomeric protein. The isoelectric point was at pH 4.5. Phospholipase B has two pH optima, one acidic (pH 2.5-3.0) and the other alkaline (pH 7.2-8.0). At acidic pH the phospholipase B activity was greatly increased in the presence of divalent metal ions, although metal ions are not a factor for enzyme activity. On the other hand, at alkaline pH the enzyme required Ca2+ or Mn2+ for activity. The pH- and thermal-stabilities at both pHs were similar. The phospholipase B hydrolyzed all diacylphospholipids tested at acidic pH, but hydrolyzed only phosphatidylcholine at alkaline pH. The hydrolysis rates of lysophospholipids were much higher (about 10-fold) than those of diacylphospholipids at both pHs.  相似文献   

8.
Trypsin(ogen) was isolated from the pyloric ceca of Greenland cod. Greenland cod trypsin catalyzed hydrolysis of N alpha-benzoyl-DL-arginine p-nitroanilide, tosyl arginine methyl ester and protein and was inhibited by the serine protease inhibitor PMSF and other well-known trypsin inhibitors. Greenland cod trypsin was more stable at alkaline pH than at acid pH; and was inactivated by relatively low thermal treatment. Like other trypsins, the enzyme was rich in potential acidic amino acid residues but poor in basic amino acid residues and had a molecular weight of 23,500; but it had less potential disulfide pairs, less alpha-helix and a lower H phi ave than other trypsins previously characterized. Reactions catalyzed by Greenland cod trypsin were not very responsive to temperature change, such that specific activity was relatively high at low reaction temperature.  相似文献   

9.
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   

10.
A study is presented on proton transfer associated with the reaction of the fully reduced, purified bovine heart cytochrome c oxidase with molecular oxygen or ferricyanide. The proton consumption associated with aerobic oxidation of the four metal centers changed significantly with pH going from approximately 3.0 H(+)/COX at pH 6.2-6.3 to approximately 1.2 H(+)/COX at pH 8.0-8.5. Rereduction of the metal centers was associated with further proton uptake which increased with pH from approximately 1.0 H(+)/COX at pH 6.2-6.3 to approximately 2.8 H(+)/COX at pH 8.0-8.5. Anaerobic oxidation of the four metal centers by ferricyanide resulted in the net release of 1.3-1.6 H(+)/COX in the pH range 6.2-8.2, which were taken up by the enzyme on rereduction of the metal centers. The proton transfer elicited by ferricyanide represents the net result of deprotonation/protonation reactions linked to anaerobic oxidoreduction of the metal centers. Correction for the ferricyanide-induced pH changes of the proton uptake observed in the oxidation and rereduction phase of the reaction of the reduced oxidase with oxygen gave a measure of the proton consumption in the reduction of O(2) to 2H(2)O. The results show that the expected stoichiometric proton consumption of 4H(+) in the reduction of O(2) to 2H(2)O is differently associated, depending on the actual pH, with the oxidation and reduction phase of COX. Two H(+)/COX are initially taken up in the reduction of O(2) to two OH(-) groups bound to the binuclear Fe a(3)-Cu(B) center. At acidic pHs the third and fourth protons are also taken up in the oxidative phase with formation of 2H(2)O. At alkaline pHs the third and fourth protons are taken up with formation of 2H(2)O only upon rereduction of COX.  相似文献   

11.
Phosphate-water oxygen exchange catalyzed by Escherichia coli alkaline phosphatase was monitored using the 18O shift on the 31P NMR signal of inorganic phosphate. Different kinetic patterns were observed with native zinc enzyme and with its cobalt analogue. For native enzyme at pH values ranging from 4.4 to 10.0, the distribution of 18O species in Pi, viz. P18O4, P18O316O,P18O216O2,P18O16O3,P16O4, with time is compatible with a kinetic scheme in which E-P, the noncovalent enzyme-phosphate complex, dissociates more rapidly than it forms the covalent complex E-P. For the cobalt enzyme at pH 6.8, the distribution of 18O species in Pi with time is different and leads to the conclusion that formation of E-P is more rapid than dissociation of Pi from E-P-A computer simulation gave good quantitative agreement with the observed distribution for the time course of the cobalt enzyme reaction when the ratio of the rate of formation of E-P to dissection of E-P was assumed to be 3 +/- 0.5.  相似文献   

12.
立体选择性酰胺酶是一种重要的手性合成工具酶,在制备手性羧酸及其衍生物方面具有广阔的应用前景,日益受到重视。在酰胺酶的应用中,其立体选择性影响巨大。从底物、反应温度、pH、添加共溶剂和微生物来源5个方面综述了其对酰胺酶立体选择性的影响,对提高酰胺酶的立体选择性,扩大其在制备光学活性化合物领域的应用具有重要的意义。  相似文献   

13.
Labeling with (18)O is currently one of the most commonly used methods for incorporating a stable isotopic label into samples for comparative proteomic studies. In this approach, isotopic labeling involves the enzymatic digestion, typically performed with trypsin, of a protein population in (18)O-water, which incorporates the stable isotope into the C termini of the newly formed peptides. Although trypsin is often used to facilitate isotopic incorporation after digestion, it is typically overlooked that this same mechanism can lead to isotopic loss even under conditions such as low pH where it is assumed that trypsin is inactive. To examine the role that trypsin plays in isotopic loss, several experiments were performed on the rate of delabeling under conditions relevant to multidimensional proteomic experiments. Results from these studies demonstrate that enzyme-facilitated exchange of (18)O in the peptide with (16)O in the aqueous solvent was the major process by which the label is removed from the peptides, even under conditions of low pH and temperature where trypsin is thought to be inactive. This study brings the rapid, tryptic-facilitated exchange to the attention of laboratories using this scheme to prevent inaccuracies in quantitative labeling due to loss of the isotopic label.  相似文献   

14.
15.
Enzyme-mediated 18O/16O differential labeling of proteome samples often suffers from incomplete exchange of the carboxy-terminus oxygen atoms, resulting in ambiguity in the measurable abundance differences. In this study, an 18O/16O labeling strategy was optimized for and applied to the solution-based comparative analysis of the detergent-resistant membrane proteome (DRMP) of untreated and Iota-b (Ib)-induced Vero cells. Solubilization and tryptic digestion of the DRMP was conducted in a buffer containing 60% methanol. Unfortunately, the activity of trypsin is attenuated at this methanol concentration hampering the ability to obtain complete oxygen atom turnover. Therefore, the incorporation of the 18O atoms was decoupled from the protein digestion step by carrying out the trypsin-mediated heavy atom incorporation in a buffer containing 20% methanol; a concentration at which trypsin activity is enhanced compared to purely aqueous conditions. After isotopic labeling, the samples were combined, fractionated by strong cation exchange and analyzed by microcapillary reversed-phase liquid chromatography coupled on-line with electrospray ionization tandem mass spectrometry. In total, over 1400 unique peptides, corresponding to almost 600 proteins, were identified and quantitated, including all known caveolar and lipid raft marker proteins. The quantitative profiling of Ib-induced DRMP from Vero cells revealed several proteins with altered expression levels suggesting their possible role in Ib binding/uptake.  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.
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   

18.
Pancreatic carboxylester lipase catalyzes the exchange of 18O between water and 13,16-cis,cis-doco-sadienoic acid (DA) in monolayers at the argon-buffer interface (Muderhwa, J.M., Schmid, P.C., and Brockman, H.L. (1992) Biochemistry 31, 141). In mixed monolayers of 18O, 18O-DA and 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC), both the extent and mechanism of 18O exchange show characteristics of a critical transition in the range of 0.5-0.6 mol fraction of DA (Muderhwa, J.M., and Brockman, H. L. (1992) Biochemistry 31, 149). To determine if the regulatory behavior exhibited on this type of surface is limited to members of the carboxylester lipase gene family (cholinesterases), comparable experiments were performed with a genetically and functionally unrelated lipase, pancreatic colipase-dependent lipase (PL). PL readily catalyzed the exchange of 18O between water and the carboxyl group of DA with enzyme at either monolayer or catalytic levels in the fatty acid-buffer interface. The oxygen exchange reaction obeyed a random, sequential mechanism, indicating that the dissociation of the enzyme.DA complex is much faster than the rate-limiting step in the overall exchange process. Kinetic analysis of oxygen exchange in pure DA monolayers showed a first-order dependence on interfacial PL and DA concentrations from which kcat/Km values were calculated. The oxygen exchange reaction proceeded with a rate constant of 16 x 10(-2) cm2 pmol-1 s-1, a value comparable to that for hydrolysis of the ester substrate, 1,3-dioleoylglycerol. With a monolayer of PL adsorbed to the interfacial phase, kcat/Km for oxygen exchange was about 600-fold lower than the value obtained with catalytic levels of adsorbed enzyme, indicating a possible restriction of substrate diffusion in the protein-covered fatty acid monolayer. With constant bulk PL concentration and mixed lipid monolayers containing DA and the non-substrate lipid, POPC, the extent of oxygen exchange increased abruptly as the abundance of DA in the interface was increased from 0.5 to 0.6 mol fraction. Concomitant with this critical transition was a change in the apparent mechanism of oxygen exchange from coupled to random, sequential. For both the extent of oxygen exchange and its mechanism shift, the critical transition was independent of the lipid packing density, i.e. surface pressure, of the interface. These results show that PL responds similarly to carboxylester lipase with respect to changes in interfacial lipid mole fraction in DA-POPC surfaces.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
Gerratana B  Frey PA  Cleland WW 《Biochemistry》2001,40(9):2972-2977
The transition-state structure for the reaction catalyzed by kanamycin nucleotidyltransferase has been determined from kinetic isotope effects. The primary (18)O isotope effects at pH 5.7 (close to the optimum pH) and at pH 7.7 (away from the optimum pH) are respectively 1.016 +/- 0.003 and 1.014 +/- 0.002. Secondary (18)O isotope effects of 1.0033 +/- 0.0004 and 1.0024 +/- 0.0002 for both nonbridge oxygen atoms were measured respectively at pH 5.7 and 7.7. These isotope effects are consistent with a concerted reaction with a slightly associative transition-state structure.  相似文献   

20.
The nonenzymatic rates of deamidation of Asn residues in a series of pentapeptides with the sequences VSNXV and VXNSV, where X is one of 10 different amino acids, were determined at neutral, alkaline, and acid pH values. The results demonstrate that in neutral and alkaline solutions the amino acid residue on the amino side of the Asn had little or no effect on the rate of deamidation regardless of its charge or size. The group on the carboxyl side of Asn affected the rate of deamidation significantly. Increasing size and branching in the side chain of this residue decreased the rate of deamidation by as much as 70-fold compared to glycine in the N-G sequence, which had the greatest rate of deamidation. In acidic solution, the rate of deamidation of the Asn residue was not affected by the amino acid sequence of the peptide. The products for each deamidation reaction were tested for the formation of isoAsp residues. In neutral and alkaline solutions, all products showed that the isoAsp:Asp peptide products were formed in about a 3:1 ratio. In acidic solution, the Asp peptide was the only deamidation product formed. All peptides in which a Ser residue follows the Asn residue were found to undergo a peptide cleavage reaction in neutral and alkaline solutions, yielding a tripeptide and a dipeptide. The rate of the cleavage reaction was about 10% of the rate of the deamidation pathway at neutral and alkaline pH values. The rates of deamidation of Asn residues in the peptides studied were not affected by ionic strength, and were not specific base catalyzed. General base catalysis was observed for small bases like ammonia. A model for the deamidation reaction is proposed to account for the observed effects.  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