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 78 毫秒
1.
1, 2-Cyclohexanedione reacts specifically with the guanidino group of arginine or arginine residues at pH 8 to 9 in sodium borate buffer in the temperature range of 25-40 degrees. The single product, N-7, N-8-(1,2-dihydroxycyclohex-1,2-ylene)-L-arginine (DHCH-arginine) is stable in acidic solutions and in borate buffers (pH 8 to 9). DHCH-Arginine is converted to N-7-adipyl-L-arginine by periodate oxidation. The structures of the two compounds were elucidated by chemical and physicochemical means. Arginine or arginyl residues can be regenerated quantitatively from DHCH-arginine by incubation at 37 degrees in hydroxylamine buffer at pH 7.0 FOR 7 TO 8 hours. Analysis of native egg white lysozyme and native as well as oxidized bovine pancreatic RNase, which were treated with cyclohexanedione, showed that only arginine residues were modified. The utility of the method in sequence studies was shown on oxidized bovine pancreatic ribonuclease A. Arginine modification was complete in 2 hours at 35 degrees in borate buffer at pH 9.0 with a 15-fold molar excess of the reagent. The derived peptides showed that tryptic hydrolysis was entirely limited to peptide bonds involving lysine residues, as shown both by two-dimensional peptide patterns and by isolation of the resulting peptides. The stability of DHCH-arginyl residues permits isolation of labeled peptides.  相似文献   

2.
Activation of porcine pepsinogen at pH 2.0 was found to proceed simultaneously by two different pathways. One pathway is the direct conversion process of pepsinogen to pepsin, releasing the intact activation segment. The isolation of the released 44-residue segment was direct evidence of this one-step process. At pH 5.5 the segment bound tightly to pepsin to form a 1:1 pepsin-activation segment complex, which was chromatographically indistinguishable from pepsinogen. The other is a stepwise-activating or sequential pathway, in which pepsinogen is activated to pepsin through intermediate forms, releasing activation peptides stepwisely. These intermediate forms were isolated and characterized. The major intermediate form was shown to be generated by removal of the amino-terminal 16 residues from pepsinogen. The released peptide mixture was composed of two major peptides comprising residues 1-16 and 17-44, and hence the stepwise-activating process was deduced to be mainly a two-step process.  相似文献   

3.
It was found that at pH 5.2 and 40-fold excess of p-nitrophenyldiazonium chloride the inhibitor incorporation into the porcine pepsin molecule involves 1.9 residues, one residue being bound to tyrosine 189. Besides, tyrosines 44, 113, 154 and 174 enter the reaction. Modified pepsin retains 25% of the native enzyme activity. In the pepsinogen molecule the degree of tyrosine 189 modification diminishes 5 times; of 1.5 inhibitor molecules incorporated into the protein 0.78 residues are bound to tyrosine 113. The potential proteolytic activity of modified pepsinogen towards haemoglobin cleavage makes up to 60% of the original one. It is concluded that the activation peptide in the pepsinogen molecule masks the substrate binding site bearing tyrosine 189, thus preventing its modification with p-nitrophenyldiazonium chloride. The activation peptide in the pepsinogen molecule is presumably located in the vicinity of the wide loop bend carrying tyrosine residue 113, which may be the reason for the decreased pKa value of this residue and of its increased reactivity in the azocoupling reaction.  相似文献   

4.
Intramolecular pepsinogen activation is inhibited either by pepstatin, a potent pepsin inhibitor, or by purified globin from hemoglobin, a good pepsin substrate. Also, pepsinogen at pH 2 can be bound to a pepstatin-Sepharose column and recovered as native zymogen upon elution in pH 8 buffer. Kinetic studies of the globin inhibition of pepsinogen activation show that globin binds to a pepsinogen intermediate. This interaction gives rise to competitive inhibition of intramolecular pepsinogen activation. The evidence presented in this paper suggests that pepsinogen is converted rapidly upon acidification to the pepsinogen intermediate delta. In the absence of an inhibitor, the intermediate undergoes conformational change to bind the activation peptide portion of this same pepsinogen molecule in the active center to form an intramolecular enzyme-substrate complex (intermediate theta). This is followed by the intramolecular hydrolysis of the peptide bond between residues 44 and 45 of the pepsinogen molecule and the dissociation of the activation peptide from the pepsin. Intermediate delta apparently does not activate another pepsinogen molecule via an intermolecular process. Neither does intermediate delta hydrolyze globin substrate.  相似文献   

5.
Influence of pH on the appearance of active peptides in peptic hydrolysis of bovine haemoglobin was studied in a homogenous phase system. Six active peptides were studied: three hemorphins: LVVH-7 (beta 31-40), VVH-7 (beta 32-40), VVH-4 (beta 32-37), one bradykinin-potentiating peptide (alpha 110-125), one antibacterial peptide (alpha 1-23), and neokyotorphin (alpha 137-141). The influence of pH was investigated in the course of the hydrolysis of haemoglobin by pepsin at 23 degrees C in acetate buffer at pH 3.5, pH 4.5, and pH 5.5. The hydrolysis of haemoglobin was studied in the presence or absence of urea. The haemoglobin hydrolysis at pH 4.5 is taken as a reference. Two different mechanisms of hydrolysis were observed: "one by one" for native haemoglobin hydrolysis at pH 4.5 and 5.5, and "zipper" for denatured haemoglobin at pH 3.5, pH 4.5, and pH 5.5, and native haemoglobin at pH 3.5. Whatever the pH and medium, a selectivity change by the pepsin was noticed. In the presence of urea, there are two phenomena: some peptides are preferentially produced at pH 3.5 and other peptides at pH 5.5, which seems to favour one particular site of pepsin that is cut. In the absence of urea, these active peptides reached a higher concentration at pH 3.5. In order to prepare these six active peptides, it is suitable to hydrolyse haemoglobin in the absence of urea at pH 3.5 (this pH denatures haemoglobin) where a "zipper" mechanism is obtained, and the peptide quantity is more significant at pH 3.5 than at pH 4.5.  相似文献   

6.
Spectrophotometric titration curves were obtained at 242 nm for native and fully guanidinated horse-heart ferricytochrome c. The cytochrome c data were fit over the pH range 9-12 (I = 0.35) by a theoretical curve with pK' values of 10.35 and 11.70. The slope of the experimental data increases sharply above pH 12.5 suggesting that two tyrosine residues with pK' values greater than 12.5 are exposed by conformation change. The guanidinated cytochrome c data after correction for the alkaline spin-state transition were fit over the entire pH range 9-13.6 (I = 0.35) by a theoretical curve with pK' values 10.37, 10.78, 11.50, and 13.60. These results along with viscosity measurements indicate that the unfolding transition occurs at higher pH in the guanidinated derivative. N-Acetylimidazole was used to acetylate specific tyrosyl groups of guanidinated cytochrome c. Assignments of acetylated tyrosine residues were confirmed by peptide mapping of 14C-labelled derivatives. Spectrophotometric titrations with rapid data acquisition of two monoacetylated derivatives allowed assignments of pK'1 (10.37) to Tyr-67 and pK'4 (13.60) to Tyr-97. The basis for the large differences in acidity and chemical reactivity of these two residues is not obvious from the crystallographic structure and may arise from differences in solvent access due to motions of the polypeptide chain.  相似文献   

7.
The membrane-binding characteristics of a number of modified vitamin K-dependent proteins and peptides showed a general pattern of structural requirements. The amino-terminal peptides from human prothrombin (residues 1-41 and 1-44, 60:40) bovine factor X (residues 1-44), and bovine factor IX (residues 1-42), showed a general requirement for a free amino-terminal group, an intact disulfide, and the tyrosine homologous to Tyr44 of factor X for membrane binding. Consequently, the peptide from factor IX did not bind to membranes. Any of several modifications of the amino terminus, except reaction with trinitrobenzenesulfonic acid, abolished membrane binding by the factor X and prothrombin peptides. Calcium, but not magnesium, protected the amino terminus from chemical modification. The requirement for a free amino terminus was also shown to be true for intact prothrombin fragment 1, factor X, and factor IX. Although aggregation of the peptide-vesicle complexes greatly complicated accurate estimation of equilibrium binding constants, results with the factor X peptide indicated an affinity that was not greatly different from that of the parent protein. The most striking difference shown by the peptides was a requirement for about 10 times as much calcium as the parent proteins. In a manner similar to the parent proteins, the prothrombin and factor X peptides showed a large calcium-dependent quenching of tryptophan fluorescence. This fluorescence quenching in the peptides also required about 10 times the calcium needed by the parent proteins. Thus, the 1-45 region of the vitamin K-dependent proteins contained most of the membrane-binding structure but lacked component(s) needed for high affinity calcium binding. Protein S that was modified by thrombin cleavage at Arg52 and Arg70 showed approximately the same behavior as the amino-terminal 45-residue peptides. That is, it bound to membranes with overall affinity that was similar to native protein S but required high calcium concentrations. These results suggested that the second disulfide loop of protein S (Cys47-Cys72) and prothrombin (Cys48-Cys61) were involved in high affinity calcium binding. Since factor X lacks a homologous disulfide loop, an alternative structure must serve a similar function. A striking property of protein S was dissociation from membranes by high calcium. While this property was shared by all the vitamin K-dependent proteins, protein S showed this most dramatically and supported protein-membrane binding by calcium bridging.  相似文献   

8.
Selective oxidation of methionine residues in proteins.   总被引:7,自引:0,他引:7  
Methionine residues in peptides and proteins were oxidized to methionine sulfoxides by mild oxidizing reagents such as chloramine-T and N-chlorosuccinimide at neutral and slightly alkaline pH. With chloramine-T cysteine was also oxidized to cystine but no other amino acid was modified; with N-chlorosuccinimide tryptophans were oxidized as well. In peptides and denaturated proteins all methionine residues were quantitatively oxidized, while in native proteins only exposed methionine residues could be modified. Extent of oxidation of methionine residues was determined by quantitative modification of the unoxidized methionine residues with cyanogen bromide (while methionine sulfoxide residues remained intact), followed by acid hydrolysis and amino acid analysis. Methionine was determined as homoserine and methionine sulfoxide was reduced back to methionine. Sites of oxidation were identified in a similar way by cleaving the unoxidized methionyl peptide bonds with cyanogen bromide, followed by quantitative end-group analysis of the new amino-terminal amino acids (by an automatic sequencer).  相似文献   

9.
C-type lysozyme (cLZ) is an antimicrobial enzyme that plays a major defense role in many human secretions. Recently, we have identified a helix-loop-helix antimicrobial peptide fragment of cLZ. This finding suggests that processing by coexisting proteases might be a relevant physiological process for generating peptides that contribute to the in vivo mucosal defense role of cLZ. In this study, we found that pepsin, under condition relevant to the newborn stomach (pH 4.0), generated various peptides from cLZ with potent bactericidal activity against several strains of Gram-negative and Gram-positive bacteria. Microsequencing and mass spectral analysis revealed that pepsin cleavage occurred at conserved loops within the alpha-domain of cLZ. We found that the bactericidal domain, which was isolated by gel filtration and reversed-phase HPLC, contains two cationic alpha-helical peptides generated from a helix-loop-helix domain (residues 1-38 of cLZ) by nicking at leucine17. A third peptide consisting of an alpha-helix (residues 18-38) and a two-stranded beta-sheet (residues 39-56) structure was also identified. These peptides share structural motifs commonly found in different innate immune defenses. Functional cellular studies with outer membrane-, cytoplasmic membrane vitality- and redox-specific fluorescence dyes revealed that the lethal effect of the isolated antimicrobial peptides is due to membrane permeabilization and inhibition of redox-driven bacterial respiration. The results provide the first demonstration that pepsin can fine-tune the antimicrobial potency of cLZ by generating multiple antimicrobial peptide motifs, delineating a new molecular switch of cLZ in the mucosal defense systems. Finally, this finding offers a new strategy for the design of antibiotic peptide drugs with potential use in the treatment of infectious diseases.  相似文献   

10.
Three peptides modelling a highly potent, 35-residue chymotrypsin inhibitor (Schistocerca gregaria chymotrypsin inhibitor) were designed and synthesized by convergent peptide synthesis. For each model peptide, the inhibitory constant (Ki) on chymotrypsin and the solution structure were determined. In addition, molecular dynamics calculations were performed for all of them. Two models containing approximately half of the parent inhibitor (17 of 35 residues) were designed and subsequently found to have no substantial inhibitory activity (Ki values in the mM range). The third model composed of 24 amino acid residues proved to be an effective (Ki approximately 10(-7)) inhibitor of bovine chymotrypsin. Both the solution structure properties determined by NMR spectroscopy and the dynamic behaviour of the latter model system are comparable to the native inhibitor. In contrast, the structure and dynamics of the first two related model peptides show characteristic differences. We suggest that the conformation and flexibility of the modelled protease inhibitor are crucial for its biological efficiency. Moreover, the structural and dynamic features of the binding loop (28-33) and those of the rest of the molecule appear to be interdependent. Most importantly, these structural characteristics can be rationally modified, at least partially, by peptide design.  相似文献   

11.
A neutral protease from Bacillus subtilis var. amylosacchariticus was modified with tetranitromethane (TNM) at pH 8.0 for 1 h at 25 degrees C, by which treatment the proteolytic activity toward casein was markedly reduced, whereas activity changes toward N-blocked peptide substrates were variable depending upon the substrate used. The modified enzyme was digested with a Staphylococcus aureus V8 protease at pH 7.9 and the resultant peptides were separated by HPLC. Two peptides which contain nitrotyrosyl residue(s) were purified. One of the peptides was found to have an amino acid sequence of Thr-Ala-Asn-Leu-Ile-Tyr-Glu, which corresponds to residue Nos. 153-159 of the neutral protease, and Tyr-158 was identified as PTH-nitrotyrosine. The other one was the amino-terminal peptide of residue Nos. 1-22, and Tyr-21 was shown to be nitrated. From a comparison with the active site structure of thermolysin, which is a zinc metalloprotease with a high sequence homology to B. subtilis neutral proteases, nitration of Tyr-158 was inferred to be closely related to the activity changes of the neutral protease from B. subtilis var. amylosacchariticus.  相似文献   

12.
Plasmodium falciparum malaria protein peptides were synthesised in the search for more effective routes for inducing a protective immune response against this deadly parasite and this information has been associated with such molecules' three-dimensional structure. These peptides had high red blood cell binding activity and their carboxy- and amino-terminal extremes were elongated for determining their immunogenic and protection-inducing activity against this disease in the Aotus monkey experimental model. 1H-NMR was used for analysing their three-dimensional structure; FAST ELISA, immunofluorescence antibody test, and Western blot were used for identifying their antibody inducing capacity and these previously immunised Aotus were inoculated with a highly infective P. falciparum strain to determine whether these elongated peptides were able to induce protection. This was aimed at establishing an association or correlation between long peptides' three-dimensional structure and their immunogenic and protection-inducing response in these monkeys. Peptides 20026 (25 residue), 20028 (30 residue), and 20030 (35 residues) were synthesised based on elongating the amino-terminal region of the 10022 highly immunogenic and protection-inducing modified peptide. 1H-NMR studies revealed that the first three had Classical type III beta-turn structures, different from the 20-amino acid long modified peptide 10022 which had a distorted type III beta-turn. Humoral immune response analysis showed that even when some antibodies could be generated against the parasite, none of the immunised Aotus could be protected with elongated peptides suggesting that elongating them eliminated modified peptide 10022 immunogenic and protection-inducing capacity.  相似文献   

13.
A murine monoclonal antibody (designated H-11) produced by injecting mice with purified human protein C was found to bind several human vitamin K-dependent proteins. Using a solid-phase competitive radioimmunoassay with antibody immobilized onto microtiter plates, binding of 125I-labeled protein C to the antibody was inhibited by increasing amounts of protein C, prothrombin, and Factors X and VII over a concentration range of 1 X 10(-8) to 1 X 10(-6) M. Other vitamin K-dependent proteins including Factor IX and protein S did not inhibit or inhibited only at the highest concentration binding of radiolabeled protein C to the immobilized antibody. Chemical treatment of prothrombin with a variety of agents including denaturation by sodium dodecyl sulfate, reduction with mercaptoethanol followed by carboxymethylation with iodoacetic acid, citraconylation of lysine residues, removal of metal ion with EDTA, or heat decarboxylation did not destroy the antigenic site recognized by the antibody as measured by immunoblotting of prothrombin or prothrombin derivative immobilized onto nitrocellulose. Immunoblotting of purified vitamin K-dependent polypeptides with the monoclonal antibody following sodium dodecyl sulfate-polyacrylamide gel electrophoresis and electrophoretic transfer to nitrocellulose indicated that the antigenic site was found on the light chains of protein C and Factor X. Chymotrypsin digestion of prothrombin and isolation on QAE-Sephadex of the peptide representing amino-terminal residues 1-44 of prothrombin further localized the antigenic site recognized by the monoclonal antibody to the highly conserved gamma-carboxyglutamic acid-containing domain. The exact location of the antigenic determinant for antibody H-11 was established using synthetic peptides. Antibody H-11 bound specifically to synthetic peptides corresponding to residues 1-12 of Factor VII and 1-22 of protein C. Comparison of protein sequences of bovine and human vitamin K-dependent proteins suggests that the sequence Phe-Leu-Glu-Glu-Xaa-Arg/Lys is required for antibody binding. The glutamic acid residues in this peptide segment are the first 2 gamma-carboxyglutamic acid residues near the amino-terminal end in the native proteins. Increasing concentrations of Ca2+, Mg2+, or Mn2+ partially inhibited binding of 125I-protein C to the antibody in a solid-phase assay system with half-maximal binding observed at divalent metal ion concentrations of 2, 4, and 0.6 mM, respectively.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

14.
Trypsin [EC 3.4.21.4] modified (reactive site cleaved) Vicia angustifolia proteinase inhibitor was prepared at pH 3 with a catalytic amount of trypsin and purified using columns of Sephadex G-50 and DEAE-Sephadex A-25. The modified inhibitor, which still retained antitryptic activity, lost its activity upon treatment with carboxypeptidase B or citraconic anhydride. End-group analyses revealed that the carboxyl-terminal Arg and the amino-terminal Ser residues were newly exposed end-groups in the modified inhibitor. It takes a much longer incubation time (about 1 h) to exhibit the maximal inhibitory activity against trypsin. Reduction and carboxymethylation of the modified inhibitor produced two fragments on Sephadex G-50 chromatography. The smaller fragment consisted of about 32 amino acid residues and possessed a new carboxyl-terminal Arg residue. The larger fragment consisted of about 80 residues and possessed a Ser residue at its amino-terminus. These results indicate that the small fragment was derived from the amino-terminal portion of the modified inhibitor and the large fragment from the carboxyl-terminal. It is also concluded that an Arg-Ser bond is the reactive site as well as the inhibitory site of the V. angustifolia inhibitor against trypsin. The sequence around the antitryptic site exhibits high degrees of homology with other double-headed inhibitors of legume origin, such as the Bowman-Birk inhibitor, lima beam inhibitor, and the major inhibitor in chick-peas.  相似文献   

15.
Y Paterson 《Biochemistry》1985,24(4):1048-1055
Two regions of rodent cytochrome c, one within the first four residues of the molecule, which is N-acetylated, and one at a beta bend around residue 44, are known to be immunogenic and antigenic in rabbits. Using sequential peptide synthesis, we have determined the residues required for linear synthetic peptides within these sequences to bind to antibody raised in rabbits to intact rat cytochrome c. The residues that were important in binding the N-terminal peptides were N-acetylglycine at position 1 and valine at position 3. The smallest peptide sequence around residue 44 that would bind to antibodies was Gln-Ala-Ala-Gly-Phe. A theoretical conformational analysis of these peptides showed that the amino-terminal tetrapeptide adopts a wide statistical ensemble of conformational states and that the addition of residues beyond 41 and 45 in the other sequence does not appear to stabilize longer peptides in the native beta-bend conformation. Thus, the antigenicity conferred by Phe-46 and Gln-42 in this peptide is most likely due to the direct interaction of the side chains of these residues with the antibody binding site. The demonstration here that native conformation is not essential for antigenic peptides to bind to antibodies raised against the whole protein indicates that the association energy between antigen and antibody can be sufficient to induce conformation in conformationally flexible peptides. This supports the concept that anti-protein and anti-peptide antibodies may invoke conformational changes in cross-reactive protein antigens and may explain why longer peptides, which may adopt stable nonnative secondary structure, often do not bind to antibodies raised to the whole molecule.  相似文献   

16.
We have evaluated the effect of lysine guanidination in peptides and proteins on the dissociation of protonated ions in the gas phase. The dissociation of guanidinated model peptide ions compared to their unmodified forms showed behavior consistent with concepts of proton mobility as a major factor in determining favored fragmentation channels. Reduction of proton mobility associated with lysine guanidination was reflected by a relative increase in cleavages occurring C-terminal to aspartic acid residues as well as increases in small molecule losses. To evaluate the effect of guanidination on the dissociation behavior of whole protein ions, bovine ubiquitin was selected as a model. Essentially, all of the amide bond cleavages associated with the +10 charge state of fully guanidinated ubiquitin were observed to occur C-terminal to aspartic acid residues, unlike the dissociation behavior of the +10 ion of the unmodified protein, where competing cleavage N-terminal to proline and nonspecific amide bond cleavages were also observed. The +8 and lower charge states of the guanidinated protein showed prominent losses of small neutral molecules. This overall fragmentation behavior is consistent with current hypotheses regarding whole protein dissociation that consider proton mobility and intramolecular charge solvation as important factors in determining favored dissociation channels, and are also consistent with the fragmentation behaviors observed for the guanidinated model peptide ions. Further evaluation of the utility of condensed phase guanidination of whole proteins is necessary but the results described here confirm that guanidination can be an effective strategy for enhancing C-terminal aspartic acid cleavages. Gas phase dissociation exclusively at aspartic acid residues, especially for whole protein ions, could be useful in identifying and characterizing proteins via tandem mass spectrometry of whole protein ions.  相似文献   

17.
We studied fusion induced by a 20-amino acid peptide derived from the amino-terminal segment of hemagglutinin of influenza virus A/PR/8/34 [Murata, M., Sugahara, Y., Takahashi, S., & Ohnishi, S. (1987) J. Biochem. (Tokyo) 102, 957-962]. To extend the study, we have prepared several water-soluble amphiphilic peptides derived from the HA peptide; the anionic peptides D4, E5, and E5L contain four and five acidic residues and the cationic peptide K5 has five Lys residues in place of the five Glu residues in E5. Fusion of egg phosphatidylcholine large unilamellar vesicles induced by these peptides is assayed by two different fluorescence methods, lipid mixing and internal content mixing. Fusion is rapid in the initial stage (12-15% within 20 s) and remains nearly the same or slightly increasing afterward. The anionic peptides cause fusion at acidic pH lower than 6.0-6.5, and the cationic peptide causes fusion at alkaline pH higher than 9.0. Leakage and vesiculation of vesicles are also measured. These peptides are bound and associated with vesicles as shown by Ficoll discontinuous gradients and by the blue shift of tryptophan fluorescence. They take an alpha-helical structure in the presence of vesicles. They become more hydrophobic in the pH regions for fusion. When the suspension is made acidic or alkaline, the vesicles aggregate, as shown by the increase in light scattering. The fusion mechanism suggests that the amphiphilic peptides become more hydrophobic by neutralization due to protonation of the carboxyl groups or deprotonation of the lysyl amino groups, aggregate the vesicles together, and interact strongly with lipid bilayers to cause fusion. At higher peptide concentrations, E5 and E5L cause fusion transiently at acidic pH followed by vesiculation.  相似文献   

18.
The primary structure of Aspergillus niger acid proteinase A   总被引:1,自引:0,他引:1  
The complete amino acid sequence of the acid proteinase A, a non-pepsin type acid proteinase from the fungus Aspergillus niger var. macrosporus, was determined by protein sequencing. The enzyme was first dissociated at pH 8.5 into a light (L) chain and a heavy (H) chain, and the L chain was sequenced completely. Further sequencing was performed with the reduced and pyridylethylated or aminoethylated derivative of the whole protein, using peptides obtained by digestions with Staphylococcus aureus V8 protease, trypsin, chymotrypsin, and lysylendopeptidase. The location of the two disulfide bonds was determined by analysis of cystine-containing peptides obtained from a chymotryptic digest of the unmodified protein. These results established that the protein consists of a 39-residue L chain and a 173-residue H chain that associate noncovalently to form the native enzyme of 212 residues (Mr 22,265). This is, to our knowledge, the first time that such a protein with a rather short peptide chain associated noncovalently has been found. No sequence homology is found with other acid or aspartic proteinases, except for Scytalidium lignicolum acid proteinase B, an enzyme unrelated to pepsin by sequence, which has about 50% identity with the present enzyme. These two enzymes, however, are remarkably different from each other in some structural features.  相似文献   

19.
Rabbit muscle aldolase (E.C. 4. 1. 2. 13) was guanidinated by reaction with O-methylisourea. Up to 60% of the lysine residues can be guanidinated without any dissociation of the tetramer but with a complete loss of enzymatic activity. Native and guanidinated aldolase can be dissociated into monomers in 2.4 m MgCl2 with only slight change in conformation of the subunit. Nitrotroponylation of guanidinated aldolase in dilute buffer gives no reaction whereas in 2.4 m MgCl2 nitrotroponlylation modifies another 8–12% of the lysine residues. Removal of MgCl2 by dialysis affords 100% recovery of activity and tetrameric structure for native aldolase and 100% recovery of tetrameric structure for guanidinated aldolase. In contrast nitrotroponylated and guanidinated aldolase remains monomeric before precipitating as the MgCl2 concentration is lowered. It is concluded that lysine may be involved in the protein-protein interaction of the subunit contact domains of muscle aldolase.  相似文献   

20.
Rabbit liver cytosolic serine hydroxymethyltransferase exists in several subforms which have different isoelectric points. Incubation of the purified enzyme with chymotrypsin cleaves the enzyme at Trp14. The released amino-terminal 14-mer peptide was shown to exist in three forms of equal concentration. The peptides differ in structure only at the asparaginyl residue at position 5. In addition to asparagine at this position we found both aspartyl and isoaspartyl residues. The deamidation of Asn5 does not appear to occur during the purification of the enzyme. The in vitro rate of deamidation of Asn5 in the enzyme is more than 5-fold slower than the rate of deamidation of this residue in the free 14-mer peptide. The isoaspartyl residue at position 5 serves as a substrate for protein carboxyl methyltransferase both in the free 14-mer peptide and the native enzyme. The enzyme which has had the amino-terminal 14 residues removed by digestion with chymotrypsin still exists in several forms with different isoelectric points. Reaction of peptides from this enzyme with carboxyl methyltransferase suggests that there is at least one more asparaginyl residue in this enzyme other than Asn5 which has undergone deamidation with the formation of isoaspartyl bonds.  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