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.
探索生物转化法制备L-天冬酰胺的技术与工艺。通过分子生物学方法,克隆来源于大肠杆菌(Escherichia coli, E.coli)JM109的天冬酰胺合成酶A基因asnA,并于E. coli BL21(DE3)中表达,利用构建的E.coli基因工程菌E.coli BL21(DE3)/pET28a(+)-asnA全细胞高密度催化L-天冬氨酸生产L-天冬酰胺,以PITC柱前衍生-高效液相检测底物和产物。表达的蛋白质分子质量约为37kDa,与预期大小相符,比酶活力为1786.6U/g。L-天冬氨酸转化率为95.8%,L-天冬酰胺产量可达126.5g/L,生产速率为15.81g/(L·h)。结果表明,已成功构建高效表达天冬酰胺合成酶A基因工程菌株,并用于催化L-天冬氨酸转化生产L-天冬酰胺,解决了L-天冬酰胺生物转化生产工艺中ATP成本过高的难题,为L-天冬酰胺制备提供新的绿色途径。  相似文献   

2.
Abstract

Experiments using equilibrium dialysis and fluorescence quenching provided direct evidence that approximately four moles of L-aspartic acid were bound per mole of tetrameric L-asparaginase from Escherichia coli, with a dissociation constant on the order of 60-160 μM. In addition, a set of weaker binding sites with a dissociation constant in the millimolar range were detected. Kinetic studies also revealed that L-aspartic acid inhibited L-asparaginase competitively, with an inhibition constant of 80 μM at micromolar concentrations of L-asparagine; at millimolar concentrations of the amide, an increase in maximal velocity but a decrease in affinity for L-asparagine were observed. L-Aspartic acid at millimolar levels again displayed competitive inhibition. These and other observations suggest that L-aspartic acid binds not only to the active site but also a second site with lower intrinsic affinity for it. The observed “substrate activation” is most likely attributable to the binding of a second molecule of L-asparagine rather than negative cooperativity among the tight sites of the subunits of this tetrameric enzyme. Further support for L-aspartic acid binding to the active site comes from experiments in which the enzyme, when exposed to various group-specific reagents suffered parallel loss of catalytic activity and in its ability to bind L-aspartic acid. Different commercial preparations of Escherichia coli L-asparaginase were found to contain ~ 2-4 moles of L-aspartic acid; these were incompletely removed by dialysis, but could be removed by transamination or decarboxylation. Efficiency of dialysis increased with increasing pH. Taken together, this set of results is consistent with the existence of a covalent β-aspartyl enzyme intermediate.  相似文献   

3.
Experiments using equilibrium dialysis and fluorescence quenching provided direct evidence that approximately four moles of L-aspartic acid were bound per mole of tetrameric L-asparaginase from Escherichia coli, with a dissociation constant on the order of 60-160 microM. In addition, a set of weaker binding sites with a dissociation constant in the millimolar range were detected. Kinetic studies also revealed that L-aspartic acid inhibited L-asparaginase competitively, with an inhibition constant of 80 microM at micromolar concentrations of L-asparagine; at millimolar concentrations of the amide, an increase in maximal velocity but a decrease in affinity for L-asparagine were observed. L-Aspartic acid at millimolar levels again displayed competitive inhibition. These and other observations suggest that L-aspartic acid binds not only to the active site but also a second site with lower intrinsic affinity for it. The observed "substrate activation" is most likely attributable to the binding of a second molecule of L-asparagine rather than negative cooperativity among the tight sites of the subunits of this tetrameric enzyme. Further support for L-aspartic acid binding to the active site comes from experiments in which the enzyme, when exposed to various group-specific reagents suffered parallel loss of catalytic activity and in its ability to bind L-aspartic acid. Different commercial preparations of Escherichia coli L-asparaginase were found to contain approximately 2-4 moles of L-aspartic acid; these were incompletely removed by dialysis, but could be removed by transamination or decarboxylation. Efficiency of dialysis increased with increasing pH. Taken together, this set of results is consistent with the existence of a covalent beta-aspartyl enzyme intermediate.  相似文献   

4.
Human glycoasparaginase (N4-(beta-N-acetyl-D-glucosaminyl)-L-asparaginase, EC 3.5.1.26) hydrolyzes a series of compounds that contain L-asparagine residue with free alpha-amino and alpha-carboxyl groups. Substrates include high mannose and complex type glycoasparagines as well as those that lack the di-N-acetylchitobiose moiety, L-aspartic acid beta-methyl ester and L-aspartic acid beta-hydroxamate. The enzyme is inactive toward L-asparagine and L-glutamine and glycoasparagines containing substituted alpha-amino and/or alpha-carboxyl groups. In the presence of the acyl acceptor hydroxylamine, glycoasparaginase catalyzes the synthesis of L-aspartic acid beta-hydroxamate from aspartyl-glucosamine, L-aspartic acid beta-methyl ester, and L-aspartic acid. 13C NMR studies using 18O-labeled L-aspartic acid demonstrate that glycoasparaginase catalyzes an oxygen exchange between water and the carboxyl group at C-4 of L-aspartic acid. These results indicate that glycoasparaginase reacts as an exo-hydrolase toward the L-asparagine moiety of the substrates and the free alpha-amino and alpha-carboxyl groups are required for the enzyme reaction. The results are consistent with an L-asparaginase-like reaction pathway which involves a beta-aspartyl enzyme intermediate. Since glycoasparaginase is active toward a series of structurally different glycoasparagines, we suggest the revised systematic name of N4-(beta-glycosyl)-L-asparaginase for the enzyme.  相似文献   

5.
Mucochloric and mucobromic acids are powerful inhibitors of tumoral and pancreatic L-asparagine synthetases. Two nitrogen donors, L-glutamine and ammonia, can be used by these enzymes; at a concentration of 1 mmol/l, mucochloric and mucobromic acids preferentially inhibit the utilization of ammonia as opposed to L-glutamine in vitro. Using the tumoral enzyme, kinetic analysis revealed that mucochloric acid produced inhibition which was apparently noncompetitive with ammonia but competitive with L-glutamine. In molar excess, L-glutamine and dithiothreitol effectively antagonized such inhibition; dialysis, however, failed to reverse established inhibition. These findings, suggest that the drugs operate by covalent attachment to crucial sulfhydryl functions on the enzyme.  相似文献   

6.
The thermophilic cyanobacterium Synechococcus sp. strain MA19 contained the structural genes for cyanophycin synthetase (cphA) and cyanophycinase (cphB), which were identified, cloned, and sequenced in this study. The translation products of cphA and cphB exhibited high levels of similarity to corresponding proteins of other cyanobacteria, such as Anabaena variabilis and Synechocystis sp. Recombinant cells of Escherichia coli harboring cphA colinear with lacPO accumulated cyanophycin that accounted for up to 25% (wt/wt) of the dry cell matter in the presence of isopropyl-beta-D-thiogalactopyranoside (IPTG). The cyanophycin synthetase was enriched 123-fold to electrophoretic homogeneity from the soluble fraction of the recombinant cells by anion-exchange chromatography, affinity chromatography, and gel filtration chromatography. The purified cyanophycin synthetase maintained the parental thermophilic character and was active even after prolonged incubation at 50 degrees C; in the presence of ectoine the enzyme retained 90% of its activity even after 2 h of incubation. The in vitro activity of the enzyme depended on ATP, primers, and both substrates, L-arginine and L-aspartic acid. In addition to native cyanophycin, the purified enzyme accepted a modified cyanophycin containing less arginine, alpha-arginyl aspartic acid dipeptide, and poly-alpha,beta-DL-aspartic acid as primers and also incorporated beta-hydroxyaspartic acid instead of L-aspartic acid or L-canavanine instead of L-arginine at a significant rate. The lack of specificity of this thermostable enzyme with respect to primers and substrates, the thermal stability of the enzyme, and the finding that the enzyme is suitable for in vitro production of cyanophycin make it an interesting candidate for biotechnological processes.  相似文献   

7.
In continuing studies of patulin biosynthesis, the first enzyme of the pathway, 6-methylsalicylic acid synthetase, was found to be far more labile than were the later enzymes of the pathway. Attempts were made to stabilize 6-methylsalicylic acid synthetase in vitro. The combined addition of the cofactor NADPH, the substrates acetyl-CoA and malonyl-CoA, the reducing agent dithiothreitol, and the proteinase inhibitor phenylmethylsulfonyl fluoride to cell-free extracts was found to prolong the half-life of the enzyme as much as 12-fold. This suggested that proteolysis and the conformational integrity of the enzyme may play an important role in controlling the duration of antibiotic biosynthesis in vivo. This was in agreement with the finding that the intracellular proteinase content of antibiotic-producing cells of Penicillium urticae rapidly increased just before the loss of 6-methylsalicylic acid synthetase content. These in vitro stabilization studies have provided some insight into the metabolic conditions that may stabilize these enzymes in vivo, and into possible ways of extending the life of these catalysts.  相似文献   

8.
The mode of action of the antibiotic pseudomonic acid has been studied in Escherichia coli. Pseudomonic acid strongly inhibits protein and RNA synthesis in vivo. The antibiotic had no effect on highly purified DNA-dependent RNA polymerase and showed only a weak inhibitory effect on a poly(U)-directed polyphenylalanine-forming ribosomal preparation. Chloramphenicol reversed inhibition of RNA synthesis in vivo. Pseudomonic acid had little effect on RNA synthesis in a regulatory mutant, E. coli B AS19 RC(rel), whereas protein synthesis was strongly inhibited. In pseudomonic acid-treated cells, increased concentrations of ppGpp, pppGpp and ATP were observed, but the GTP pool size decreased, suggesting that inhibition of RNA synthesis is a consequence of the stringent control mechanism imposed by pseudomonic acid-induced deprivation of an amino acid. Of the 20 common amino acids, only isoleucine reversed the inhibitory effect in vivo. The antibiotic was found to be a powerful inhibitor of isoleucyl-tRNA synthetase both in vivo and in vitro. Of seven other tRNA synthetases assayed, only a weak inhibitory effect on phenylalanyl-tRNA synthetase was observed; this presumably accounted for the weak effect on polyphenylalanine formation in a ribosomal preparation. Pseudomonic acid also significantly de-repressed threonine deaminase and transaminase B activity, but not dihydroxyacid dehydratase (isoleucine-biosynthetic enzymes) by decreasing the supply of aminoacylated tRNA(Ile). Pseudomonic acid is the second naturally occurring inhibitor of bacterial isoleucyl-tRNA synthetase to be discovered, furanomycin being the first.  相似文献   

9.
L-Asparaginase is known to catalyze the hydrolysis of L-asparagine to L-aspartic and ammonia, but little is known about its action on peptides. When we incubated L-asparaginases purified either from Escherichia coli or Erwinia chrysanthemi - commonly used as chemotherapeutic agents because of their antitumour activity - with eight small beta-aspartylpeptides such as beta-aspartylserineamide, beta-aspartylalanineamide, beta-aspartylglycineamide and beta-aspartylglycine, we found that both L-asparaginases could catalyze the hydrolysis of five of them yielding L-aspartic acid and amino acids or peptides. Our data show that L-asparaginases can hydrolyze beta-aspartylpeptides and suggest that L-asparaginase therapy may affect the metabolism of beta-aspartylpeptides present in human body.  相似文献   

10.
l-Asparagine synthetase was partially purified from mouse pancreas to a final mean specific activity of 0.10 unit/mg of protein. The enzyme exhibited an l-glutaminase activity which was not affected by l-asparate, NH(4)Cl, ATP-MgCl(2), l-glutamate, AMP (sodium salt) or sodium pyrophosphate. The l-glutamine-dependent l-asparagine synthetase activity of the partially purified enzyme from mouse pancreas was markedly decreased by freezing for 7 days at -87 degrees C in the presence of 1mm-dithiothreitol, but effectively protected from inactivation by high concentrations (10mm) of the thiol reagent. The l-glutaminase activity of the enzyme was inhibited by antagonists of l-glutamine (e.g. 6-diazo-5-oxo-l-norleucine, 5-chloro-4-oxo-l-norvaline, 5-diazo-4-oxo-l-norvaline and NSC-163501) and thiol-reactive compounds (e.g. 2-amino-4-arsenophenol hydrochloride, maleimide, mucochloric acid and ZnCl(2)), but not by aminomalonic acid, the next lower homologue of l-aspartate, nor by l-homoserine beta-adenylate, an analogue of the presumed transitory covalent intermediate. The complete forward reaction catalysed by l-asparagine synthetase from mouse pancreas appears to be irreversible and essentially stoicheiometric under the conditions examined. Mouse pancreas contains a proteolytic inhibitor of l-asparagine synthetase separable from the enzyme by ion-exchange column chromatography. The inhibitor is activated by incubation at 4 degrees C for 110h and inactivated by soya-bean trypsin inhibitor, di-isopropyl phosphorofluoridate and boiling.  相似文献   

11.
Asparagine synthetase catalyzes the ATP-dependent formation of L-asparagine from L-aspartate and L-glutamine, via a beta-aspartyl-AMP intermediate. Since interfering with this enzyme activity might be useful for treating leukemia and solid tumors, we have sought small-molecule inhibitors of Escherichia coli asparagine synthetase B (AS-B) as a model system for the human enzyme. Prior work showed that L-cysteine sulfinic acid competitively inhibits this enzyme by interfering with L-aspartate binding. Here, we demonstrate that cysteine sulfinic acid is also a partial substrate for E. coli asparagine synthetase, acting as a nucleophile to form the sulfur analogue of beta-aspartyl-AMP, which is subsequently hydrolyzed back to cysteine sulfinic acid and AMP in a futile cycle. While cysteine sulfinic acid did not itself constitute a clinically useful inhibitor of asparagine synthetase B, these results suggested that replacing this linkage by a more stable analogue might lead to a more potent inhibitor. A sulfoximine reported recently by Koizumi et al. as a competitive inhibitor of the ammonia-dependent E. coli asparagine synthetase A (AS-A) [Koizumi, M., Hiratake, J., Nakatsu, T., Kato, H., and Oda, J. (1999) J. Am. Chem. Soc. 121, 5799-5800] can be regarded as such a species. We found that this sulfoximine also inhibited AS-B, effectively irreversibly. Unlike either the cysteine sulfinic acid interaction with AS-B or the sulfoximine interaction with AS-A, only AS-B productively engaged in asparagine synthesis could be inactivated by the sulfoximine; free enzyme was unaffected even after extended incubation with the sulfoximine. Taken together, these results support the notion that sulfur-containing analogues of aspartate can serve as platforms for developing useful inhibitors of AS-B.  相似文献   

12.
L-asparagine uptake in Escherichia coli.   总被引:1,自引:1,他引:0       下载免费PDF全文
The uptake of L-asparagine by Escherichia coli K-12 is characterized by two kinetic components with apparent Km values of 3.5 muM and 80 muM. The 3.5 muM Km system displays a maximum velocity of 1.1 nmol/min per mg of protein, which is a low value when compared with derepressed levels of other amino acid transport systems but is relatively specific for L-asparagine. Compounds providing effective competition for L-asparagine uptake were 4-carbon analogues of the L-isomer with alterations at the beta-amide position, i.e., 5-diazo-4-oxo-L-norvaline (Ki = 4.6 muM), beta-hydroxyamyl-L-aspartic acid (Ki = 10 muM), and L-aspartic acid (Ki = 50 muM). Asparagine uptake is energy dependent and is inhibited by a number of metabolic inhibitors. In a derived strain of E. coli deficient in cytoplasmic asparaginase activity asparagine can be accumulated several-fold above the apparent biosynthetic pool of the amino acid and 100-fold above the external medium. The high affinity system is repressed by culture of cells with L-asparagine supplements in excess of 1 mM and is suggested to be necessary for growth of E. coli asparagine auxotrophs with lower supplement concentrations.  相似文献   

13.
Consequences of aspartase deficiency in Yersinia pestis.   总被引:2,自引:0,他引:2       下载免费PDF全文
Growing cells of Yersinia pseudotuberculosis, but not those of closely related Yersinia pestis, rapidly destroyed exogenous L-aspartic and L-glutamic acids, thus prompting a comparative study of dicarboxylic amino acid catabolism. Rates of amino acid metabolism by resting cells of both species were determined at pH 5.5, 7.0, and 8.5. Regardless of pH, Y. pseudotuberculosis destroyed L-glutamic acid, L-glutamine, L-aspartic acid, and L-asparagine at rates greater than those observed for Y. pestis. Although rates of proline degardation were similar, its metabolism by Y. pestis at pH 8.5 resulted in excretion of glutamic and aspartic acids. Similarly, Y. pestis excreted aspartic acid when incubated with L-glutamic acid (pH 8.5) or L-asparagine (pH 5.5, 7.0, and 8.5). Aspartase activity was not detected in extracts of 10 strains of Y. pestis but was present in all 11 isolates of Y. pseudotuberculosis. The latter contained significantly more glutaminase, asparaginase, and L-glutamate-oxalacetate transminase activity than did extracts of Y. pestis; specific activities of L-glutamate dehydrogenase and alpha-ketoglutarate dehydrogenase were similar. The observed differences in dicarboxylic amino acid metabolism are traceable to asparatase deficiency in Y. pestis and may account for the slow doubling time of this organism relative to Y. pseudotuberculosis.  相似文献   

14.
The pathway of 3,4-dihydroxyphenylalanine undergoing metabolism via transamination and subsequent oxidative rearrangement to 2,4,5-trihydroxyphenylacetate was investigated. 3-Methoxytyrosine does not pursue an identical course, since its corresponding keto acid is not subject to action of p-hydroxyphenylpyruvate hydroxylase. The radiochemical synthesis of 3-methoxy-4-hydroxy[carboxy-14C]pyruvate was accomplished. This metabolite was used to demonstrate that this keto acid does not proceed through oxidative rearrangement both in vitro and in vivo. The keto acid was found to be a competitive inhibitor of the hydroxylase and can help account for some of the metabolites observed in the urine of patients treated with 3,4-dihyroxyphenylalanine.  相似文献   

15.
本文从几方面探讨了产物L-天冬氨酸和L-天冬酰胺酶之间的相互作用。产物与珀琥酸对酶荧光影响的比较研究以及产物存在下对碘离子淬灭酶荧光性质的观察表明产物和酶之间存在直接的相互作用。产物保护天冬酰胺酶抗胰蛋白酶水解及抗热失活作用暗示它结合在酶的活性部位附近。  相似文献   

16.
The glutamate analog, alpha-aminomethylglutaric acid, was synthetized by Michael addition of ammonia to 2-methylene glutaronitrile followed by hydrolysis of the intermediate alpha-aminomethylglutaryl nitrile; the analog cyclizes readily on heating to 2-piperidone-5-carboxylic acid. Sheep brain glutamine synthetase utilizes one isomer of DL-alpha-aminomethylglutarate at about 10% of the rate with L-glutamate. gamma-Glutamylcysteine synthetase uses both isomers of DL-alpha-aminomethylglutarate, preferentially acting on the same isomer used by glutamine synthetase. gamma-(alpha-Aminomethyl)glutaryl-alpha-aminobutyrate, prepared enzymatically with gamma-glutamylcysteine synthetase, was found to be a substrate and an inhibitor of glutathione synthetase. alpha-Aminomethylglutarate does not inhibit gamma-glutamyl cyclotransferase and gamma-glutamyl transpeptidase appreciably. When alpha-aminomethylglutarate was administered to mice, there were substantial decreases in the levels of glutamine, glutathione, glutamate, and glycine in the kidney, and of glutamine and glutamate in the liver, indicating that this glutamate analog is effective as an inhibitor of glutamine and glutathione synthesis in vivo, and suggesting that it may also inhibit other enzymes.  相似文献   

17.
An important process for the regulation of auxin levels in plants is the inactivation of indole-3-acetic acid (IAA) by conjugation to amino acids. The conjugation reaction is catalysed by IAA-amido synthetases belonging to the family of GH3 proteins. Genetic approaches to study the biological significance of these enzymes have been hampered by large gene numbers and a high degree of functional redundancy. To overcome these difficulties a chemical approach based on the reaction mechanism of GH3 proteins was employed to design a small molecule inhibitor of IAA-amido synthetase activity. Adenosine-5'-[2-(1H-indol-3-yl)ethyl]phosphate (AIEP) mimics the adenylated intermediate of the IAA-conjugation reaction and was therefore proposed to compete with the binding of MgATP and IAA in the initial stages of catalysis. Two grapevine IAA-amido synthetases with different catalytic properties were chosen to test the inhibitory effects of AIEP in vitro. GH3-1 has previously been implicated in the grape berry ripening process and is restricted to two amino acid substrates, whereas GH3-6 conjugated IAA to 13 amino acids. AIEP is the most potent inhibitor of GH3 enzymes so far described and was shown to be competitive against MgATP and IAA binding to both enzymes with K(i)-values 17-68-fold lower than the respective K(m)-values. AIEP also exhibited in vivo activity in an ex planta test system using young grape berries. Exposure to 5-20 μM of the inhibitor led to decreased levels of the common conjugate IAA-Asp and reduced the accumulation of the corresponding Asp-conjugate upon treatment with a synthetic auxin. AIEP therefore represents a novel chemical probe with which to study IAA-amido synthetase function.  相似文献   

18.
19.
Aminoacyl-tRNA synthetases, a family of enzymes essential for protein synthesis, are promising targets of antimicrobials. Indolmycin, a secondary metabolite of Streptomyces griseus and a selective inhibitor of prokaryotic tryptophanyl-tRNA synthetase (TrpRS), was used to explore the mechanism of inhibition and to explain the resistance of a naturally occurring strain. Streptomyces coelicolor A3(2), an indolmycin-resistant strain, contains two trpS genes encoding distinct TrpRS enzymes. We show that TrpRS1 is indolmycin-resistant in vitro and in vivo, whereas TrpRS2 is sensitive. The lysine (position 9) in the enzyme tryptophan binding site is essential for making TrpRS1 indolmycin-resistant. Replacement of lysine 9 by glutamine, which at this position is conserved in most bacterial TrpRS proteins, abolished the ability of the mutant trpS gene to confer indolmycin resistance in vivo. Molecular modeling suggests that lysine 9 sterically hinders indolmycin binding to the enzyme. Tryptophan recognition (assessed by k(cat)/K(M)) by TrpRS1 is 4-fold lower than that of TrpRS2. Examination of the mRNA for the two enzymes revealed that only TrpRS2 mRNA is constitutively expressed, whereas mRNA for the indolmycin-resistant TrpRS1 enzyme is induced when the cells are exposed to indolmycin.  相似文献   

20.
We have synthesized a novel heteropolymer double-stranded RNA (dsRNA) molecule of defined length and strandedness (dsRNA309) and evaluated its ability to induce cytokine gene expression, activate dsRNA-dependent enzymes, and inhibit both tumor cell growth and virus replication. Unlike the conventionally studied synthetic homopolymer dsRNAs, polyinosinic acid:polycytidylic acid (poly(I-C)) and its mismatched analogue polyinosinic:polycytidylic, uridylic acid (poly(I-C12,U), dsRNA309 possessed restricted biological activity. dsRNA309 was unable to inhibit tumor cell growth or efficiently induce cytokine (i.e. interferon-beta and interleukin-1 alpha) gene expression. However, dsRNA309 was able to inhibit virus replication and activate dsRNA-dependent intracellular enzymes, 2'-5' oligoadenylate synthetase (2'-5' A synthetase) and the dsRNA-activated inhibitor kinase in in vitro assay systems. Overall, dsRNA309 provided a means for examining the mechanisms governing the dsRNA-regulated antiviral and antiproliferative responses, and studies with dsRNA309 demonstrated that the ability of a synthetic dsRNA to activate dsRNA-dependent intracellular enzymes does not necessarily predict the same gene inducing capacity.  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