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 15 毫秒
1.
Glycosylasparaginase catalyzes the hydrolysis of the N-glycosylic bond in N(4)-(2-acetamido-2-deoxy-beta-D-glucopyranosyl)-L-asparagine in the catabolism of N-linked oligosaccharides. A deficiency, or absence, of enzyme activity gives rise to aspartylglycosaminuria, the most common disorder of glycoprotein metabolism. The enzyme catalyzes the hydrolysis of a variety of asparagine and aspartyl compounds containing a free alpha-carboxyl group and a free alpha-amino group; computational studies suggest that the alpha-amino group actively participates in the catalytic mechanism. In order to study the importance of the alpha-carboxyl group and the alpha-amino group on the natural substrate to the reaction catalyzed by the enzyme, 14 analogues of the natural substrate were studied where the structure of the aspartyl group of the substrate was changed. The incremental binding energy (DeltaDeltaGb) for those analogues that were substrates was calculated. The results show that the alpha-amino group may be substituted with a group of comparable size, for the alpha-amino group contributes little, if any, to the transition state binding energy of the natural substrate. The alpha-amino group position acts as an "anchor" in the binding site for the substrate. On the other hand, the alpha-carboxyl group is necessary for enzyme activity; removal of the alpha-carboxyl group or changing it to an alpha-carboxamide group results in no hydrolysis reaction. Also, N-acetyl-D-glucosamine is not sufficient for binding to the active site for efficient hydrolysis by the enzyme. These results provide supporting evidence for a proposed intramolecular autoproteolytic activation reaction for the enzyme. However, the results raise a question as to an important role for the alpha-amino group in the catalytic mechanism as indicated in computational studies.  相似文献   

2.
The orthogonal synthesis of N alpha-Boc-L-aspartic acid-gamma-fluorenylmethyl ester and N alpha-Boc-L-glutamic acid-delta-fluorenylmethyl ester is reported. This is a four-step synthesis that relies on the selective esterification of the side-chain carboxyl groups on N alpha-CBZ-L-aspartic acid and N alpha-CBZ-L-glutamic acid. Such selectivity is accomplished by initially protecting the alpha-carboxyl group through the formation of the corresponding 5-oxo-4-oxazolidinone ring. Following side-chain esterification, the alpha-carboxyl and alpha-amino groups are deprotected with acidolysis. Finally, the alpha-amino group is reprotected with the t-butyl-oxycarbonyl (Boc) group. Thus aspartic acid and glutamic acid have their side-chain carboxyl groups protected with the base-labile fluorenylmethyl ester (OFm) and their alpha-amino groups protected with the acid-labile Boc group. These residues, when used in conjunction with N alpha-Boc-N epsilon-Fmoc-L-lysine, are important in the formation of side-chain to side-chain cyclizations, via an amide bridge, during solid-phase peptide synthesis.  相似文献   

3.
The use of Escherichia coli asparaginase II as a drug for the treatment of acute lymphoblastic leukemia is complicated by the significant glutaminase side activity of the enzyme. To develop enzyme forms with reduced glutaminase activity, a number of variants with amino acid replacements in the vicinity of the substrate binding site were constructed and assayed for their kinetic and stability properties. We found that replacements of Asp248 affected glutamine turnover much more strongly than asparagine hydrolysis. In the wild-type enzyme, N248 modulates substrate binding to a neighboring subunit by hydrogen bonding to side chains that directly interact with the substrate. In variant N248A, the loss of transition state stabilization caused by the mutation was 15 kJ mol(-1) for L-glutamine compared to 4 kJ mol(-1) for L-aspartic beta-hydroxamate and 7 kJ mol(-1) for L-asparagine. Smaller differences were seen with other N248 variants. Modeling studies suggested that the selective reduction of glutaminase activity is the result of small conformational changes that affect active-site residues and catalytically relevant water molecules.  相似文献   

4.
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   

5.
The substrate specificity of aspartokinase I has been examined by using both steady-state kinetic analyses and phosphorus-31 NMR spectroscopic studies. Analogues in which the alpha-amino group is either derivatized or replaced are not substrates or inhibitors for the enzyme, indicating the importance of the alpha-amino group as a binding determinant. The alpha-carboxyl group is not required for substrate recognition, and the alpha-amide or alpha-esters are competent alternative substrates. In addition, beta-derivatized structural analogues, such as the beta-hydroxamate, the beta-amide, or beta-esters, were found to be viable substrates. This was unexpected since the beta-carboxyl group is the usual site of phosphorylation. The nature of the acyl phosphate products obtained from these beta-derivatized alternative substrates has been characterized by coupled enzyme assays, oxygen-18-labeling studies, and phosphorus-31 NMR spectroscopy. These beta-derivatized analogues are capable of productive binding to aspartokinase through a reversal of regiospecificity to make the alpha-carboxyl group available as a phosphoryl acceptor. Many, but not all, of these alpha-acyl phosphates have also been shown to be viable substrates for the next two enzyme-catalyzed steps in this metabolic pathway. This raises the possibility of producing enzyme-generated alternative substrates that can serve as antimetabolites for the downstream reactions in this biosynthetic pathway.  相似文献   

6.
The apparent active site of human leukocyte glycoasparaginase (N4-(beta-acetylglucosaminyl)-L-asparaginase EC 3.5.1.26) has been studied by labeling with an asparagine analogue, 5-diazo-4-oxo-L-norvaline. Glycoasparaginase was purified 4,600-fold from human leukocytes with an overall recovery of 12%. The purified enzyme has a Km of 110 microM, a Vmax of 34 mumol x l-1 x min-1, and a specific activity of 2.2 units/mg protein with N4-(beta-N-acetylglucosaminyl)-L-asparagine as substrate. The carbohydrate content of the enzyme is 15%, and it exhibits a broad pH maximum between 7 and 9. The 88-kDa native enzyme is composed of 19-kDa light (L) chains and 25-kDa heavy (H) chains and it has a heterotetrameric structure of L2H2-type. The glycoasparaginase activity decreases rapidly and irreversibly in the presence of 5-diazo-4-oxo-L-norvaline. At any one concentration of the compound, the inactivation of the enzyme is pseudo-first-order with time. The inhibitory constant, K1, is 80 microM and the second-order rate constant 1.25 x 10(3) M-1 min-1 at pH 7.5. The enzyme activity is competitively protected against this inactivation by its natural substrate, aspartylglucosamine, indicating that this inhibitor binds to the active site or very close to it. The covalent incorporation of [5-14C]diazo-4-oxo-L-norvaline paralleled the loss of the enzymatic activity and one inhibitor binding site was localized to each L-subunit of the heterotetrameric enzyme. Four peptides with the radioactive label were generated, purified by high performance liquid chromatography, and sequenced by Edman degradation. The sequences were overlapping and all contained the amino-terminal tripeptide of the L-chain. By mass spectrometry, the reacting group of 5-diazo-4-oxo-L-norvaline was characterized as 4-oxo-L-norvaline that was bound through an alpha-ketone ether linkage to the hydroxyl group of the amino-terminal amino acid threonine.  相似文献   

7.
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   

8.
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   

9.
Reaction of an alpha-amino acid (alpha-AA) with 1,1-diphenylborinic acid (DPBA) leads to the formation of a kinetically stable adduct at pH 2-5 in which both the alpha-amino and the alpha-carboxyl groups are bound to boron forming a cyclic mixed anhydride termed a boroxazolidone. In this adduct, the greater than N:B bond is coordinate, involving the free electron pair of nitrogen, thereby satisfying the octet rule for the second electron shell of boron (Group IIIA). Consequently, the alpha-amino function of the boroxazolidone can be primary, secondary, or tertiary, as demonstrated by boroxazolidone formation with glycine, N-methylglycine, and N,N-dimethylglycine. On reaction with DPBA, the alpha-AA moiety of N-terminal gamma-glutamyl peptides is also derivatized as demonstrated by the formation of a glutathione boroxazolidone. The 1,1-diphenylboroxazolidone adducts of alpha-AA may be separated by reversed-phase (RP)-HPLC (AA-DPBA/RP-HPLC) enabling the derivatization procedure to be used as a precolumn reaction for alpha-AA analysis. Under the conditions we describe here, DPBA is not stably reactive with the epsilon-amino group of lysine. Furthermore, it does not complex with amide bonds of the peptide backbone or to any side chains of the common amino acids. Reaction of an alpha-AA mixture with DPBA, followed by RP-HPLC (AA-DPBA/RP-HPLC) is then a simple method by which to analyze alpha-AA in a mixture with peptides and amines. Precolumn reaction with DPBA may be used to separate peptides from alpha-AA and from those peptides which contain an alpha-AA moiety. Unreacted peptides are bound only weakly to the HPLC column and thus are separated from reacted alpha-amino acids which are retained as 1,1-diphenylboroxazolidones until their selective elution. This method is particularly suited for the analysis of alpha-amino acids that are derived from post-translational modification of protein side chains.  相似文献   

10.
L-Asparagine uptake by Stemphylium botryosum is mediated by two distinct energy- and temperature-dependent transport systems. One permease is relatively specific for L-asparagine and L-glutamine and is present in nutrient-sufficient mycelium. The specific permease shows an optimum pH at 5.2, saturation kinetics (Km = 4.4 x 10(-4) M, Vmax = 1.1 mumol/g per min), competitive gradient of L-asparagine, and higher affinity towards the L-isomer of asparagine. Amide derivatives of L-asparagine (5-diazo-4-oxo-L-norvaline or L-aspartyl hydroxamate) are the most effective competitors, alpha-amino derivative (N-acetyl asparagine) is a moderate competitor, and alpha-carboxyl derivative (L-asparagine-t-butylester) shows only slight inhibition of the specific permease. Derivatives of L-glutamine are significantly less effective competitors than those of L-asparatine. The level of the specific permease is affected by nitrogen sources and increases approximately threefold upon starvation. The nonspecific permease possesses an optimum pH at 6.8, saturation kinetics (Km = 7 x 10(-5) M, Vmax = 5 mumol/g per min, Kt = 7.4 x 10(-5) M for L-leucine), and high affinity towards various types of amino acids.  相似文献   

11.
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   

12.
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   

13.
The enzyme L-asparaginase (ASNASE), which hydrolyzes L-asparagine (L-Asn) to ammonia and L-aspartic acid (L-Asp), is commonly used for remission induction in acute lymphoblastic leukemia. To correlate ASNASE activity with L-Asn reduction in human serum, sensitive methods for the determination of ASNASE activity are required. Using L-aspartic beta-hydroxamate (AHA) as substrate we developed a sensitive plate reader-based method for the quantification of ASNASE derived from Escherichia coli and Erwinia chrysanthemi and of pegylated E. coli ASNASE in human serum. ASNASE hydrolyzed AHA to L-Asp and hydroxylamine, which was determined at 710 nm after condensation with 8-hydroxyquinoline and oxidation to indooxine. Measuring the indooxine formation allowed the detection of 2 x 10(-5)U ASNASE in 20 microl serum. Linearity was observed within 2.5-75 and 75-1,250 U/L with coefficients of correlation of r(2)>0.99. The coefficients of variation for intra- and interday variability for the three different ASNASE enzymes were 1.98 to 8.77 and 1.73 to 11.0%. The overall recovery was 101+/-9.92%. The coefficient of correlation for dilution linearity was determined as r(2)=0.986 for dilutions up to 1:20. This method combined with sensitive methods for the quantification of L-Asn will allow bioequivalence studies and individualized therapeutic drug monitoring of different ASNASE preparations.  相似文献   

14.
Site-specific mutagenesis has been used to probe amino acid residues proposed to be critical in catalysis by Escherichia coli asparaginase II. Thr12 is conserved in all known asparaginases. The catalytic constant of a T12A mutant towards L-aspartic acid beta-hydroxamate was reduced to 0.04% of wild type activity, while its Km and stability against urea denaturation were unchanged. The mutant enzyme T12S exhibited almost normal activity but altered substrate specificity. Replacement of Thr119 with Ala led to a 90% decrease of activity without markedly affecting substrate binding. The mutant enzyme S122A showed normal catalytic function but impaired stability in urea solutions. These data indicate that the hydroxyl group of Thr12 is directly involved in catalysis, probably by favorably interacting with a transition state or intermediate. By contrast, Thr119 and Ser122, both putative target sites of the inactivator DONV, are functionally less important.  相似文献   

15.
T-protein is a component of the glycine cleavage system and catalyzes the tetrahydrofolate-dependent reaction. To determine the folate-binding site on the enzyme, 14C-labeled methylenetetrahydropteroyltetraglutamate (5,10-CH2-H4PteGlu4) was enzymatically synthesized from methylenetetrahydrofolate (5, 10-CH2-H4folate) and [U-14C]glutamic acid and subjected to cross-linking with the recombinant Escherichia coli T-protein using 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide, a zero-length cross-linker between amino and carboxyl groups. The cross-linked product was digested with lysylendopeptidase, and the resulting peptides were separated by reversed-phase high performance liquid chromatography. Amino acid sequencing of the labeled peptides revealed that three lysine residues at positions 78, 81, and 352 were involved in the cross-linking with polyglutamate moiety of 5, 10-CH2-H4PteGlu4. The comparable experiment with 5,10-CH2-H4folate revealed that Lys-81 and Lys-352 were also involved in cross-linking with the monoglutamate form. Mutants with single or multiple replacement(s) of these lysine residues to glutamic acid were constructed by site-directed mutagenesis and subjected to kinetic analysis. The single mutation of Lys-352 caused similar increase (2-fold) in Km values for both folate substrates, but that of Lys-81 affected greatly the Km value for 5,10-CH2-H4PteGlu4 rather than for 5,10-CH2-H4folate. It is postulated that Lys-352 may serve as the primary binding site to alpha-carboxyl group of the first glutamate residue nearest the p-aminobenzoic acid ring of 5,10-CH2-H4folate and 5,10-CH2-H4PteGlu4, whereas Lys-81 may play a key role to hold the second glutamate residue through binding to alpha-carboxyl group of the second glutamate residue.  相似文献   

16.
Site-specific mutagenesis was used to replace the three histidine residues of Escherichia coli asparaginase II (EcA2) with other amino acids. The following enzyme variants were studied: [H87A]EcA2, [H87L]EcA2, [H87K]EcA2, [H183L]EcA2 and [H197L]EcA2. None of the mutations substantially affected the Km for L-aspartic acid beta-hydroxamate or impaired aspartate binding. The relative activities towards L-Asn, L-Gln, and l-aspartic acid beta-hydroxamate were reduced to the same extent, with residual activities exceeding 10% of the wild-type values. These data do not support a number of previous reports suggesting that histidine residues are essential for catalysis. Spectroscopic characterization of the modified enzymes allowed the unequivocal assignment of the histidine resonances in 1H-NMR spectra of asparaginase II. A histidine signal previously shown to disappear upon aspartate binding is due to His183, not to the highly conserved His87. The fact that [H183L]EcA2 has normal activity but greatly reduced stability in the presence of urea suggests that His183 is important for the stabilization of the native asparaginase tetramer. 1H-NMR and fluorescence spectroscopy indicate that His87 is located in the interior of the protein, possibly adjacent to the active site.  相似文献   

17.
H A Milman  R Muth  D A Cooney 《Enzyme》1979,24(1):36-47
Aminomalonic acid is a strong in vitro inhibitor of L-asparagine synthetase from Leukemia 5178Y/AR and from mouse pancreas; the agent is formally competitive with L-aspartic acid (Ki = 0.0023 M and 0.0015 M for the tumoral and pancreatic enzymes, respectively). Since aminomalonic acid is unstable and inert in vivo as an inhibitor of L-asparagine synthetase, attempts were made to deliver it to the site of its intended action via precursors: the diamide (2-aminomalonamide), the diester (diethylaminomalonate), and the keto acid (ketomalonic acid). Each of these putative 'pro drugs' was shown to be susceptible to metabolism to aminomalonate by mammalian and bacterial enzymes, in vitro. In vivo, aminomalonamide failed to inhibit tumoral L-asparagine synthetase at any time period up to 24 h after its oral or intraperitoneal administration. The diester and keto acid were similarly inactive. However, with specialized techniques it was possible to demonstrate that the diamide significantly inhibited the amidation and/or incorporation of L-aspartic acid into the L-asparaginyl residues of protein. Chemical manipulations of aminomalonic acid aimed at introducing irreversibly reacting functions are warranted.  相似文献   

18.
Acyl-peptide hydrolase from rat liver. Characterization of enzyme reaction   总被引:5,自引:0,他引:5  
Acyl-peptide hydrolase, which catalyzes the hydrolysis of an N-terminally acetylated peptide to release an N-acetylamino acid, was isolated from rat liver and found to be N-terminally blocked. The kinetics of the hydrolysis of acetyl (Ac)-Ala-Ala, Ac-Ala-Ala-Ala, acetylalanine p-nitroanilide, and acetylalanine beta-naphthylamide were investigated. The Km values were between 1 and 9 mM, and the Vmax values were between 100 and 500 nmol/min/micrograms of enzyme. The enzyme activity toward acetylalanine p-nitroanilide and acetylalanine beta-naphthylamide was activated by the presence of Cl- and SCN- at concentrations between 0.1 and 0.5 M. By contrast, the activity toward Ac-Ala-Ala and Ac-Ala-Ala-Ala was inhibited by these anions. Among a series of divalent cations, Zn2+ was demonstrated to be the most potent inhibitor. The enzyme was inactivated by the addition of diisopropyl fluorophosphate, diethyl pyrocarbonate. Woodward's Reagent K, and glycine methyl ester/carbodiimide. Titration by diisopropyl fluorophosphate showed 0.7 mol of active serine/mol of enzyme subunit, which was confirmed by the incorporation of [3H]diisopropyl fluorophosphate into the enzyme. Acetylalanine chloromethyl ketone inactivated the enzyme following pseudo-first order kinetics; and Ac-Ala, a competitive inhibitor, protected the enzyme from this inactivation. Acyl-peptide hydrolase appears to be a serine protease utilizing a charge relay system involving serine, histidine, and, probably, a carboxyl group(s). Two series of acetyl dipeptides, acetylamino acid p-nitroanilides and acetylamino acid beta-naphthylamides, were prepared in order to determine enzyme specificity. The enzyme preferentially removed Ac-Ala, Ac-Met, and Ac-Ser, the most common acetylated N-terminal residues (Persson, B., Flinta, C., von Heijne, G., and J?rnvall, H. (1985) Eur. J. Biochem. 152, 523-527). The enzyme was shown to be useful for deblocking peptides (e.g. alpha-melanocyte-stimulating hormone and acetyl-renin substrate), and the crude enzyme/substrate mixtures were amenable to direct protein sequence analysis.  相似文献   

19.
Porcine liver aminopeptidase B[EC 3.4.11.6] is highly specific for hydrolysis of beta-naphthylamides of basic L-amino acids; the Km values for L-arginine beta-naphthylamide and L-lysine beta-naphthylamide were 0.035 and 0.12 mM, respectively. The enzyme was inhibited by various alpha-amino acids. Among basic amino acids, L-homoarginine and L-arginine were the most potent inhibitors, L-lysine and L-norarginine (alpha-amino-gamma-guanidinobutyric acid) being less inhibitory. Hydrophobic amino acids also inhibited the enzyme competitively. This suggests that there is a hydrophobic region that binds the side chain of the substrates or inhibitors in the specificity site of the enzyme. Studies on the inhibitions by L-arginine derivatives showed that blocking of the alpha-carboxyl or the alpha-amino group reduced the inhibitory effect of L-arginine. Porcine liver aminopeptidase B was not inhibited by puromycin, whereas bestatin inhibited the enzyme competitively with a Ki value of 1.4 X 10(-8) M. This enzyme had no kinin-converting activity.  相似文献   

20.
Glycolato and S-lactato complexes containing the dioxomolybdenum(VI) moiety have been synthesized for studies on the role of the alpha-hydroxycarboxylato anion in the iron molybdenum cofactor of nitrogenase. The ligands in these complexes, vis K2[MoO2(glyc)2].H2O (H2glyc=glycolic acid, C2H4O3) (1) and (Na2[MoO2(S-lact)2])3.13H2O (H2lact=lactic acid, C3H6O3) (2) chelate through their alpha-alkoxyl and alpha-carboxyl oxygen atoms. In contrast, octanuclear K6[(MoO2)8(glyc)6(Hglyc)2].10H2O (3) formed by the reduction of the glycolato complex (1), features three different ligand binding modes: (i) non-bridging and bridging bidentate coordination of alpha-alkoxyl and alpha-carboxyl groups, and (ii) bidentate bridging using alpha-carboxyl group, leaving the alpha-alkoxyl group free. The octanuclear skeleton shows strong metal-metal interactions. The coordination modes in (1) and (2) mimic that of homocitrate to the iron molybdenum cofactor (FeMo-co) of nitrogenase. The bidentate coordination of alpha-alkoxyl and alpha-carboxyl groups shows that bond of alpha-carboxyl group to Mo is less susceptible to the oxidation state of molybdenum compared with the Mo-alpha-alkoxyl bond. This is supported by the dinuclear coordination of alpha-carboxyl group with free alpha-alkoxyl group in glycolato molybdate(V) (3).  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