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 359 毫秒
1.
The active sites of aspartate transcarbamoylase from Escherichia coli were titrated by measuring the decrease in the enzyme-catalyzed arsenolysis of N-carbamoyl-L-aspartate caused by the addition of the tight-binding inhibitor, N-phosphonacetyl-L-aspartate. Because the enzyme is a poor catalyst for this non-physiological reaction, high concentrations are required for the assays (more than 1000-fold the dissociation constant of the reversibly bound inhibitor) and, therefore, virtually all of the bisubstrate analog is bound. From the endpoint of the titration, 5.7 active sites were calculated, in excellent agreement with the number, six, based on the structure of the enzyme. Simple inhibition was observed only when the molar ratio of inhibitor to enzyme exceeded five; under these conditions, as shown in earlier physical chemical studies, the R-conformational state of the enzyme is the sole or predominant species. At low ratios of inhibitor to enzyme, the addition of inhibitor caused an increase in activity which is attributable to the conversion of the enzyme from the low-activity T-state to the much more active R-state. Comparison of the linear increase in activity as a function of inhibitor concentration at the low molar ratio (0.01, i.e. 1 inhibitor/600 active sites) with the activity lost at the high ratio provided a direct value for the mean number of active sites converted from the T-state to the R-state as a result of the binding of one bisubstrate analog to an enzyme molecule. This number was four with Mg X ATP or carbamoyl phosphate present and 4.7 for the enzyme in the presence of Mg X PPi, values approaching or identical to the theoretical maximum, 4.7, for a concerted transition with all of the active sites of the molecule changing from the T- to R-state upon the formation of a binary complex of hexameric enzyme with a single inhibitor. With the enzyme in the absence of effectors or with Mg X CTP present, the titrations showed that an average of two and one sites, respectively, of 4.7 possible, changed conformation upon ligand binding. These results were interpreted as a manifestation of an equilibrium between a sub-population of T- and R-state enzyme complexes containing one bound inhibitor molecule. The R-state species would represent 40% of the population for aspartate transcarbamoylase in the absence of extraneous ligands.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
Dansylated tight-binding inhibitors are effective fluorophoric probes for detecting conformational changes of enzyme active sites. In this study they have been employed to examine the effect of anions on the conformation of angiotensin-converting enzyme. The efficiency of radiationless energy transfer between enzyme tryptophan residues and an active site-bound dansyl inhibitor has been shown to be enhanced by the addition of chloride. Half-maximal fluorescence enhancement occurs at about 2 mM chloride and is the same for both N-(1-carboxyl-5-dansylamino-pentyl)-glycyl-L-phenylalanine [Ki,app = 50 nM (pH 7.5, 300 mM NaCl)] and N-(1-carboxyl-5-dansylamino-pentyl)-glycyl-L-lysine (Ki,app = 5.7 nM). Other activating anions also evoke similar increases in enzyme-inhibitor energy transfer. Fluorescence changes are not due to binding additional inhibitor molecules but rather to an anion-induced change in protein conformation.  相似文献   

3.
The telomeric G-rich single-stranded DNA can adopt in vitro an intramolecular quadruplex structure, which has been shown to directly inhibit telomerase activity. The reactivation of this enzyme in immortalized and most cancer cells suggests that telomerase is a relevant target in oncology, and telomerase inhibitors have been proposed as new potential anticancer agents. In this paper, we have analyzed the selectivity of four ethidium derivatives and ethidium itself toward different G-quadruplex species, with electrospray mass spectrometry and competitive equilibrium dialysis and evaluated their inhibitory properties against telomerase. A selectivity profile may be obtained through electrospray ionization mass spectrometry (ESI-MS), which is in fair agreement with competitive equilibrium dialysis data. It also provides unambiguous data on the number of binding sites per nucleic acid (maximal number of two ethidium derivatives per quadruplex, in agreement with external stacking). Our experiments also demonstrate that one compound (4) is the most active and selective G-quadruplex ligand within this series and the most selective telomerase inhibitor in a modified TRAP-G4 assay.  相似文献   

4.
Some preparations of both native aspartate transcarbamylase from Escherichia coli and catalytic subunit have fewer tight binding sites per oligomer for carbamyl-P than the number of catalytic peptide chains. In contrast, the number of sites for the tight-binding inhibitor N-(phosphonacetyl)-L-aspartate does equal the number of catalytic chains in each case. Binding of the labile carbamyl-P was determined using rapid gel filtration, with conversion to stable carbamyl-L-aspartate during collection. Native enzyme (six catalytic chains) obtained from cells grown under the conditions of J.C. Gerhart and H. Holoubek (J. Biol. Chem. (1967) 242, 2886-2892) has 5.4 tight sites for carbamyl-P at pH 8.0 (KD = 9.9 muM), whereas native enzyme from cells grown with higher concentrations of glucose, uracil, and histidine (to yield more enzyme per unit volume of culture) has only 1.9 tight sites at pH 8.0 (KD = 4.6 muM) and only 2.3 tight sites at pH 7.0 (KD = 2.6 muM). At pH 8.0, catalytic subunit (three catalytic chains) obtained from the former native enzyme has 2.2 tight sites for carbamyl-P (KD = 2.4 muM) and the number of sites is 2.3 in the presence of 35 mM succinate, whereas catalytic subunit obtained from the latter native enzyme has 1.8 tight sites (KD = 3.6 muM) in the absence of succinate and 2.3 tight sites in its presence. The number of tight binding sites is also less than the number of subunit peptide chains in 19F nuclear magnetic resonance experiments performed with catalytic subunit and two fluorinated analogs of carbamyl-P at comparable concentrations of analogs and active sites. A model is proposed in which incomplete removal of formylmethionine from the NH2 termini of the enzyme under conditions of extreme depression affects affinity for ligands.  相似文献   

5.
Abstract

A simple novel method was introduced for determination of an inhibitor binding constant (Kj) and enthalpy of binding by isothermal titration microcalorimetry technique. This method was applied to the binding of fluoride ion, as an inhibitor, with the active sites of jack bean urease at pH = 7.0 (Tris 30 mM) and T = 300°K. The dissociation equilibrium constant measured by this method was markedly consistent with the inhibition constant obtained from assay of enzyme activity in the presence of fluoride ion.  相似文献   

6.
Reaction microcalorimetry and equilibrium dialysis have been used to study the binding of AMP and IMP to glycogen phosphorylase b (EC 2.4.1.1) at 25 degrees C and pH 6.9. The combination of both techniques has enabled us to obtain some of the thermodynamic parameters for these binding processes. Four binding sites were found to be present in the dimeric active enzyme for both AMP and IMP. The binding to two high-affinity sites, which, in our opinion, correspond to the activator sites, seems to be cooperative. The two low-affinity sites, which would then correspond to the inhibitor sites, appear to be independent when the nucleotides bind to the enzyme. The negative delta G0 of binding/site at 25 degrees C is the result in all cases of a balance between negative enthalpy and entropy changes. The large differences in delta H and delta S0 for the binding of AMP to the activator sites (-27 and -70 kJ mol-1; -22 and -150 J X K-1 mol-1) suggest the existence of rather extensive conformational changes taking place in phosphorylase b on binding with the allosteric activator. Whereas the affinity of AMP for the activator sites is about 1 order of magnitude higher than that of IMP, the affinity of both nucleotides, including their delta H and delta S0 values, seems to be the same for the inhibitor sites.  相似文献   

7.
Binding of the potent radiolabelled competitive inhibitor 3H RU 44403 to pure human kidney angiotensin-I converting enzyme was examined in equilibrium and non equilibrium conditions. Equilibrium dialysis experiments indicate that, despite the duplicated structure of the enzyme and the presence of two putative active sites, 3H RU 44403 interacts with a single high affinity (Kd = 0.44 +/- 0.05 x 10-9 M, n = 3) binding site. This suggests that only one of the two putative active sites is functional, and can bind substrates or inhibitors. Sodium chloride plays an essential role in the enzyme-inhibitor interaction. The formation of the complex is only slightly influenced by NaCl, but the kinetic of dissociation is dramatically dependent on NaCl concentration. In a Nacl free medium the complex is unstable and dissociates rapidly. These results are consistent with the hypothesis that chloride ion influences isomerization of the complex toward a more stable form.  相似文献   

8.
The interaction of leech-derived tryptase inhibitor (LDTI) with bovine liver capsule tryptase (BLCT) and bovine trypsin has been studied using both thermodynamic and kinetic approaches. Several differences were detected: (i) the equilibrium affinity of LDTI for BLCT (Ka = 8.9 x 10(5) M(-1)) is about 600-fold lower than that for bovine trypsin (Ka = 5.1 x 10(8) M(-1)); (ii) LDTI behaves as a purely non-competitive inhibitor of BLCT, while it is a purely competitive inhibitor of bovine trypsin. These functional data are compared with those previously reported for the LDTI binding to human tryptase, where tight inhibition occurs at two of the four active sites of the tetramer (Ka = 7.1 x 10(8) M(-1)). Amino acid sequence alignment of BLCT, human betaII-tryptase and bovine trypsin allows us to infer some possible structural basis for the observed functional differences.  相似文献   

9.
The transient state kinetics of catalysis for dihydrofolate reductase (DHFR) from several enzyme sources including highly purified recombinant human enzyme (rHDHFR) have been examined. Like DHFR from Escherichia coli, the enzyme from Lactobacillus casei, and isoenzyme 2 from Streptococcus faecium exhibit a slow increase in activity upon addition of substrates to enzyme. No slow hysteresis of this type was detected with recombinant human DHFR (rHDHFR) or DHFR from chicken or bovine liver or L1210 mouse leukemia cells (MDHFR). In contrast, both rHDHFR and MDHFR exhibited a very rapid decrease in activity (t1/2 = 30 and 20 ms, respectively) during a phase that occurred after the first turnover of the enzyme but before establishment of the steady state. This intermediate phase was not observed for the bacterial enzymes or the avian enzyme, nor was it observed with a mutant of rHDHFR in which Phe-31 has been replaced by leucine. For rHDHFR the intermediate phase is not a consequence of product inhibition, substrate depletion, or enzyme instability. It may therefore be concluded that this unusual transient state kinetic behavior results from the existence of two conformers of the enzyme, one of which has a higher turnover number than the other with the equilibrium shifting in favor of the less active conformer during the course of catalysis. The equilibrium is particularly favorable for the less active conformer when NADP is present in the active site of rHDHFR, whereas bound tetrahydrofolate favors the more active conformer. The more active conformer has a 6-fold higher Km for dihydrofolate than does the less active conformer. The existence of these conformers is likely to produce cooperative behavior by rHDHFR in vivo.  相似文献   

10.
The mode of binding of acetyl-pepstatin to the protease from the human immunodeficiency virus type 1 (HIV-1) has been determined by x-ray diffraction analysis. Crystals of an acetyl-pepstatin-HIV-1 protease complex were obtained in space group P2(1)2(1)2 (unit cell dimensions a = 58.39 A, b = 86.70 A, c = 46.27 A) by precipitation with sodium chloride. The structure was phased by molecular replacement methods, and a model for the structure was refined using diffraction data to 2.0 A resolution (R = 0.176 for 12901 reflections with I greater than sigma (I); deviation of bond distances from ideal values = 0.018 A; 172 solvent molecules included). The structure of the protein in the complex has been compared with the structure of the enzyme without the ligand. A core of 44 amino acids in each monomer, including residues in the active site and residues at the dimer interface, remains unchanged on binding of the inhibitor (root mean square deviation of alpha carbon positions = 0.39 A). The remaining 55 residues in each monomer undergo substantial rearrangement, with the most dramatic changes occurring at residues 44-57 (these residues comprise the so-called flaps of the enzyme). The flaps interact with one another and with the inhibitor so as to largely preserve the 2-fold symmetry of the protein. The inhibitor is bound in two approximately symmetric orientations. In both orientations the peptidyl backbone of the inhibitor is extended; a network of hydrogen bonds is formed between the inhibitor and the main body of the protein as well as between the inhibitor and the flaps. Hydrophobic side chains of residues in the body of the protein form partial binding sites for the side chains of the inhibitor; hydrophobic side chains of residues in the flaps complete these binding sites.  相似文献   

11.
CO dehydrogenase from Clostridium thermoaceticum is a nickel-containing enzyme that catalyzes both the reversible conversion of CO2 to CO (for incorporation into the carbonyl group of acetate) and the synthesis of acetyl-CoA from methyl corrinoid, CO, and CoASH. The latter activity is conveniently assayed by monitoring the exchange of [1-14C]acetyl-CoA (carbonyl group) with 12CO. Kinetic parameters for the highly oxygen sensitive exchange activity have been determined: Km (acetyl-CoA) = 600 microM; Vmax = 440 min-1. In addition, coenzyme A analogues have been tested as inhibitors of the exchange to probe the active site of the enzyme; each has no effect on the CO2 in equilibrium CO activity of CO dehydrogenase. Coenzyme A, the substrate for acetate biosynthesis, is a potent competitive inhibitor, KI = 7 microM. Comparison of this value with that for desulfo-CoA (KI = 6000 microM) suggests that a key mode of binding is through the sulfur atom, possibly to a metal site on the enzyme. The relatively high affinity of the enzyme for CoASH relative to acetyl-CoA is consistent with its proposed operation in the acetogenic direction. The differential sensitivity to oxygen and storage of the two activities of CO dehydrogenase as well as the contrasting effect of coenzyme A inhibitors suggests that acetate assemblage occurs at a site distinct from that for CO dehydrogenation.  相似文献   

12.
J K Wright  M Takahashi 《Biochemistry》1977,16(8):1541-1548
The aspartokinase activity of the aspartokinase-homoserine dehydrogenase complex of Escherichia coli was affinity labeled with substrates ATP, aspartate, and feedback inhibitor threonine. Exchange-inert ternary adducts of Co(III)-aspartokinase and either ATP, aspartate or threonine were formed by oxidation of corresponding Co(II) ternary complexes with H2O2. The ternary enzyme-Co(III)-threonine adduct (I) had 3.8 threonine binding sites per tetramer, one-half that of the native enzyme. The binding of threonine to I was still cooperative as determined by equilibrium dialysis (nH = 2.2) or by studying inhibition of residual dehydrogenase activity (nH = 2.7). Threonine still protected the SH groups of I against 5,5'-dithiobis(2-nitrobenzoate) (DTNB) reaction but the number of SH groups reacting with thiol reagents (DTNB) was reduced by 1-2 per subunit in the absence of threonine. This suggests either that Co(III) is bound to the enzyme via sulfhydryl groups or that 1-2SH groups are buried or rendered inaccessible in I. The binding of threonine to sites not blocked by the affinity labeling produced changes in the circular dichroism of the complex comparable to changes produced by threonine binding to native enzyme and also protected against proteolytic digestion. The major conformational changes produced by threonine are thus ascribable to binding at this one class of regulatory sites. The interactions of kinase substrates with various aspartokinase-Co(III) complexes containing ATP, aspartate, or threonine and a threonine-insensitive homoserine dehydrogenase produced by mild proteolysis were studied. The inhibition of homoserine dehydrogenase by kinase substrates is not due to binding of these inhibitors at the kinase active site but was shown to be due to binding to sites within the dehydrogenase domain of the enzyme. L-alpha-Aminobutyrate, a presumed threonine analogue, also inhibits the dehydrogenase by binding at the same or similar sites in the dehydrogenase domain and not at threonine regulatory site.  相似文献   

13.
S M Janes  J P Klinman 《Biochemistry》1991,30(18):4599-4605
Recent evidence has shown that the active site cofactor in bovine serum amine oxidase (BSAO) is 2,4,5-trihydroxyphenylalanine or 6-hydroxydopa [Janes et al. (1990) Science 248, 981]. However, much ambiguity remains regarding the mechanism of the enzymatic reaction. Conflicting data exist for both the number of functional active sites in the dimeric enzyme and for the oxygen dependence of product release. To resolve these questions, a new method has been developed for the purification of BSAO which leads to the isolation of specific activity greater than or equal to 0.4 unit/mg of enzyme in 2-3 weeks. This highly active enzyme has been used to quantitate both aldehyde and ammonia release in the reductive half-reaction. Anaerobic incubation of enzyme and substrate resulted in the production of 2 mol of aldehyde/mol of enzyme, indicating the presence of a cofactor at each enzyme subunit. As anticipated for an aminotransferase reaction, no ammonia release was detected under comparable conditions. Active site titration of enzyme samples of varying specific activity with phenylhydrazine extrapolates to 1 mol of inhibitor/mol of enzyme subunit for BSAO of specific activity = 0.48 unit/mg. These findings contrast with numerous, previous reports of only one functional cofactor per enzyme dimer in copper amine oxidases.  相似文献   

14.
The effect of salts on the catalytic activity of the molybdenum-containing nitrogenase complex from Azotobacter vinelandii has been investigated. NaCl was found to inhibit the reduction of the substrates, protons, acetylene, and dinitrogen by a common mechanism. The pattern of inhibition is sigmoidal, indicating a highly cooperative interaction involving multiple inhibitor sites. Sixteen other salts that were investigated also exhibited this pattern of inhibition. NaCl functions as a dead-end inhibitor without altering the number of MgATP hydrolyzed/electron transferred to substrate. The level of expressed inhibition is sensitive to MgATP concentration, the molar ratio of the MoFe-protein (Av1) to the Fe-protein (Av2), and total protein concentration. In addition, NaCl is an inhibitor of the MgATP-dependent, iron chelation of Av2. Although the inhibition is exhibited over the same salt concentration range as that for inhibition of substrate reduction, the pattern of inhibition is hyperbolic. A model based upon simple equilibrium interactions among the enzyme species, nucleotides, and inhibitor has been developed which quantitatively accounts for the observed effects of salt. In this model, the formation of the active complex between Av1 and Av2 is abolished by salts. Likewise, the apparent affinity of Av2 for MgATP is reduced. An additional prediction based upon the model is that the affinity between Av2 and Av1 is independent of nucleotide binding.  相似文献   

15.
1. From the standpoint of monomer-dimer equilibrium of hog kidney D-amino acid oxidase [EC 1.4.3.3] and the interaction between the enzyme and small molecules, the effect of pH on the binding of p-aminobenzoate to the monomer and dimer of the enzyme was studied by kinetic methods and spectrophotometric titration. 2. The maximum binding number of p-aminobenzoate to the dimer is two molecules, and there is no interaction between the two active sites of the dimer (i.e., no cooperativity) over the range of pH from 6.5 to 10. 3. The affinity of the dimer for p-aminobenzoate is several times higher than that of the monomer at pH 6.5-10, and consequently p-aminobenzoate induces dimerization in the equilibrium state of D-amino acid oxidase. The interaction energy of two subunits of the dimer is stabilized by the binding of p-aminobenzoate by 1-2 kcal/mole over the pH range studied. 4. The binding sites of the quasi-substrate, p-aminobenzoate, in the dimer and the intersubunit binding site of the dimer are clearly different, because p-aminobenzoate induces dimerization of the enzyme. 5. The pK values of ionizing groups in the free monomer and the free dimer which participate in the binding of the competitive inhibitor, p-aminobenzoate, are approximately the same, 8.7, as determined from the pH dependence of the affinity of the inhibitor for the enzyme. Furthermore, no pK for the enzyme-inhibitor complex in the pH range 6.5-10 was observed. 6. There is no interaction between the two ionizing groups of the dimer during protonation-deprotonation, because a theoretical equation involving no cooperativity between the two ionizing groups in the dimer explains the results well.  相似文献   

16.
H C Shin  D M Quinn 《Biochemistry》1992,31(3):811-818
The interaction of lipoprotein lipase with p-nitrophenyl N-alkylcarbamates [PNPOC(=O)-NHCnH2n+1; n = 4, 8, and 12] proceeds by the three-stage mechanism shown below. After reversible [formula: see text] formation of the enzyme-carbamate complex (EC), rapid carbamylation (kc) precedes slow decarbamylation. Therefore, in short-term assays (less than or equal to 30 min) of lipoprotein lipase catalyzed hydrolysis of p-nitrophenyl butyrate, activity is rapidly lost. The inhibition by p-nitrophenyl N-butylcarbamate follows saturation kinetics, which allows determination of Kc = 5.4 +/- 0.9 microM and kc = (4.9 +/- 0.7) x 10(-2)s-1. Saturation kinetics are not observed for the longer inhibitors p-nitrophenyl N-octylcarbamate and p-nitrophenyl N-dodecylcarbamate. Rather, plots of the pseudo-first-order rate constant for activity loss versus inhibitor concentration are concave upward, consistent with inhibitor binding to two sites on the enzyme. The inhibition phase is sufficiently rapid that p-nitrophenyl N-octylcarbamate can be used to titrate enzyme active sites. On the other hand, long-term assays (greater than 5 h) show sequential inhibition and activity return phases, and from the activity return phase kd is calculated. The long-term activity time course is accurately simulated by Runge-Kutta integration of the differential equations for the three-stage mechanism. These approaches have been used to characterize the kinetics of interaction of the enzyme with the carbamate inhibitors.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
The shape of the plots of product accumulation versus time (t) has been analysed for slowly equilibrating association-dissociation enzyme systems of the types 2p in equilibrium P (P is enzyme oligomer which is able to dissociate reversibly forming two identical halves p) and M in equilibrium M2 in equilibrium M2 in equilibrium... (M is monomer which has two association sites overlapping with active sites). It is assumed that the rate of equilibration between oligomeric forms is comparable with the rate of over-all enzymatic reaction and that substrate-oligomer complexes are in rapid equilibrium with free components. It has been shown that characteristic feature of kinetic behavior of slowly equilibrating association-dissociation enzyme systems is that the value of tau depends on enzyme concentration (tau is the intercept on t-axis for linear asymptota of the curve of product concentration versus time at t leads to infinity).  相似文献   

18.
The effect of methylation of histidine-57 of alpha-chymotrypsin with Streptomyces subtilisin inhibitor was examined. Methylchymotrypsin was isolated by affinity chromatography on inhibitor-Sepharose, and the interaction of this inactive enzyme with inhibitor was quantitatively analyzed by two different methods: the spectrophotometric titration of difference spectrum resulted in the complex formation and the application of competitive enzyme assay by using substrates of large Km values. The former method gave values of 8.6 . 10(-6) M as dissociation constant (Kd) of methylchymotrypsin . inhibitor complex and 0.91 as the number of binding sites (n) per inhibitor monomer, both of which were almost equivalent to those for native enzyme . inhibitor complex. By the latter novel method, values of 7.9 . 10(-6) M and 1.08 were obtained for Kd and n, respectively, for interaction of inhibitor with alpha-chymotrypsin, and 8 . 10(-6) M as Kd for methylchymotrypsin . inhibitor complex. These results indicate that methylation of histidine-57 of active site in alpha-chymotrypsin molecule does not affect essentially the binding ability to inhibitor and the modified enzyme binds stoichiometrically to inhibitor, as the native enzyme does, with a molar ratio of 1:1 per inhibitor monomer.  相似文献   

19.
It was shown that AMP, an allosteric inhibitor of fructose-1.6-bisphosphatase, decreases the apparent affinity of the enzyme for the activating cation, Mg2+, which is accompanied by a decrease of the kinetic cooperativity between the Mg2+-binding sites. In its turn, the Mg2+ increase diminishes the enzyme sensitivity to the inhibiting effect of AMP and decreases the cooperativity of the inhibitor binding. The heterotropic interactions between the allosteric inhibitor and activator binding centers are consistent with the predictions of the Monod-Wyman-Changeux model which involves two conformational states of the enzyme (of which one is catalytically inactive) differing in their affinity for the ligands. An increase in pH from 7.4 to 9.0 increases the enzyme affinity for Mg2+ and causes an equilibrium shift towards the catalytically active state of the enzyme.  相似文献   

20.
K Duncan  C T Walsh 《Biochemistry》1988,27(10):3709-3714
In Salmonella typhimurium, D-alanine:D-alanine ligase (ADP) (EC 6.3.2.4) is the second enzyme in the three enzyme D-alanine branch pathway of peptidoglycan biosynthesis. The interaction of this enzyme with a possible transition-state analogue, the (aminoalkyl)phosphinate D-3-[(1-aminoethyl)phosphinyl]-2-heptylpropionic acid [Parsons et al. (1987) Abstracts of Papers, 193rd National Meeting of the American Chemical Society, Denver, CO, MEDI 63, American Chemical Society, Washington, DC], has been studied. This compound is a potent active site directed inhibitor and is competitive with D-alanine (Ki = 1.2 microM); it exhibits time-dependent inhibition in the presence of ATP. Kinetic analysis revealed a rapid onset of steady-state inhibition (kon = 1.35 X 10(4) M-1 s-1) followed by slow dissociation of inhibitory complex(es) with a half-life of 8.2 h. The inhibitory complex was shown to consist of E...I...ATP in equilibrium with E...I, Pi, and ADP. Similar time-dependent inhibition was also observed with D-(1-aminoethyl)phosphonic acid (D-Ala-P) (Ki = 0.5 mM; kon = 27 M-1 s-1; t1/2 for regain = 1.73 min) but not with D-(1-aminoethyl)phosphinic acid, which behaved as a simple competitive inhibitor (Ki = 0.4 mM). The mechanism of inhibition is discussed in the light of the precedents of glutamine synthase inhibition by methionine sulfoximine and phosphinothricin.  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