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.
Interaction of camel lens zeta-crystallin with the hydrophobic probe 1-anilinonaphthalene-8-sulfonic acid (ANS) enhanced the ANS fluorescence and quenched the protein fluorescence. Both of these events were concentration-dependent and showed typical saturation curves suggesting specific ANS-zeta-crystallin binding. Quantitative analysis indicated that 1 mole zeta-crystallin bound at most 1 mole ANS. NADPH but not 9,10-phenanthrenequinone (PQ) was able to displace zeta-crystallin-bound ANS. These results suggested the presence of a hydrophobic domain in zeta-crystallin, possibly at the NADPH binding site. alpha-Crystallin as well as NADPH protected zeta-crystallin against thermal inactivation suggesting the importance of this site for enzyme stability. The NADPH:quinone oxidoreductase activity of zeta-crystallin was inhibited by ANS with NADPH as electron donor and PQ as electron acceptor. Lineweaver-Burk plots indicated mixed-type inhibition with respect to NADPH, with a K(i) of 2.3 microM. Secondary plots of inhibition with respect to NADPH indicated a dissociation constant (K'I) of 12 microM for the zeta-crystallin-NADPH-ANS complex. The K(i) being smaller than K'I suggested that competitive inhibition at the NADPH binding site was predominant over non-competitive inhibition. Like ANS-zeta-crystallin binding, inhibition was dependent on ANS concentration but independent of incubation time.  相似文献   

2.
Interaction of camel lens zeta-crystallin with aspirin was investigated by activity and fluorescence measurements. Aspirin minimally inhibited the oxidoreductase activity of the enzyme and weakly quenched its fluorescence. However, significant fluorescence quenching of zeta-crystallin coincided with the appearance of a fluorescence signal characteristic of salicylic acid thereby raising the possibility that salicylic acid might have been the moiety responsible for inhibition and fluorescence quenching. Direct fluorescence measurements showed that zeta-crystallin had a much higher affinity for salicylic acid than aspirin (K(i) of about 24 microM for salicylic acid versus 630 microM for aspirin). Salicylic acid was also far more effective in inhibiting zeta-crystallin than aspirin (K(i) values were 23 microM versus 820 microM, respectively). Inhibition kinetics suggested that salicylic acid interacted with zeta-crystallin via a binding site that was distinct from that of NADPH. Salicylic acid also interacted with and quenched the fluorescence of camel lens alpha-crystallin suggesting a general mode of interaction with lens proteins. Within the normal therapeutic concentrations of salicylic acid or aspirin, only crystallin-salicylic acid interactions might be significant. These results showed that camel lens zeta- and alpha-crystallin exhibited remarkable selectivity for salicylic acid over aspirin, and thus, could be considered as salicylate-binding proteins.  相似文献   

3.
Interaction of camel lens zeta-crystallin, an NADPH:quinone oxidoreductase, with several quinone derivatives was examined by fluorescence spectroscopy and activity measurements. Fluorescence of zeta-crystallin was quenched to different levels by the different quinones:juglone (5-OH, 1,4 naphthoquinone), 1,4 naphthoquinone (1,4-NQ), and 1,2 naphthoquinone (1,2-NQ) considerably quenched the fluorescence of zeta-crystallin, where as the commonly used substrate, 9,10-phenanthrenequinone (PQ) did not induce significant quenching. Activity measurements showed only PQ served as a substrate for camel lens zeta-crystallin, while juglone, 1,4-NQ, and 1,2-NQ were inhibitors. Thus quinones that interacted with zeta-crystallin directly inhibited the enzyme, whereas the substrate had very low affinity for the enzyme in the absence of NADPH. Another substrate, dichlorophenol indophenol (DCIP), conformed to the same pattern; DCIP did not quench the fluorescence of the enzyme significantly, but served as a substrate. This pattern is consistent with an ordered mechanism of catalysis with quinone being the second substrate. All three naphthoquinones were uncompetitive inhibitors with respect to NADPH and noncompetitive with respect to PQ. These kinetics are similar to those exhibited by cysteine- and/or lysine-modifying agents. Juglone, 1,4-NQ, and 1,2-NQ interacted with and quenched the fluorescence of camel lens alpha-crystallin, but to lesser extent than that of zeta-crystallin.  相似文献   

4.
Fluorescence spectrum of camel lens zeta-crystallin, a major protein in the lens of camelids and histicomorph rodents, showed maximum emission at 315 nm. This emission maximum is blue shifted compared to most proteins, including alpha-crystallin, and appeared to be due to tryptophan in highly hydrophobic environment. Interaction of NADPH with zeta-crystallin quenched the protein fluorescence and enhanced the fluorescence of bound NADPH. Analysis of fluorescence quenching suggested high-affinity interaction between NADPH and zeta-crystallin with an apparent Km<0.45 microM. This value is at least an order of magnitude lower than that suggested by activity measurements. Analysis of NADPH fluorescence showed a biphasic curve representing fluorescence of free- and bound-NADPH. The intersection between free- and bound-NADPH closely paralleled the enzyme concentration, suggesting one mole of NADPH was bound per subunit of the enzyme. Phenanthrenequinone (PQ), the substrate of zeta-crystallin, also was able to quench the fluorescence of zeta-crystallin, albeit weaker than NADPH. Quantitative analysis suggested that zeta-crystallin had low affinity for PQ in the absence of NADPH, and PQ binding induced significant conformational changes in zeta-crystallin.  相似文献   

5.
Homoserine dehydrogenase (HSD), which is required for the synthesis of threonine, isoleucine and methionine in fungi, is a potential target for novel antifungal drugs. In order to design effective inhibitors, the kinetic mechanism of Saccharomyces cerevisiae HSD and the stereochemistry of hydride transfer were examined. Product inhibition experiments revealed that yeast HSD follows an ordered Bi Bi kinetic mechanism, where NAD(P)H must bind the enzyme prior to aspartate semialdehyde (ASA) and homoserine is released first followed by NAD(P)+. H-(1,2,4-triazol-3-yl)-D,L-alanine was an uncompetitive inhibitor of HSD with respect to NADPH (K(ii)=3.04+/-0.18 mM) and a noncompetitive inhibitor with respect to ASA (K(is)=1.64+/-0.36 mM, K(ii)=3.84+/-0.46 mM), in agreement with the proposed substrate order. Both kinetic isotope and viscosity experiments provided evidence for a very rapid catalytic step and suggest nicotinamide release to be primarily rate limiting. Incubation of HSD with stereospecifically deuterated NADP[2H] and subsaturating amounts of aspartate semialdehyde revealed that the pro-S NADPH hydride is transferred to the aldehyde. The pH dependence of steady state kinetic parameters indicate that ionizable groups with basic pKs may be involved in substrate binding, consistent with the observation of Lys223 at the enzyme active site in the recently determined 3D structure [B. DeLaBarre, P.R. Thompson, G.D. Wright, A.M. Berghuis, Nat. Struct. Biol. 7 (2000) 238-244]. These findings provide the requisite foundation for future exploitation of fungal HSD in inhibitor design.  相似文献   

6.
Different classes of folate analogues have been examined with respect to the mechanism of their inhibition of dihydrofolate reductases from Escherichia coli and chicken liver. In addition, the degree of synergism between the binding of these compounds and NADPH has been investigated. Methotrexate acts as a slow, tight-binding inhibitor of both enzymes whereas trimethoprim is a slow, tight-binding inhibitor of the enzyme from E. coli and a classical inhibitor of the chicken-liver enzyme. Pyrimethamine, 2,4-diamino-6,7-dimethylpteridine, a phenyltriazine, folate and folinate exhibit classical inhibition. The degree of synergism between the binding of NADPH and the inhibitor varied from low for pyrimethamine and folate to very large for the phenyltriazine which binds to the chicken-liver enzyme almost 50 000-times more tightly in the presence of NADPH. The degree of synergism is reflected in the type of inhibition that the folate analogues yield with respect to NADPH. Compounds which exhibit slight synergism give noncompetitive inhibition whereas those with a high degree of synergism yield uncompetitive inhibition. With the exception of folinate, all compounds that act as classical inhibitors give rise to competitive inhibition with respect to dihydrofolate. Folinate exhibits competitive inhibition against NADPH and noncompetitive inhibition against dihydrofolate. These results are consistent with the formation of an enzyme-dihydrofolate-folinate complex. The (6S, alphaS)-diastereoisomer of folinate was bound at least 1000-times more tightly than the (6R, alphaS)-diastereoisomer. Consideration has been given to the possible interactions that occur between residues on the enzyme and groups on the inhibitor that give rise to slow-binding inhibition.  相似文献   

7.
Glutathione reductase (GR, type IV, Baker's yeast, E.C 1.6.4.2) is a flavoprotein that catalyzes the NADPH-dependent reduction of oxidized glutathione (GSSG) to reduced glutathione (GSH). In this study some metal ions have been tested on GR; lithium, manganese, molybdate, aluminium, barium, zinc, calcium, cadmium and nickel. Cadmium, nickel and calcium showed a good to moderate inhibitory effect on yeast GR. GR is inhibited non-competitively by Zn2+ (up to 2 mM) and activated above this concentration. Ca2+ inhibition was non-competitive with respect to GSSG and uncompetitive with respect to NADPH. Nickel inhibition was competitive with respect to GSSG and uncompetitive with respect to NADPH. The inhibition constants for these metals on GR were determined. The chelating agent EDTA recovered 90% of the GR activity inhibited by these metals.  相似文献   

8.
The alpha-aminoadipate-semialdehyde dehydrogenase (EC 1.2.1.31) of Trichosporon adeninovorans, an enzyme of lysine biosynthesis, was partially purified, some properties of the enzyme were studied and a novel regulatory pattern was found. The Km values of the enzyme were estimated to be 0.78 mM for alpha-aminoadipate, 1.0 mM for ATP, 0.23 mM for NADPH and 0.77 mM for MgCl2. It is demonstrated that the enzyme can be regulated by lysine and lysine analogues. L-Lysine (Ki of 0.09 mM), S-(beta-aminoethyl)-L-cysteine (Ki of 0.007 mM) and delta-hydroxylysine (Ki of 1.65 mM) inhibited the enzyme activity. The inhibition was competitive with respect to alpha-aminoadipate and non-competitive with respect to both ATP and NADPH.  相似文献   

9.
Parikh SL  Xiao G  Tonge PJ 《Biochemistry》2000,39(26):7645-7650
Structural and genetic studies indicate that the antibacterial compound triclosan, an additive in many personal care products, is an inhibitor of EnvM, the enoyl reductase from Escherichia coli. Here we show that triclosan specifically inhibits InhA, the enoyl reductase from Mycobacterium tuberculosis and a target for the antitubercular drug isoniazid. Binding of triclosan to wild-type InhA is uncompetitive with respect to both NADH and trans-2-dodecenoyl-CoA, with K(i)' values of 0.22+/-0.02 and 0.21+/-0.01 microM, respectively. Replacement of Y158, the catalytic tyrosine residue, with Phe, reduces the affinity of triclosan for the enzyme and results in noncompetitive inhibition, with K(i) and K(i)' values of 36+/-5 and 47+/-5 microM, respectively. Consequently, the Y158 hydroxyl group is important for triclosan binding, suggesting that triclosan binds in similar ways to both InhA and EnvM. In addition, the M161V and A124V InhA mutants, which result in resistance of Mycobacterium smegmatis to triclosan, show significantly reduced affinity for triclosan. Inhibition of M161V is noncompetitive with K(i)' = 4.3+/-0.5 microM and K(i) = 4.4+/-0.9 microM, while inhibition of A124V is uncompetitive with K(i)' = 0. 81 +/- 0.11 microM. These data support the hypothesis that the mycobacterial enoyl reductases are targets for triclosan. The M161V and A124V enzymes are also much less sensitive to isoniazid compared to the wild-type enzyme, indicating that triclosan can stimulate the emergence of isoniazid-resistant enoyl reductases. In contrast, I47T and I21V, two InhA mutations that occur in isoniazid-resistant clinical isolates of M. tuberculosis, show unimpaired inhibition by triclosan, with uncompetitive inhibition constants (K(i)') of 0.18+/-0.01 and 0.12+/- 0.01 microM, respectively. The latter result indicates that InhA inhibitors targeted at the enoyl substrate binding site may be effective against existing isoniazid-resistant strains of M. tuberculosis.  相似文献   

10.
Dihydrodipicolinate synthase (DHDPS, E.C. 4.2.1.52) mediates the first unique reaction of (S)-lysine biosynthesis in plants and microbes-the condensation of (S)-aspartate-beta-semialdehyde ((S)-ASA) and pyruvate. It has been shown that DHDPS is partially feedback inhibited by (S)-lysine; it is suggested that this mechanism regulates flux through the DAP biosynthetic pathway. Others have characterised DHDPS from Escherichia coli with respect to (S)-lysine inhibition. They have concluded that, with respect to pyruvate, the first substrate of the reaction, DHDPS shows uncompetitive inhibition: as such, they further suggest that (S)-lysine inhibits DHDPS via interaction with the binding site for the second substrate, (S)-ASA. Yet, this finding is based on the assumption that (S)-lysine is a fully uncompetitive inhibitor. In light of crystallographic studies, which lead to the proposal that (S)-lysine affects the putative proton-relay of DHDPS, we re-evaluated the inhibition mechanism of DHDPS with respect to (S)-lysine by incorporating the observed hyperbolic inhibition. Our data showed that lysine is not an uncompetitive inhibitor, but a mixed inhibitor when pyruvate and (S)-lysine concentrations were varied. Thus, consistent with the crystallographic data, (S)-lysine must have an effect on the initial steps of the DHDPS reaction, including the binding of pyruvate and Schiff base formation.  相似文献   

11.
A fatty acid synthetase multienzyme complex was purified from Euglena gracilis variety bacillaris. The fatty acid synthetase activity is specifically inhibited by antibodies against Escherichia coli acyl-carrier protein. The Euglena enzyme system requires both NADPH and NADH for maximal activity. An analysis was done of the steady-state kinetics of the reaction catalysed by the fatty acid synthetase multienzyme complex. Initial-velocity studies were done in which the concentrations of the following pairs of substrates were varied: malonyl-CoA and acetyl-CoA, NADPH and acetyl-CoA, malonyl-CoA and NADPH. In all three cases patterns of the Ping Pong type were obtained. Product-inhibition studies were done with NADP+ and CoA. NADP+ is a competitive inhibitor with respect to NADPH, and uncompetitive with respect to malonyl-CoA and acetyl-CoA. CoA is uncompetitive with respect to NADPH and competitive with respect to malonyl-CoA and acetyl-CoA. When the concentrations of acetyl-CoA and malonyl-CoA were varied over a wide range, mutual competitive substrate inhibition was observed. When the fatty acid synthetase was incubated with radiolabelled acetyl-CoA or malonyl-CoA, labelled acyl-enzyme was isolated. The results are consistent with the idea that fatty acid synthesis proceeds by a multisite substituted-enzyme mechanism involving Ping Pong reactions at the following enzyme sites: acetyl transacylase, malonyl transacylase, beta-oxo acyl-enzyme synthetase and fatty acyl transacylase.  相似文献   

12.
Heavy metal pollution can arise from many sources and damage many organisms. Exposure to the metal ions can leads to a reduction in cellular antioxidant enzyme activities and lowers cellular defense against oxidative stress. In this study we have tested effects of the some metal ions on the purified bovine kidney cortex glutathione reductase (GR). Cadmium (Cd2+), nickel (Ni2+), and zinc (Zn2+) showed inhibitory effect on the enzyme. The obtained IC?? values of Cd2+, Ni2+, and Zn2+ are 0.027, 0.8, and 1 mM, respectively. Kinetic characterization of the inhibition is also investigated. Cd2+ inhibition is noncompetitive with respect to both oxidized glutathione (GSSG) (Ki(GSSG) 0.060 ± 0.005 mM) and NADPH (Ki(NADPH) 0.025 ± 0.002 mM). Ni2+ inhibition is noncompetitive with respect to GSSG (Ki(GSSG) 0.329 ± 0.016 mM) and uncompetitive with respect to NADPH (Ki(NADPH) 0.712 ± 0.047 mM). The effect of Zn2+ on GR activity is consistent with noncompetitive inhibition pattern when the varied substrate is the GSSG (Ki(GSSG) 0.091 ± 0.005 mM) and the NADPH (Ki(NADPH) 0.226 ± 0.01 mM), respectively. GR inhibition studies may be useful for understanding the mechanisms for oxidative damage associated with heavy metal toxicity.  相似文献   

13.
Glutathione reductase (GR, type IV, Baker's yeast, E.C 1.6.4.2) is a flavoprotein that catalyzes the NADPH-dependent reduction of oxidized glutathione (GSSG) to reduced glutathione (GSH). In this study some metal ions have been tested on GR; lithium, manganese, molybdate, aluminium, barium, zinc, calcium, cadmium and nickel. Cadmium, nickel and calcium showed a good to moderate inhibitory effect on yeast GR. GR is inhibited non-competitively by Zn2 + (up to 2 mM) and activated above this concentration. Ca2 + inhibition was non-competitive with respect to GSSG and uncompetitive with respect to NADPH. Nickel inhibition was competitive with respect to GSSG and uncompetitive with respect to NADPH. The inhibition constants for these metals on GR were determined. The chelating agent EDTA recovered 90% of the GR activity inhibited by these metals.  相似文献   

14.
Nitrate reductase (NADPH:nitrate oxidoreductase; EC 1.6.6.1-3) was purified to apparent homogeneity from mycelium of Penicillium chrysogenum. The final preparation catalyzed the NADPH-dependent, FAD-mediated reduction of nitrate with a specific activity of 170-225 units X mg of protein-1. Gel filtration and glycerol density centrifugation yielded, respectively, a Stokes radius of 6.3 nm and an s20,w of 7.4. The molecular weight was calculated to be 199,000. On sodium dodecyl sulfate gels, the enzyme displayed two almost contiguous dye-staining bands corresponding to molecular weights of about 97,000 and 98,000. The enzyme prefers NADPH to NADH (kspec ratio = 2813), FAD to FMN (kspec ratio = 141), FAD (+ NADPH) to FADH2 (kspec ratio = 12,000), and nitrate to chlorate (kspec ratio = 4.33), where the kspec (the specificity constant for a given substrate) represents Vmax/Km. The Penicillium enzyme will also catalyze te NADPH-dependent, FAD-mediated reduction of cytochrome c with a specific activity of 647 units X mg of protein-1 (Kmcyt = 1.25 X 10(-5) M), and the reduced methyl viologen (MVH2, i.e. methyl viologen + dithionite)-dependent, NADPH and FAD-independent reduction of nitrate with a specific activity of 250 units X mg of protein-1 kmMVH2 = 3.5 X 10(-6) M). Initial velocity studies showed intersecting NADPH-FAD and nitrate-FAD reciprocal plot patterns. The NADPH-nitrate pattern was a series of parallel lines at saturating and unsaturating FAD levels. NADP+ was competitive with NADPH, uncompetitive with nitrate (at saturating and unsaturating FAD levels), and a mixed-type inhibitor with respect to FAD. Nitrite was competitive with nitrate, uncompetitive with NADPH (at saturating and unsaturating FAD levels), and a mixed-type inhibitor with respect to FAD. At unsaturating nitrate and FAD, NADPH exhibited substrate inhibition, perhaps as a result of binding to the FAD site(s). At very low FAD concentrations, low concentrations of NADP+ activated the reaction slightly. The initial velocity and product inhibition patterns are consistent with either of the two kinetic mechanisms. One (rather unlikely) mechanism involves the rapid equilibrium random binding of all ligands with (a) NADP+ and NADPH mutually exclusive, (b) nitrate and nitrite mutually exclusive, (c) the binding of NADPH strongly inhibiting the binding of nitrate and vice versa, (d) the binding of NADPH strongly promoting the binding of nitrite and vice versa, and (e) the binding of nitrate strongly promoting the binding of NADP+ and vice versa...  相似文献   

15.
Aspirin may exert part of its antithrombotic effects through platelet-independent mechanisms. Diabetes is a condition in which the beneficial effects of aspirin are less prominent or absent - a phenomenon called "aspirin resistance". We investigated whether acetylation and glycation occur at specific sites in fibrinogen and if competition between glucose and aspirin in binding to fibrinogen occurs. Our hypothesis was that such competition might be one explanation to "aspirin resistance" in diabetes. After incubation of fibrinogen in vitro with aspirin (0.8 mM, 24 h) or glucose (100 mM, 5-10 days), we found 12 modified sites with mass spectrometric techniques. Acetylations in the α-chain: αK191, αK208, αK224, αK429, αK457, αK539, αK562, in the β-chain: βK233, and in the γ-chain: γK170 and γK273. Glycations were found at βK133 and γK75, alternatively γK85. Notably, the lysine 539 is a site involved in FXIII-mediated cross-linking of fibrin. With isotope labeling in vitro, using [(14)C-acetyl]salicylic acid and [(14)C]glucose, a labeling of 0.013-0.084 and 0.12-0.5 mol of acetylated and glycated adduct/mol fibrinogen, respectively, was found for clinically (12.9-100 μM aspirin) and physiologically (2-8 mM glucose) relevant plasma concentrations. No competition between acetylation and glycation could be demonstrated. Thus, fibrinogen is acetylated at several lysine residues, some of which are involved in the cross-linking of fibrinogen. This may mechanistically explain why aspirin facilitates fibrin degradation. We find no support for the idea that glycation of fibrin(ogen) interferes with acetylation of fibrinogen.  相似文献   

16.
The inhibition of xanthine oxidase by its reaction product, uric acid, was studied by steady state kinetic analysis. Uric acid behaved as an uncompetitive inhibitor of xanthine oxidase with respect to the reducing substrate, xanthine. Under 50 microM xanthine and 210 microM oxygen, the apparent K(i) for uric acid was 70 microM. Uric acid-mediated xanthine oxidase inhibition also caused an increase in the percentage of univalent reoxidation of the enzyme (superoxide radical production). Steady-state rate equations derived by the King-Altman method support the formation of an abortive-inhibitory enzyme-uric acid complex (dead-end product inhibition). Alternatively, inhibition could also depend on the reversibility of the classical ping-pong mechanism present in xanthine oxidase-catalyzed reactions.  相似文献   

17.
2-C-Methyl-D-erythritol-4-phosphate synthase (MEP synthase) catalyzes the rearrangement/reduction of 1-D-deoxyxylulose-5-phosphate (DXP) to methylerythritol-4-phosphate (MEP) as the first pathway-specific reaction in the MEP biosynthetic pathway to isoprenoids. Recombinant E. coli MEP was purified by chromatography on DE-52 and phenyl-Sepharose, and its steady-state kinetic constants were determined: k(cat) = 116 +/- 8 s(-1), K(M)(DXP) = 115 +/- 25 microM, and K(M)(NADPH) = 0.5 +/- 0.2 microM. The rearrangement/reduction is reversible; K(eq) = 45 +/- 6 for DXP and MEP at 150 microM NADPH. The mechanism for substrate binding was examined using fosmidomycin and dihydro-NADPH as dead-end inhibitors. Dihydro-NADPH gave a competitive pattern against NADPH and a noncompetitive pattern against DXP. Fosmidomycin was an uncompetitive inhibitor against NADPH and gave a pattern representative of slow, tight-binding competitive inhibition against DXP. These results are consistent with an ordered mechanism where NADPH binds before DXP.  相似文献   

18.
Efforts to identify novel compounds capable of blocking the steroid 5 alpha-reductase (SR) catalyzed conversion of testosterone (T) to 5 alpha-dihydrotestosterone have resulted in the development of 17 beta-substituted-3-androstene-3-carboxylic acids as potent inhibitors of the rat prostatic enzyme. The dead-end inhibition patterns of one of these steroidal acrylates, 17 beta-N-(2-methyl-2-propyl)-carbamoyl-androst-3,5-diene-3-carboxylic acid were best evaluated with a linear uncompetitive kinetic model vs both T (Kii = 11 +/- 1 nM) and NADPH (Kii = 22 +/- 2 nM). To interpret these observations, the kinetic mechanism of the rat prostatic SR was shown to involve the binding of NADPH prior to that of T through a series of dead-end and product inhibition experiments. Within the context of this preferentially ordered kinetic mechanism, it is proposed that the uncompetitive inhibition patterns result from the association of the steroidal acrylate to an enzyme complex containing NADP+ in formation of a dead-end ternary complex of enzyme, NADP+, and inhibitor.  相似文献   

19.
Xu H  West AH  Cook PF 《Biochemistry》2006,45(39):12156-12166
Kinetic data have been measured for the histidine-tagged saccharopine dehydrogenase from Saccharomyces cerevisiae, suggesting the ordered addition of nicotinamide adenine dinucleotide (NAD) followed by saccharopine in the physiologic reaction direction. In the opposite direction, the reduced nicotinamide adenine dinucleotide (NADH) adds to the enzyme first, while there is no preference for the order of binding of alpha-ketoglutarate (alpha-Kg) and lysine. In the direction of saccharopine formation, data also suggest that, at high concentrations, lysine inhibits the reaction by binding to free enzyme. In addition, uncompetitive substrate inhibition by alpha-Kg and double inhibition by NAD and alpha-Kg suggest the existence of an abortive E:NAD:alpha-Kg complex. Product inhibition by saccharopine is uncompetitive versus NADH, suggesting a practical irreversibility of the reaction at pH 7.0 in agreement with the overall K(eq). Saccharopine is noncompetitive versus lysine or alpha-Kg, suggesting the existence of both E:NADH:saccharopine and E:NAD:saccharopine complexes. NAD is competitive versus NADH, and noncompetitive versus lysine and alpha-Kg, indicating the combination of the dinucleotides with free enzyme. Dead-end inhibition studies are also consistent with the random addition of alpha-Kg and lysine. Leucine and oxalylglycine serve as lysine and alpha-Kg dead-end analogues, respectively, and are uncompetitive against NADH and noncompetitive against alpha-Kg and lysine, respectively. Oxaloacetate (OAA), pyruvate, and glutarate behave as dead-end analogues of lysine, which suggests that the lysine-binding site has a higher affinity for keto acid analogues than does the alpha-Kg site or that dicarboxylic acids have more than one binding mode on the enzyme. In addition, OAA and glutarate also bind to free enzyme as does lysine at high concentrations. Glutarate gives S-parabolic noncompetitive inhibition versus NADH, indicating the formation of a E:(glutarate)2 complex as a result of occupying both the lysine- and alpha-Kg-binding sites. Pyruvate, a slow alternative keto acid substrate, exhibits competitive inhibition versus both lysine and alpha-Kg, suggesting the combination to the E:NADH:alpha-Kg and E:NADH:lysine enzyme forms. The equilibrium constant for the reaction has been measured at pH 7.0 as 3.9 x 10(-7) M by monitoring the change in NADH upon the addition of the enzyme. The Haldane relationship is in very good agreement with the directly measured value.  相似文献   

20.
The transport of lysine has been investigated in epithelial cells isolated from chicken jejunum. The kinetics of lysine transport and the pattern of interaction with zwitterionic amino acids were consistent with system b(0,+) activity, the broad-spectrum and Na(+)-independent amino acid transporter. The half-saturation constant for lysine entry (K(m)+/-S.E.) was 0.029+/-0.002 mM and the flux was not affected significantly by Na(+) replacement with choline. Lysine influx was inhibited by L-leucine both in Na(+) and choline medium with inhibition constants (K(i)+/-S.E.) 0.068+/-0.006 mM (in Na(+)) and 0.065+/-0.009 mM (in choline). Other inhibitory amino acids (K(i)+/-S.E.) were (mM): L-tyrosine (0.073+/-0.018), L-methionine (0.15+/-0.015), L-cystine (0.42+/-0.04), L-cysteine (1.1+/-0.07), L-isoleucine (1.1+/-0.09), L-glutamine (1.8+/-0.16) and L-valine (2.5+/-0.13). Lysine exit was trans-accelerated (approx. 20 fold) by 2 mM L-lysine and L-leucine. The flux was resistant to pretreatment of the cells with p-chloromercuriphenylsulfonate (0.2 mM), which is an inhibitor of system y(+)L, the broad-spectrum and cation-modulated transporter.  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