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 21 毫秒
1.
In the x-ray structure of the human dihydrofolate reductase, phenylalanine 31 and phenylalanine 34 have been shown to be involved in hydrophobic interactions with bound substrates and inhibitors. Using oligonucleotide-directed mutagenesis and a bacterial expression system producing the wild-type and mutant human dihydrofolate reductases at levels of 10% of the bacterial protein, we have constructed, expressed, and purified a serine 31 (S31) mutant and a serine 34 (S34) mutant. Fluorescence titration experiments indicated that S31 bound the substrate H2folate 10-fold tighter and the coenzyme NADPH 2-fold tighter than the wild-type human dihydrofolate reductase. The serine 31 mutation had little effect on the steady-state kinetic properties of the enzyme but produced a 100-fold increase in the dissociation constant (Kd) for the inhibitor methotrexate. The serine 34 mutant had much greater alterations in its properties than S31; specifically, S34 had a 3-fold reduction in the Km for NADPH, a 24-fold increase in the Km for H2folate, a 3-fold reduction in the overall reaction rate kcat, and an 80,000-fold increase in the Kd for methotrexate. In addition, the pH dependence of the steady-state kinetic parameters of S34 were different from that of the wild-type enzyme. These results suggest that phenylalanine 31 and phenylalanine 34 make very different contributions to ligand binding and catalysis in the human dihydrofolate reductase.  相似文献   

2.
3.
We have previously described methotrexate-resistant Chinese hamster ovary cells which appear to contain normal levls of a structurally altered dihydrofolate reductase (EC 1.5.1.3) (Flintoff, W.F., Davidson, S.V., and Siminovitch, L. (1976) Somatic Cell Genet.2,245-261). By selecting for increased resistance form these class I cells, class III resistant cells were isolated which appeared to possess an increased activity of the altered enzyme. In the report, we describe the purification and several properties of the reductase from wild-type cells, two independently selected class I cells, and class III resistant cell. The reductases from wild-type and resistant cells had similar specific activities using folate and dihydrofolate as substrates, and similar molecular weights as determined by sodium dodecyl sulfate gel electrophoresis. The mutant enzymes, however, were about six- to eight-fold more resistant to inhibition by methotrexate than the wild-type enzyme, suggesting a decreased affinity of the mutant reductases to methotrexate-binding. Small differences between various enzymes were also seen in other physicochemical properties such as pH optima and Km values for folate, and in their heat stabilities, which suggest that different structural alterations may lead to the same mutant phenotype. As expected from earlier studies with crude extracts, class III cells did produce a higher (about 10-fold) yield of the reductase than the class I or wild-type cells.  相似文献   

4.
Oligonucleotide-directed, site-specific mutagenesis was used to convert phenylalanine-31 of human recombinant dihydrofolate reductase (DHFR) to leucine. This substitution was of interest in view of earlier chemical modification studies (Kumar et al., 1981) and structural studies based on X-ray crystallographic data (Matthews et al., 1985a,b) which had implicated the corresponding residue in chicken liver DHFR, Tyr-31, in the binding of dihydrofolate. Furthermore, this particular substitution allowed testing of the significance of protein sequence differences between mammalian and bacterial reductases at this position with regard to the species selectivity of trimethoprim. Both wild-type (WT) and mutant (F31L) enzymes were expressed and purified by using a heterologous expression system previously described (Prendergast et al., 1988). Values of the inhibition constants (Ki values) for trimethoprim were 1.00 and 1.08 microM for WT and F31L, respectively. Thus, the presence of phenylalanine at position 31 in human dihydrofolate reductase does not contribute to the species selectivity of trimethoprim. The Km values for nicotinamide adenine dinucleotide phosphate (reduced) (NADPH) and dihydrofolate were elevated 10.8-fold and 9.4-fold, respectively, for the mutant enzyme, whereas the Vmax increased only 1.8-fold. Equilibrium dissociation constants (KD values) were obtained for the binding of NADPH and dihydrofolate in binary complexes with each enzyme. The KD for NADPH is similar in both WT and F31L, whereas the KD for dihydrofolate is 43-fold lower in F31L. Values for dihydrofolate association rate constants (kon) with enzyme and enzyme-NADPH complexes were measured by stopped-flow techniques.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Pediococcus cerevisiae/AMr, resistant to amethopterin, possesses a higher dihydrofolate reductase (5, 6, 7, 8-tetrahydrofolate: NADP+ oxidoreductase, EC 1.5.1.3) activity than the parent, a folate-permeable and thus amethopterin-susceptible strain and than the wild-type. The properties of dihydrofolate reductase from the three strains have been compared. Temperature, pH optima, heat stability, as well amethopterin binding did not reveal significant differences between the enzymes from the susceptible and resistant strains. The enzyme from the wild-type was 10 times more sensitive to inhibition by amethopterin and more susceptible to heat denaturation. The apparent Km values for dihydrofolate in enzymes from the three strains were in the range of 4.8--7.2 muM and for NADPH 6.5--8.0 muM. The amethopterin-resistant strain exhibited cross-resistance to trimethoprim and was about 40-fold more resistant to the latter than the sensitive parent and the wild-type. The resistance to trimethoprim appears to be a direct result of the increased dihydrofolate reductase activity. Inhibition of dihydrofolate reductase activity by this drug was similar in the three strains. 10--20 nmol caused 50% inhibition of 0.02 enzyme unit. Trimethoprim was about 10 000 times less effective inhibitor of dihydrofolate reductase than amethopterin. The cell extract of the AMr strain possessed a folate reductase activity three times higher than that of the sensitive strain. The activities of other folate-related enzymes like thymidylate synthetase and 10-formyltetrahydrofolate synthetase (formate: tetrahydrofolate ligase (ADP-forming), EC 6.3.4.3) were similar in the three strains studied.  相似文献   

6.
Arginine-70 of human dihydrofolate reductase (hDHFR) is a highly conserved residue which X-ray crystallographic data have shown to interact with the alpha-carboxylate of the terminal L-glutamate moiety of either folic acid or methotrexate (MTX). The rationale for this study was to introduce a conservative amino acid residue change at position 70 (Arg----Lys) which might function as a titratable group and, thus, reveal possible quantitative changes in ligand binding and kinetic parameters as a function of pH. Such a mutant enzyme (R70K) has been constructed and expressed by using site-directed mutagenesis techniques. This substitution has a dramatic effect on the binding of MTX, which displays a 22,600-fold increase in the dissociation constant (KD) at pH 7.5 compared to that of the reported wild-type enzyme value. At this pH, the KD value for dihydrofolate (FAH2) for the R70K enzyme shows only a 7-fold increase over that for the wild-type hDHFR. The pH profiles of the Michaelis and dissociation constants for FAH2 and KD values for MTX for the mutant enzyme all show a 7-8-fold increase from pH 7.5 to 8.5 as compared to its wild-type counterpart. The binding of NADPH or the nonclassical inhibitor trimetrexate (TMQ) to either the wild-type or the mutant enzyme does not show such pH-dependent characteristics. Thus, since FAH2 and MTX interact with the guanidinium side chain of arginine-70 in the wild-type hDHFR, the replacement of this residue with a lysine in the R70K mutant appears to have resulted in the introduction of a titratable group with a perturbed pKa value of ca. 8.3.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
We have applied site-directed mutagenesis methods to change the conserved tryptophan-22 in the substrate binding site of Escherichia coli dihydrofolate reductase to phenylalanine (W22F) and histidine (W22H). The crystal structure of the W22F mutant in a binary complex with the inhibitor methotrexate has been refined at 1.9-A resolution. The W22F difference Fourier map and least-squares refinement show that structural effects of the mutation are confined to the immediate vicinity of position 22 and include an unanticipated 0.4-A movement of the methionine-20 side chain. A conserved bound water-403, suspected to play a role in the protonation of substrate DHF, has not been displaced by the mutation despite the loss of a hydrogen bond with tryptophan-22. Steady-state kinetics, stopped-flow kinetics, and primary isotope effects indicate that both mutations increase the rate of product tetrahydrofolate release, the rate-limiting step in the case of the wild-type enzyme, while slowing the rate of hydride transfer to the point where it now becomes at least partially rate determining. Steady-state kinetics show that below pH 6.8, kcat is elevated by up to 5-fold in the W22F mutant as compared with the wild-type enzyme, although kcat/Km(dihydrofolate) is lower throughout the observed pH range. For the W22H mutant, both kcat and kcat/Km(dihydrofolate) are substantially lower than the corresponding wild-type values. While both mutations weaken dihydrofolate binding, cofactor NADPH binding is not significantly altered. Fitting of the kinetic pH profiles to a general protonation scheme suggests that the proton affinity of dihydrofolate may be enhanced upon binding to the enzyme. We suggest that the function of tryptophan-22 may be to properly position the side chain of methionine-20 with respect to N5 of the substrate dihydrofolate.  相似文献   

8.
We identify His381 of Pseudomonas mevalonii 3-hydroxy-3-methylglutaryl-coenzyme A (HMG-CoA) reductase as the basic residue functional in catalysis. The catalytic domain of 20 HMG-CoA reductases contains a single conserved histidine (His381 of the P. mevalonii enzyme). Diethyl pyrocarbonate inactivated the P. mevalonii enzyme, and hydroxylamine partially restored activity. We changed His381 to alanine, lysine, asparagine, and glutamine. The mutant proteins were overexpressed, purified to homogeneity, and characterized. His381 mutant enzymes were not inactivated by diethyl pyrocarbonate. All four mutant enzymes exhibited wild-type crystal morphology and chromatographed on substrate affinity supports like wild-type enzyme. The mutant enzymes had low catalytic activity (Vmax 0.06-0.5% that of wild-type enzyme), but Km values approximated those for wild-type enzyme. For wild-type enzyme and mutant enzymes H381A, H381N, and H381Q, Km values at pH 8.1 were 0.45, 0.27, 3.7, and 0.71 mM [(R,S)-mevalonate]; 0.05, 0.03, 0.20, and 0.11 mM [coenzyme A]; 0.22, 0.14, 0.81, and 0.62 mM [NAD+]. Km values at pH 11 for wild-type enzyme and mutant enzyme H381K were 0.32 and 0.75 mM [(R,S)-mevalonate]; 0.24 and 0.50 mM [coenzyme A]; 0.15 and 1.23 mM [NAD+]. Both pK values for the enzyme-substrate complex increased relative to wild-type enzyme (by 1-2.5 pH units for pK1 and by 0.5-1.3 pH units for pK2). For mutant enzyme H381K, the pK1 of 10.2 is consistent with lysine acting as a general base at high pH. His381 of P. mevalonii HMG-CoA reductase, and consequently the histidine of the consensus Leu-Val-Lys-Ser-His-Met-Xaa-Xaa-Asn-Arg-Ser motif of the catalytic domain of eukaryotic HMG-CoA reductases, thus is the general base functional in catalysis.  相似文献   

9.
Using oligonucleotide-directed mutagenesis of the gene encoding the small subunit (rbcS) from Anacystis nidulans mutant enzymes have been generated with either Trp-54 of the small subunit replaced by a Phe residue, or with Trp-57 replaced by a Phe residue, whereas both Trp-54 and Trp-57 have been replaced by Phe residues in a double mutant. Trp-54 and Trp-57 are conserved in all amino acid sequences or the small subunit (S) that are known at present. The wild-type and mutant forms of Rubisco have all been purified to homogeneity. The wild-type enzyme, purified from Escherichia coli is indistinguishable from enzyme similarly purified from A. nidulans in subunit composition, subunit molecular mass and kinetic parameters (Vmax CO2 = 2.9 U/mg, Km CO2 = 155 microM). The single Trp mutants are indistinguishable from the wild-type enzyme by criteria (a) and (b). However, whereas, Km CO2 is also unchanged, Vmax CO2 is 2.5-fold smaller than the value for the wild-type enzyme for both mutants, demonstrating for the first time that single amino acid replacements in the non-catalytic small subunit influence the catalytic rate of the enzyme. The specificity factor tau, which measures the partitioning of the active site between the carboxylase and oxygenase reactions, was found to be invariant. Since tau is not affected by these mutations we conclude that S is an activating not a regulating subunit.  相似文献   

10.
Site-directed mutagenesis was used to generate mutants of recombinant mouse dihydrofolate reductase to test the role of some amino acids in the binding of two inhibitors, methotrexate and trimethoprim. Eleven mutations changing eight amino acids at positions all involved in hydrogen bonding or hydrophobic interactions with dihydrofolate or one of the two inhibitors were tested. Nine mutants were obtained by site-directed mutagenesis and two were spontaneous mutants previously obtained by in vivo selection (Grange, T., Kunst, F., Thillet, J., Ribadeau-Dumas, B., Mousseron, S., Hung, A., Jami, J., and Pictet, R. (1984) Nucleic Acids Res. 12, 3585-3601). The choice of the mutated positions was based on the knowledge of the active site of chicken dihydrofolate reductase established by x-ray crystallographic studies since the sequences of all known eucaryotic dihydrofolate reductases are greatly conserved. Enzymes were produced in great amounts and purified using a plasmid expressing the mouse cDNA into a dihydrofolate reductase-deficient Escherichia coli strain. The functional properties of recombinant mouse dihydrofolate reductase purified from bacterial extracts were identical to those of dihydrofolate reductase isolated from eucaryotic cells. The Km(NADPH) values for all the mutants except one (Leu-22----Arg) were only slightly modified, suggesting that the mutations had only minor effects on the ternary conformation of the enzyme. In contrast, all Km(H2folate) values were increased, since the mutations were located in the dihydrofolate binding site. The catalytic activity was also modified for five mutants with, respectively, a 6-, 10-, 36-, and 60-fold decrease of Vmax for Phe-31----Arg, Ile-7----Ser, Trp 24----Arg and Leu-22----Arg mutants and a 2-fold increase for Val-115----Pro. All the mutations affected the binding of methotrexate and six, the binding of trimethoprim: Ile-7----Ser, Leu-22----Arg, Trp-24----Arg, Phe-31----Arg, Gln-35----Pro and Phe-34----Leu. The relative variation of Ki for methotrexate and trimethoprim were not comparable from one mutant to the next, reflecting the different binding modes of the two inhibitors. The mutations which yielded the greatest increases in Ki are those which involved amino acids making hydrophobic contacts with the inhibitor.  相似文献   

11.
Serine 127 of human NADH-cytochrome b5 reductase was replaced by proline and alanine by site-directed mutagenesis. The former mutation has been found in the genes of patients with hereditary deficiency of the enzyme. Both the mutant enzymes (Ser-127----Pro mutant and Ser-127----Ala mutant) were overproduced in Escherichia coli and purified to homogeneity. The two purified mutant enzymes showed indistinguishable spectral properties which differed from those of the wild-type enzyme. The mutant enzymes showed higher molecular extinction coefficients at 462 nm than that of the wild-type enzyme. Quenching of FAD fluorescence in these mutant enzymes was significantly less than that in the wild-type enzyme. Furthermore, circular dichroism spectra of the mutant enzymes were different, in both the visible and ultraviolet regions, from that of the wild-type enzyme. The spectra of the mutant enzymes in the visible region were restored to almost the same spectrum as the wild type upon reduction with NADH. Ser-127----Pro mutant and Ser-127----Ala mutant showed very low Kcat/Km (NADH) values (5 x 10(7) and 3.5 x 10(7) s-1 M-1, respectively) with cytochrome b5 as an electron acceptor, than that of the wild-type enzyme (Kcat/Km (NADH) = 179 x 10(7) s-1 M-1), while the Kcat/Km (cytochrome b5) value for each enzyme was similar. The mutant enzymes were less thermostable than the wild-type enzyme. These results indicate that serine 127 plays an important role to maintain the structure of the NADH-binding site in the enzyme.  相似文献   

12.
Growth of a Dihydrofolate Reductaseless Mutant of Bacteriophage T4   总被引:5,自引:5,他引:0       下载免费PDF全文
A mutant of bacteriophage T4 was isolated which was unable to induce virus-specific dihydrofolate reductase in infected cells. The mutant was able to form several other early enzymes of pyrimidine metabolism. Growth of the mutant in a wild-type host, Escherichia coli B, was compared with that of the parent strain, T4BO(1), and T4td8, a mutant which lacks the ability to induce thymidylate synthetase. Growth studies were carried out in minimal medium, which gave higher growth rates and phage yields than the supplemented media used in previous studies. The reductase mutant formed deoxyribonucleic acid and plaque-forming particles at a rate slightly higher than the synthetase mutant but 1.5-to 2-fold lower than that of the wild-type phage under all conditions studied. The addition of thymine to a culture infected by the mutant increased the growth rate significantly, suggesting that the genetic lesion leads to a partial thymidylate deficiency. Like other viral genes controlling steps in thymidylate metabolism, the dihydrofolate reductase gene appears to be useful but not completely essential for growth.  相似文献   

13.
Pediococcus cerevisiae/AMr, resistant to amethopterin, possesses a higher dihydrofolate reductase (5, 6, 7, 8-tetrahydrofolate: NADP+ oxidoreductase, EC 1.5.1.3) activity than the parent, a folate-permeable and thus amethopterin-susceptible strain and than the wild-type. The properties of dihydrofolate reductase from the three strains have been compared. Temperature, pH optima, heat stability, as well amethopterin binding did not reveal significant differences between the enzymes from the susceptible and resistant strains. The enzyme from the wild-type was 10 times more sensitive to inhibition by amethopterin and more susceptible to heat denaturation. The apparent Km values for dihydrofolate in enzymes from the three strains were in the range of 4.8–7.2 μM and for NADPH 6.5–8.0 μM. The amethopterin-resistant strain exhibited cross-resistance to trimethoprim and was about 40-fold more resistant to the latter than the sensitive parent and the wild-type. The resistance to trimethoprim appears to be a direct result of the increased dihydrofolate reductase activity. Inhibition of dihydrofolate reductase activity by this drug was similar in the three strains. 10–20 nmol caused 50% inhibition of 0.02 enzyme unit. Trimethoprim was about 10 000 times less effective inhibitor of dihydrofolate reductase than amethopterin. The cell extract of the AMr strain possessed a folate reductase activity three times higher than that of the sensitive strain. The activities of other folate-related enzymes like thymidylate synthethase and 10-formyltetra-hydrofolate synthetase (formate: tetrahydrofolate ligase (ADP)-forming), EC 6.3.4.3) were similar in the three strains studied.  相似文献   

14.
We have explored the substrate protonation mechanism of Escherichia coli dihydrofolate reductase by changing the location of the proton donor. A double mutant was constructed in which the proton donor of the wild-type enzyme, aspartic acid-27, has been changed to serine and simultaneously an alternative proton donor, glutamic acid, has replaced threonine at position 113. The active site of the resulting variant enzyme molecule should therefore somewhat resemble that proposed for the R67 plasmid-encoded dihydrofolate reductase [Matthews, D. A., Smith, S. L., Baccanari, D. P., Burchall, J. J., Oatley, S. J., & Kraut, J. (1986) Biochemistry 25, 4194]. At pH 7, the double-mutant enzyme has a 3-fold greater kcat and an unchanged Km(dihydrofolate) as compared with the single-mutant Asp-27----Ser enzyme described previously [Howell, E. E., Villafranca, J. E., Warren, M. S., Oatley, S. J., & Kraut, J. (1986) Science (Washington, D.C.) 231, 1123]. Additionally, its activity vs pH profiles together with observed deuterium isotope effects, suggest that catalysis depends on an acidic group with a pKa of 8. It is concluded that the dihydropteridine ring of a bound substrate molecule can indeed be protonated by a glutamic acid side chain at position 113 (instead of an aspartic acid side chain at position 27), but with greatly decreased efficiency: at pH 7, the double mutant still has a 25-fold lower kcat (1.2 s-1) and a 2900-fold lower kcat/km(dihydrofolate) (8.6 X 10(3) s-1 M-1) than the wild-type enzyme.  相似文献   

15.
Human aldose reductase (EC 1.1.1.21) and aldehyde reductase (EC 1.1.1.2) are implicated in the development of diabetic complications by a variety of mechanisms, and a number of drugs to inhibit these enzymes have been proposed for the therapy and prevention of these complications. To probe the structure and function of these two enzymes, we used site-directed mutagenesis in the cDNAs of both enzymes to replace lysine 262 with methionine. Wild-type and mutant enzymes were overexpressed in Escherichia coli and purified by anion exchange and affinity chromatography. N-terminal sequence analysis, Western blots, and kinetic studies confirmed the identity of the recombinant wild-type enzymes with the native human placental and liver enzymes. Recombinant aldose reductase (hAR) and aldehyde reductase (hGR) have apparent kinetic constants virtually identical to their respective native enzymes. The mutant aldose reductase (hARK262 greater than M) shows a 66-fold increase in Km for NADPH with respect to the wild type (1.9 +/- 0.4 microM versus 125 +/- 14 microM), whereas the Km for DL-glyceraldehyde increased 35-fold (20 +/- 2 versus 693 +/- 41 microM). The same constants for the mutant aldehyde reductase (hGRK262 greater than M) increased 97- and 86-fold, respectively (from 2.0 +/- 0.4 to 194 +/- 16 microM and from 1.6 +/- 0.4 to 137 +/- 3 mM). These results indicate that lysine 262 in aldose reductase and aldehyde reductase is crucial to their catalytic activity by affecting co-factor binding.  相似文献   

16.
Characterization of Candida albicans dihydrofolate reductase   总被引:3,自引:0,他引:3  
Dihydrofolate reductase from Candida albicans was purified 31,000-fold and characterized. In addition, the C. albicans dihydrofolate reductase gene was cloned into a plasmid vector and expressed in Escherichia coli, and the enzyme was purified from this source. Both preparations showed a single protein-staining band with a molecular weight of about 25,000 on sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzymes were stable and had an isoelectric point of pH 7.1 on gel isoelectric focusing. Kinetic characterization showed that the enzymes from each source had similar turnover numbers (about 11,000 min-1) and Km values for NADPH and dihydrofolate of 3-4 microM. Like other eukaryotic dihydrofolate reductases, the C. albicans enzyme exhibited weak binding affinity for the antibacterial agent trimethoprim (Ki = 4 microM), but further characterization showed that the inhibitor binding profile of the yeast and mammalian enzymes differed. Methotrexate was a tight binding inhibitor of human but not C. albicans dihydrofolate reductase; the latter had a relatively high methotrexate Ki of 150 pM. The yeast and vertebrate enzymes also differed in their interactions with KCl and urea. These two agents activate vertebrate dihydrofolate reductases but inhibited the C. albicans enzyme. The sequence of the first 36 amino-terminal amino acids of the yeast enzyme was also determined. This portion of the C. albicans enzyme was more similar to human than to E. coli dihydrofolate reductases (50% and 30% identity, respectively). Some key amino acid residues in the C. albicans sequence, such as E-30 (human enzyme numbering), were "vertebrate-like" whereas others, such as I-31, were not. These results indicate that there are physical and kinetic differences between the eukaryotic mammalian and yeast enzymes.  相似文献   

17.
The structure of pentaerythritol tetranitrate (PETN) reductase in complex with the nitroaromatic substrate picric acid determined previously at 1.55 A resolution indicated additional electron density between the indole ring of residue Trp-102 and the nitro group at C-6 of picrate. The data suggested the presence of an unusual bond between substrate and the tryptophan side chain. Herein, we have extended the resolution of the PETN reductase-picric acid complex to 0.9 A. This high-resolution analysis indicates that the active site is partially occupied with picric acid and that the anomalous density seen in the original study is attributed to the population of multiple conformational states of Trp-102 and not a formal covalent bond between the indole ring of Trp-102 and picric acid. The significance of any interaction between Trp-102 and nitroaromatic substrates was probed further in solution and crystal complexes with wild-type and mutant (W102Y and W102F) enzymes. Unlike with wild-type enzyme, in the crystalline form picric acid was bound at full occupancy in the mutant enzymes, and there was no evidence for multiple conformations of active site residues. Solution studies indicate tighter binding of picric acid in the active sites of the W102Y and W102F enzymes. Mutation of Trp-102 does not impair significantly enzyme reduction by NADPH, but the kinetics of decay of the hydride-Meisenheimer complex are accelerated in the mutant enzymes. The data reveal that decay of the hydride-Meisenheimer complex is enzyme catalyzed and that the final distribution of reaction products for the mutant enzymes is substantially different from wild-type enzyme. Implications for the mechanism of high explosive degradation by PETN reductase are discussed.  相似文献   

18.
Single tryptophan-containing mutants of low adenylylation state Escherichia coli glutamine synthetase have been studied by frequency-domain fluorescence spectroscopy in the presence of various substrates and inhibitors. At pH 6.5, the Mn-bound wild-type enzyme (wild type has two tryptophans/subunit) and the mutant enzymes exhibit heterogeneous fluorescence decay kinetics; the individual tryptophans are adequately described by a triple exponential decay scheme. The recovered lifetime values are 5.9 ns, 2.6 ns, and 0.4 ns for Trp-57 and 5.8 ns, 2.3 ns, and 0.4 ns for Trp-158. These values are nearly identical to the previously reported results at pH 7.5 (Atkins, W.M., Stayton, P.S., & Villafranca, J.J., 1991, Biochemistry 30, 3406-3416). In addition, Trp-57 and Trp-158 both exhibit an ATP-induced increase in the relative fraction of the long lifetime component, whereas only Trp-57 is affected by this ligand at pH 7.5. The transition-state analogue L-methionine-(R,S)-sulfoximine (MSOX) causes a dramatic increase in the fractional intensity of the long lifetime component of Trp-158. This ligand has no effect on the W158S mutant protein and causes a small increase in the fractional intensity of the long lifetime component of the W158F mutant protein. Addition of glutamate to the ATP complex, which affords the gamma-glutamylphosphate-ADP complex, results in the presence of new lifetime components at 7, 3.2, and 0.5 ns for Trp-158, but has no effect on Trp-57. Similar results were obtained when ATP was added to the MSOX complex; Trp-57 exhibits heterogeneous fluorescence decay with lifetimes of 7, 3.5, and 0.8 ns. Decay kinetics of Trp-158 are best fit to a nearly homogeneous decay with a lifetime of 5.5 ns in the MSOX-ATP inactivated complex. These results provide a model for the sequence of structural and dynamic changes that take place at the Trp-57 loop and the central loop (Trp-158) during several intermediate stages of catalysis.  相似文献   

19.
Site-directed mutagenesis was employed to investigate the role of Cys566 in the catalytic mechanism of rat liver NADPH-cytochrome P-450 oxidoreductase. Rat NADPH-cytochrome P-450 oxidoreductase and mutants containing either alanine or serine at position 566 were expressed in Escherichia coli and purified to homogeneity. Substitution of alanine at position 566 had no effect on enzymatic activity with the acceptors cytochrome c and ferricyanide but did increase trans-hydrogenase activity with 3-acetylpyridine adenine dinucleotide phosphate by 79%. The Km for NADPH was increased 2.5-fold, and the NADP+ KI was increased 4.8-fold compared with that found for the wild-type enzyme. The conservative substitution, Ser566, produced a 50% decrease in cytochrome c reductase activity whereas activity with ferricyanide was decreased 57%, and 3-acetylpyridine adenine dinucleotide phosphate activity was unaffected. The NADPH Km was increased 4.6-fold, and the NADP+ KI increased 7.6-fold. The dependence of cytochrome c reductase activity on the KCl concentration was markedly altered by the Cys566 substitutions. Maximum activity for the wild-type enzyme was observed at approximately 0.18 M KCl whereas maximum activity for the mutant enzymes was observed between 0.04 and 0.09 M KCl. The pH dependence of cytochrome c reductase activity, cytochrome c Km, and flavin content were unaffected by these substitutions. These results demonstrate that Cys566 is not essential for activity of rat liver NADPH-cytochrome P-450 oxidoreductase although the cysteine side chain does affect the interaction of NADPH with the enzyme.  相似文献   

20.
Proline 17 in the glycine-rich region of adenylate kinase was replaced by Gly (the Gly-mutant) or Val (the Val-mutant) by site-directed mutagenesis. The mutant enzymes were purified to homogeneous states on sodium dodecyl sulfate-gel electrophoresis after solubilization of the proteins from the pellets of cell lysates of Escherichia coli. The apparent Km values of the Gly- and the Val-mutants for AMP increased approximately 7- and 24-fold, respectively, as compared with that of the wild-type enzyme. The apparent Km values for ATP also increased 7- and 42-fold in the Gly- and Val-mutants, respectively. In contrast, Vmax values of both mutant enzymes were comparable to that of the wild-type enzyme. These results suggest that Pro-17 plays an important role for the binding of substrates, but not for catalytic efficiency, although it does not directly interact with substrates. Adenosine diphosphopyridoxal, which specifically modifies Lys-21 in adenylate kinase (Tagaya, M., Yagami, T., and Fukui, T. (1987) J. Biol. Chem. 262, 8257-8261), inactivated the wild-type and mutant enzymes at almost the same rates. Interestingly, both mutant enzymes showed higher specificities for adenine nucleotides than the wild-type enzyme. Both mutant enzymes were less resistant than the wild-type enzyme against inactivation at elevated temperatures or by treatment with trypsin. It would appear that most of the properties of the mutant enzymes may be explained on the basis of a need for conformational flexibility of the loop which includes Pro-17 for substrate binding.  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