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 734 毫秒
1.
The reaction of pentaerythritol tetranitrate reductase with reducing and oxidizing substrates has been studied by stopped-flow spectrophotometry, redox potentiometry, and X-ray crystallography. We show in the reductive half-reaction of pentaerythritol tetranitrate (PETN) reductase that NADPH binds to form an enzyme-NADPH charge transfer intermediate prior to hydride transfer from the nicotinamide coenzyme to FMN. In the oxidative half-reaction, the two-electron-reduced enzyme reacts with several substrates including nitroester explosives (glycerol trinitrate and PETN), nitroaromatic explosives (trinitrotoluene (TNT) and picric acid), and alpha,beta-unsaturated carbonyl compounds (2-cyclohexenone). Oxidation of the flavin by the nitroaromatic substrate TNT is kinetically indistinguishable from formation of its hydride-Meisenheimer complex, consistent with a mechanism involving direct nucleophilic attack by hydride from the flavin N5 atom at the electron-deficient aromatic nucleus of the substrate. The crystal structures of complexes of the oxidized enzyme bound to picric acid and TNT are consistent with direct hydride transfer from the reduced flavin to nitroaromatic substrates. The mode of binding the inhibitor 2,4-dinitrophenol (2,4-DNP) is similar to that observed with picric acid and TNT. In this position, however, the aromatic nucleus is not activated for hydride transfer from the flavin N5 atom, thus accounting for the lack of reactivity with 2,4-DNP. Our work with PETN reductase establishes further a close relationship to the Old Yellow Enzyme family of proteins but at the same time highlights important differences compared with the reactivity of Old Yellow Enzyme. Our studies provide a structural and mechanistic rationale for the ability of PETN reductase to react with the nitroaromatic explosive compounds TNT and picric acid and for the inhibition of enzyme activity with 2,4-DNP.  相似文献   

2.
The roles of His181, His184 and Tyr186 in PETN reductase have been examined by mutagenesis, spectroscopic and stopped-flow kinetics, and by determination of crystallographic structures for the Y186F PETN reductase and reduced wild-type enzyme-progesterone complex. Residues His181 and His184 are important in the binding of coenzyme, steroids, nitroaromatic ligands and the substrate 2-cyclohexen-1-one. The H181A and H184A enzymes retain activity in reductive and oxidative half-reactions, and thus do not play an essential role in catalysis. Ligand binding and catalysis is not substantially impaired in Y186F PETN reductase, which contrasts with data for the equivalent mutation (Y196F) in Old Yellow Enzyme. The structure of Y186F PETN reductase is identical to wild-type enzyme, with the obvious exception of the mutation. We show in PETN reductase that Tyr186 is not a key proton donor in the reduction of alpha/beta unsaturated carbonyl compounds. The structure of two electron-reduced PETN reductase bound to the inhibitor progesterone mimics the catalytic enzyme-steroid substrate complex and is similar to the structure of the oxidized enzyme-inhibitor complex. The reactive C1-C2 unsaturated bond of the steroid is inappropriately orientated with the flavin N5 atom for hydride transfer. With steroid substrates, the productive conformation is achieved by orientating the steroid through flipping by 180 degrees , consistent with known geometries for hydride transfer in flavoenzymes. Our data highlight mechanistic differences between Old Yellow Enzyme and PETN reductase and indicate that catalysis requires a metastable enzyme-steroid complex and not the most stable complex observed in crystallographic studies.  相似文献   

3.
One of the three tryptophan residues per subunit of thermostable D-amino acid transaminase, Trp-139, is close to the active-site Lys-145 in the sequence of the protein. This tryptophan has been changed to several other types of residues by site-directed mutagenesis. The only mutant protein that was sufficiently active and stable for study had Phe substituted for Trp (W139F). The spectroscopic properties of this mutant enzyme differed from those of the wild-type transaminase. For example, denatured W139F showed the expected decrease in fluorescence emission intensity at 350 nm due to the deletion of one Trp residue, but the fluorescence emission of the wild-type and W139F enzymes in the native state did not differ in intensity. This result suggests that the fluorescence of Trp-139 in the native, wild-type enzyme is not manifested perhaps due to its proximity to the coenzyme, pyridoxal phosphate. Results of energy-transfer studies at several wavelengths could also be interpreted as due to the proximity of Trp-139 and the coenzyme. Circular dichroism studies indicated that the negative Cotton effect at 420 nm due to the coenzyme was still present in W139F. However, the 280-nm optically active band present in the wild-type enzyme was greatly diminished in W139F. The mutant protein with Asp at position 139 (W139D) could not be isolated presumably because it was degraded. The other mutant enzymes, W139P, W139A, and W139H, were isolated with partial activities (15-35%) that were slowly lost upon storage at 4 degrees C. Overall, these results indicate the importance of Trp-139 in the thermostable D-amino acid transaminase.  相似文献   

4.
The kinetics of flavin reduction in two mutant forms of human cytochrome P450 reductase have been studied by stopped-flow spectroscopy with absorption and fluorescence detection. The mutant enzymes were altered at the position of Trp-676, which, by analogy with the structure of rat CPR, is close to the isoalloxazine ring of the enzyme-bound FAD. We show that mutant CPRs in which Trp-676 has been changed to histidine (W676H) and alanine (W676A) can be reduced by NADPH only to the two-electron level in single mixing stopped-flow experiments. The concentration dependence of the rate of hydride transfer indicates that the second, noncatalytic NADPH-binding site present in wild-type CPR is retained in the mutant enzymes. Detailed studies of W676H CPR indicate that further reduction of the enzyme beyond the two electron level is prevented due to the slow release of NADP(+) from the active site following the first hydride transfer from NADPH, owing to the stability of a reduced enzyme-NADP(+) charge-transfer complex. Reduction to the four-electron level is achieved in a sequential mixing stopped-flow experiment. In this procedure, W676H CPR is reacted first with a stoichiometric amount of NADPH, and then, following a delay of 100 ms, with excess NADPH. The data indicate that occupancy of the noncatalytic coenzyme site also hinders NADP(+) release from reduced enzyme. Fluorescence stopped-flow studies of the W676H and wild-type CPR enzymes reveal that the complex signals associated with reduction of wild-type CPR by NADPH are attributable to changes in the environment of residue W676. From these studies, a model is proposed for nicotinamide binding in wild-type CPR. In this model W676 serves as a trigger to release NADP(+) from the active site following hydride transfer. In the W676H enzyme, the slow release of NADP(+) is a consequence of the combined effects of (i) removing W676 by mutagenesis (thus removing the trigger for displacement) and (ii) the binding of NADPH in the noncatalytic site, thus trapping NADP(+) in the catalytic site.  相似文献   

5.
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   

6.
Enterobacter cloacae PB2 was originally isolated on the basis of its ability to utilize nitrate esters, such as pentaerythritol tetranitrate (PETN) and glycerol trinitrate, as the sole nitrogen source for growth. The enzyme responsible is an NADPH-dependent reductase designated PETN reductase. E. cloacae PB2 was found to be capable of slow aerobic growth with 2,4,6-trinitrotoluene (TNT) as the sole nitrogen source. Dinitrotoluenes were not produced and could not be used as nitrogen sources. Purified PETN reductase was found to reduce TNT to its hydride-Meisenheimer complex, which was further reduced to the dihydride-Meisenheimer complex. Purified PETN reductase and recombinant Escherichia coli expressing PETN reductase were able to liberate nitrogen as nitrite from TNT. The ability to remove nitrogen from TNT suggests that PB2 or recombinant organisms expressing PETN reductase may be useful for bioremediation of TNT-contaminated soil and water.  相似文献   

7.
Mutants of the active site residues Trp-116 and Tyr-114 of the molybdenum-containing Me(2)SO reductase from Rhodobacter capsulatus have been examined spectroscopically and kinetically. The Y114F mutant has an increased rate constant for oxygen atom transfer from Me(2)SO to reduced enzyme, the result of lower stability of the E(red).Me(2)SO complex. The absorption spectrum of this species (but not that of either oxidized or reduced enzyme) is significantly perturbed in the mutant relative to wild-type enzyme, consistent with Tyr-114 interacting with bound Me(2)SO. The as-isolated W116F mutant is only five-coordinate, with one of the two equivalents of the pyranopterin cofactor found in the enzyme dissociated from the molybdenum and replaced by a second Mo=O group. Reduction of the mutant with sodium dithionite and reoxidation with Me(2)SO, however, regenerates the long-wavelength absorbance of functional enzyme, although the wavelength maximum is shifted to 670 nm from the 720 nm of wild-type enzyme. This "redox-cycled" mutant exhibits a Me(2)SO reducing activity and overall reaction mechanism similar to that of wild-type enzyme but rapidly reverts to the inactive five-coordinate form in the course of turnover.  相似文献   

8.
Adenylylation of Tyr-397 of each subunit of Escherichia coli glutamine synthetase (GS) down-regulates enzymatic activity in vivo. The overall structure of the enzyme consists of 12 subunits arranged as two hexamers, face to face. Research reported in this paper addresses the question of whether the covalently attached adenylyl group interacts with neighboring amino acid residues to produce the regulatory phenomenon. Wild-type GS has two Trp residues (positions 57 and 158) and the adenylylation site lies within 7-8 A of the Trp-57 loop in the adjacent subunit of the same hexameric ring; Trp-158 is about 35 A from the site of adenylylation. Fluorescence lifetimes and quantum yields have been determined for two fluorophores with wild-type and mutant GS. One fluorophore is epsilon-AMP adenylylated GS (at Tyr-397), and the other fluorophore is the intrinsic protein residue Trp-57. These experiments were conducted in order to detect possible intersubunit interactions between adenylyl groups and the neighboring Trp-57 to search for a role for the Trp-57 loop in the regulation of GS. The fluorescence due to epsilon-AMP of two adenylylated enzymes, wild-type GS and the W158F mutant, exhibits heterogeneous decay kinetics; the data adequately fit to a double exponential decay model with recovered average lifetime values of 18.2 and 2.1 ns, respectively. The pre-exponential factors range from 0.66 to 0.73 for the long lifetime component, at five emission wavelengths. The W57L-epsilon-AMP enzyme yields longer average lifetime values of 19.5 and 2.4 ns, and the pre-exponential factors range from 0.82 to 0.85 for the long lifetime component. An additional residue in the Trp-57 loop, Lys-58, has been altered and the K58C mutant enzyme has been adenylylated with epsilon-AMP on Tyr-397. Lys-58 is near the ATP binding site and may represent a link by which the adenylyl group controls the activity of GS. The fluorescence of epsilon-AMP-adenylylated K58C mutant GS is best described by a triple exponential decay with average recovered lifetime values of 19.9, 4.6, and 0.58 ns, with the largest fraction being the median lifetime component. Relative quantum yields of epsilon-AMP-Tyr-397 were measured in order to determine if static quenching occurs from adenine-indole stacking in the wild-type GS. The relative quantum yield of the epsilon-AMP-adenylylated W57L mutant is larger than the wild-type protein by the amount predicted from the difference in lifetime values: thus, no static quenching is evident.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
NADPH-cytochrome P450 oxidoreductase catalyzes transfer of electrons from NADPH, via two flavin cofactors, to various cytochrome P450s. The crystal structure of the rat reductase complexed with NADP(+) has revealed that nicotinamide access to FAD is blocked by an aromatic residue (Trp-677), which stacks against the re-face of the isoalloxazine ring of the flavin. To investigate the nature of interactions between the nicotinamide, FAD, and Trp-677 during the catalytic cycle, three mutant proteins were studied by crystallography. The first mutant, W677X, has the last two C-terminal residues, Trp-677 and Ser-678, removed; the second mutant, W677G, retains the C-terminal serine residue. The third mutant has the following three catalytic residues substituted: S457A, C630A, and D675N. In the W677X and W677G structures, the nicotinamide moiety of NADP(+) lies against the FAD isoalloxazine ring with a tilt of approximately 30 degrees between the planes of the two rings. These results, together with the S457A/C630A/D675N structure, allow us to propose a mechanism for hydride transfer regulated by changes in hydrogen bonding and pi-pi interactions between the isoalloxazine ring and either the nicotinamide ring or Trp-677 indole ring. Superimposition of the mutant and wild-type structures shows significant mobility between the two flavin domains of the enzyme. This, together with the high degree of disorder observed in the FMN domain of all three mutant structures, suggests that conformational changes occur during catalysis.  相似文献   

10.
The crystal structure of the NADH-dependent bacterial flavoenzyme morphinone reductase (MR) has been determined at 2.2-A resolution in complex with the oxidizing substrate codeinone. The structure reveals a dimeric enzyme comprising two 8-fold beta/alpha barrel domains, each bound to FMN, and a subunit folding topology and mode of flavin-binding similar to that found in Old Yellow Enzyme (OYE) and pentaerythritol tetranitrate (PETN) reductase. The subunit interface of MR is formed by interactions from an N-terminal beta strand and helices 2 and 8 of the barrel domain and is different to that seen in OYE. The active site structures of MR, OYE, and PETN reductase are highly conserved reflecting the ability of these enzymes to catalyze "generic" reactions such as the reduction of 2-cyclohexenone. A region of polypeptide presumed to define the reducing coenzyme specificity is identified by comparison of the MR structure (NADH-dependent) with that of PETN reductase (NADPH-dependent). The active site acid identified in OYE (Tyr-196) and conserved in PETN reductase (Tyr-186) is replaced by Cys-191 in MR. Mutagenesis studies have established that Cys-191 does not act as a crucial acid in the mechanism of reduction of the olefinic bond found in 2-cyclohexenone and codeinone.  相似文献   

11.
Rhodococcus erythropolis HL 24-2, which was originally isolated as a 2,4-dinitrophenol-degrading bacterium, could also utilize picric acid as a nitrogen source after spontaneous mutation. During growth, the mutant HL PM-1 transiently accumulated an orange-red metabolite, which was identified as a hydride-Meisenheimer complex of picric acid. This complex was formed as the initial metabolite and further converted with concomitant liberation of nitrite. 2,4,6-Trinitrocyclohexanone was identified as a dead-end metabolite of the degradation of picric acid, indicating the addition of two hydride ions to picric acid.  相似文献   

12.
Two homologous Delta5-3-ketosteroid isomerases from Comamonas testosteroni (TI-WT) and Pseudomonas putida biotype B (PI-WT) exhibit different pH activity profiles. TI-WT loses activity below pH 5.0 due to the protonation of the conserved catalytic base, Asp-38, while PI-WT does not. Based on the structural analysis of PI-WT, the critical catalytic base, Asp-38, was found to form a hydrogen bond with the indole ring NH of Trp-116, which is homologously replaced with Phe-116 in TI-WT. To investigate the role of Trp-116, we prepared the F116W mutant of TI-WT (TI-F116W) and the W116F mutant of PI-WT (PI-W116F) and compared kinetic parameters of those mutants at different pH levels. PI-W116F exhibited significantly decreased catalytic activity at acidic pH like TI-WT, whereas TI-F116W maintained catalytic activity at acidic pH like PI-WT and increased the kcat/Km value by 2.5- to 4.7-fold compared with TI-WT at pH 3.8. The crystal structure of TI-F116W clearly showed that the indole ring NH of Trp-116 could form a hydrogen bond with the carboxyl oxygen of Asp-38 like that of PI-WT. The present results demonstrate that the activities of both PI-WT and TI-F116W at low pH were maintained by a tryptophan, which was able not only to lower the pKa value of the catalytic base but also to increase the substrate affinity. This is one example of the strategy nature can adopt to evolve the diversity of the catalytic function in the enzymes. Our results provide insight into deciphering the molecular evolution of the enzyme and creating novel enzymes by protein engineering.  相似文献   

13.
Rhodococcus erythropolis HL 24-2, which was originally isolated as a 2,4-dinitrophenol-degrading bacterium, could also utilize picric acid as a nitrogen source after spontaneous mutation. During growth, the mutant HL PM-1 transiently accumulated an orange-red metabolite, which was identified as a hydride-Meisenheimer complex of picric acid. This complex was formed as the initial metabolite and further converted with concomitant liberation of nitrite. 2,4,6-Trinitrocyclohexanone was identified as a dead-end metabolite of the degradation of picric acid, indicating the addition of two hydride ions to picric acid.  相似文献   

14.
The binding site residue Trp-24 is conserved in all vertebrate and bacterial dihydrofolate reductases of known sequence. To determine its effects on enzyme properties, a Trp-24 to Phe-24 mutant (W-24-F) of human dihydrofolate reductase has been constructed by oligodeoxynucleotide site-directed mutagenesis. The W-24-F mutant enzyme appears to have a more open or flexible conformation as compared to the wild-type human dihydrofolate reductase on the basis of results of a number of studies. These studies include competitive ELISA using peptide-specific antibodies against human dihydrofolate reductase, thermal stability, and protease susceptibility studies of both mutant W-24-F and wild-type enzymes. It is concluded that Trp-24 is important for maintaining the structural integrity of the native enzymes. Changes in relative fluorescence quantum yield indicate that Trp-24 is buried and its fluorescence quenched relative to the other two tryptophan residues in the wild-type human reductase. Kinetic studies indicate that kcat values for W-24-F are increased in the pH range of 4.5-8.5 with a 5-fold increase at pH 7.5 as compared to the wild-type enzyme. However, the catalytic efficiency of W-24-F decreases rapidly as the pH is increased from 7.5 to 9.5. The Km values for dihydrofolate are also increased for W-24-F in the pH range of 4.5-9.5 with a 30-fold increase at pH 7.5, while the Km value for NADPH increases only ca. 1.4-fold at pH 7.5 as compared to the wild type.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
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   

16.
M Kubo  Y Mitsuda  M Takagi    T Imanaka 《Applied microbiology》1992,58(11):3779-3783
On the basis of three-dimensional information, many amino acid substitutions were introduced in the thermostable neutral protease (NprM) of Bacillus stearothermophilus MK232 by site-directed mutagenesis. When Glu at position 143 (Glu-143), which is one of the proposed active sites, was substituted for by Gln and Asp, the proteolytic activity disappeared. F114A (Phe-114 to Ala), Y110W (Tyr-110 to Trp), and Y211W (Tyr-211 to Trp) mutant enzymes had higher activity (1.3- to 1.6-fold) than the wild-type enzyme. When an autolysis site, Tyr-93, was replaced by Gly and Ser, the remaining activities of those mutant enzymes were higher than that of the wild-type enzyme.  相似文献   

17.
Single-tryptophan-containing mutants of low adenylation state Escherichia coli glutamine synthetase (wild type has two tryptophans at positions 57 and 158) have been constructed and studied by multifrequency phase/modulation fluorescence spectroscopy. The W57L mutant (retains tryptophan at residue 158) and the W158S mutant (retains tryptophan at residue 57) are both characterized by heterogeneous exponential decay kinetics. Global analysis indicates that for the Mn-bound form of the enzyme at pH 7.4 the fluorescence of both tryptophans is best described by a sum of three discrete expontials with recovered lifetimes of 4.77, 1.72, and 0.10 ns for Trp-57 and 5.04, 2.28, and 0.13 ns for Trp-158. The wild-type enzyme also exhibits decay kinetics described by a triple-exponential model with similar lifetime components. The individual tryptophans are distinguishable by the fractional intensities of the resolvable lifetimes. The wild-type and W158S enzymes are dominated by the 5-ns component which provides nearly 60% and 65%, respectively, of the fractional intensity at five wavelengths spanning the emission spectrum. In contrast, the W57L enzyme demonstrates a larger fraction of the 2-ns lifetime species (60%) and only 35% of the longer lifetime component. The substrate ATP induces a shift to approximately 90% of the 5-ns component for the wild-type and W158S enzymes, whereas the W57L protein is essentially unaffected by this ligand. Steady-state quenching studies with iodide indicate that addition of ATP results in a 3.0-3.5-fold decrease in the apparent Stern-Volmer quenching constants for the wild-type and W158S enzymes. Phase/modulation experiments at several iodide concentrations indicate that the median, 2 ns, lifetime component is selectively quenched compared to the 5-ns lifetime component. These results suggest a model where ATP binding results in a shift in the equilibrium distribution of microconformational states populated by Trp-57. ATP shifts this equilibrium nearly completely to the states exhibiting the long-lifetime component which, based on quenching studies, is less solvent-accessible than the conformational states associated with the other lifetime components.  相似文献   

18.
Sirtuins are emerging as the key regulators of metabolism and aging, and their potential activators and inhibitors are being explored as therapeutics for improving health and treating associated diseases. Despite the global structural similarity among all seven isoforms of sirtuins (of which most of them catalyze the deacetylation reaction), SIRT5 is the only isoform that catalyzes the cleavage of negatively charged acylated substrates, and the latter feature appears to be encoded by the presence of Tyr102 and Arg105 residues at the active site pocket of the enzyme. To determine the contributions of the above residues in SIRT5 (vis a vis the corresponding residues of SIRT1) on substrate selectivity, inhibition by EX527 and nicotinamide, secondary structural features and thermal stability of the enzymes, we created single and double mutations (viz. Y102A, R105l, and Y102A/R105I) in SIRT5. The kinetic data revealed that while Y102A mutant enzyme catalyzed both deacetylation and desuccinylation reactions with comparable efficiencies, R105I and Y102A/R105I mutant enzymes favored the deacetylase reaction. Like SIRT1, the nicotinamide inhibition of SIRT5 double mutant (Y102A/R105I) exhibited the mixed non-competitive behavior. On the other hand, the desuccinylation reaction of both wild-type and Y102A mutant enzymes conformed to the competitive inhibition model. The inhibitory potency of EX527 progressively increased from Y102A, R105I, to Y102A/R105 mutant enzymes in SIRT5, but it did not reach to the level obtained with SIRT1. The CD spectroscopic data for the wild-type and mutant enzymes revealed changes in the secondary structural features of the enzymes, and such changes were more pronounced on examining their thermal denaturation patterns. A cumulative account of our experimental data reveal mutual cooperation between Y102 and R105 residues in promoting the desuccinylation versus deacetylation reaction in SIRT5, and the overall catalytic feature of the enzyme is manifested via the mutation induced modulation in the protein structure.  相似文献   

19.
On the basis of three-dimensional information, many amino acid substitutions were introduced in the thermostable neutral protease (NprM) of Bacillus stearothermophilus MK232 by site-directed mutagenesis. When Glu at position 143 (Glu-143), which is one of the proposed active sites, was substituted for by Gln and Asp, the proteolytic activity disappeared. F114A (Phe-114 to Ala), Y110W (Tyr-110 to Trp), and Y211W (Tyr-211 to Trp) mutant enzymes had higher activity (1.3- to 1.6-fold) than the wild-type enzyme. When an autolysis site, Tyr-93, was replaced by Gly and Ser, the remaining activities of those mutant enzymes were higher than that of the wild-type enzyme.  相似文献   

20.
Time-resolved polarized flavin fluorescence was used to study the active site dynamics of Escherichia coli glutathione reductase (GR). Special consideration was given to the role of Tyr177, which blocks the access to the NADPH binding-site in the crystal structure of the enzyme. By comparing wild-type GR with the mutant enzymes Y177F and Y177G, a fluorescence lifetime of 7 ps that accounts for approximately 90% of the fluorescence decay could be attributed to quenching by Y177. Based on the temperature invariance for this lifetime, and the very high quenching rate, electron transfer from Y177 to the light-excited isoalloxazine part of flavin adenine dinucleotide (FAD) is proposed as the mechanism of flavin fluorescence quenching. Contrary to the mutant enzymes, wild-type GR shows a rapid fluorescence depolarization. This depolarization process is likely to originate from a transient charge transfer interaction between Y177 and the light-excited FAD, and not from internal mobility of the flavin, as has previously been proposed. Based on the fluorescence lifetime distributions, the mutants Y177F and Y177G have a more flexible protein structure than wild-type GR: in the range of 223 K to 277 K in 80% glycerol, both tyrosine mutants mimic the closely related enzyme dihydrolipoyl dehydrogenase. The fluorescence intensity decays of the GR enzymes can only be explained by the existence of multiple quenching sites in the protein. Although structural fluctuations are likely to contribute to the nonexponential decay and the probability of quenching by a specific site, the concept of conformational substates need not be invoked to explain the heterogeneous fluorescence dynamics.  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