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.
The thermal dependence of kinetic parameters has been determined in purified or partially purified preparations of cold-hardiness-specific glutathione reductase isozymes from red spruce (Picea rubens Sarg.) needles to investigate a possible functional adaptation of these isozymes to environmental temperature. We have previously purified glutathione reductase isozymes specific for nonhardened (GR-1NH) or hardened (GR-1H) needles. Isozymes that were distinct from GR-1NH and GR-1H, but appeared to be very similar to each other, were also purified from nonhardened (GR-2NH) or hardened (GR-2H) needles (A. Hausladen, R.G. Alscher [1994] Plant Physiol 105: 205-213). GR-1NH had 2-fold higher Km values for NADPH and 2- to 4-fold lower Km values for oxidized glutathione (GSSG) than GR-2NH, and a similar difference was found between GR-1H and GR-2H. However, no differences in Km values were found between the hardiness-specific isozymes GR-1NH and GR-1H. There was only a small effect of temperature on the Km(GSSG) of GR-1H and GR-2H, and no significant temperature effect on Km(NADPH) or Km(GSSG) could be found for the other isozymes. These results are discussed with respect to "thermal kinetic windows," and it is proposed that the relative independence of Km values to temperature ensures adequate enzyme function in a species that is exposed to extreme temperature differences in its natural habitat. A variety of substrates has been tested to characterize any further differences among the isozymes, but all isozymes are highly specific for their substrates, NADPH and GSSG. The reversible reductive inactivation by NADPH (redox interconversion) is more pronounced in GR-1H than in GR-2H. Reduced, partially inactive GR-1H is further deactivated by H2O2, whereas GR-2H is fully reactivated by the same treatment. Both isozymes are reactivated by GSSG or reduced glutathione. It is proposed that this property of GR-2H ensures enzyme function under oxidative conditions, and that in vivo the enzyme may exist in its partially inactive form and be activated in the presence of increased levels of GSSG or oxidants.  相似文献   

2.
Tyrosine 34 is a prominent and conserved residue in the active site of the manganese superoxide dismutases in organisms from bacteria to man. We have prepared the mutant containing the replacement Tyr 34 --> Phe (Y34F) in human manganese superoxide dismutase (hMnSOD) and crystallized it in two different crystal forms, orthorhombic and hexagonal. Crystal structures of hMnSOD Y34F have been solved to 1.9 A resolution in a hexagonal crystal form, denoted as Y34Fhex, and to 2.2 A resolution in an orthorhombic crystal form, denoted as Y34Fortho. Both crystal forms give structures that are closely superimposable with that of wild-type hMnSOD, with the phenyl rings of Tyr 34 in the wild type and Phe 34 in the mutant very similar in orientation. Therefore, in Y34F, a hydrogen-bonded relay that links the metal-bound hydroxyl to ordered solvent (Mn-OH to Gln 143 to Tyr 34 to H2O to His 30) is broken. Surprisingly, the loss of the Tyr 34 hydrogen bonds resulted in large increases in stability (measured by Tm), suggesting that the Tyr 34 hydroxyl does not play a role in stabilizing active-site architecture. The functional role of the side chain hydroxyl of Tyr 34 can be evaluated by comparison of the Y34F mutant with the wild-type hMnSOD. Both wild-type and Y34F had kcat/Km near 10(9) M-1 s-1, close to diffusion-controlled; however, Y34F showed kcat for maximal catalysis smaller by 10-fold than the wild type. In addition, the mutant Y34F was more susceptible to product inhibition by peroxide than the wild-type enzyme. This activity profile and the breaking of the hydrogen-bonding chain at the active site caused by the replacement Tyr 34 --> Phe suggest that Tyr 34 is a proton donor for O2* - reduction to H2O2 or is involved indirectly by orienting solvent or other residues for proton transfer. Up to 100 mM buffers in solution failed to enhance catalysis by either Y34F or the wild-type hMnSOD, suggesting that protonation from solution cannot enhance the release of the inhibiting bound peroxide ion, likely reflecting the enclosure of the active site by conserved residues as shown by the X-ray structures. The increased thermostability of the mutant Y34F and equal diffusion-controlled activity of Y34F and wild-type enzymes with normal superoxide levels suggest that evolutionary conservation of active-site residues in metalloenzymes reflects constraints from extreme rather than average cellular conditions. This new hypothesis that extreme rather than normal substrate concentrations are a powerful constraint on residue conservation may apply most strongly to enzyme defenses where the ability to meet extreme conditions directly affects cell survival.  相似文献   

3.
The enzymatic properties and the three-dimensional structure of spinach glycolate oxidase which has the active-site Tyr129 replaced by Phe (Y129F glycolate oxidase) has been studied. The structure of the mutant is unperturbed which facilitates interpretation of the biochemical data. Y129F glycolate oxidase has an absorbance spectrum with maxima at 364 and 450 nm (epsilon max = 11400 M-1 cm-1). The spectrum indicates that the flavin is in its normal protonated form, i.e. the Y129F mutant does not lower the pKa of the N(3) of oxidized flavin as does the wild-type enzyme [Macheroux, P., Massey, V., Thiele, D. J., and Volokita, M. (1991) Biochemistry 30, 4612-4619]. This was confirmed by a pH titration of Y129F glycolate oxidase which showed that the pKa is above pH 9. In contrast to wild-type glycolate oxidase, oxalate does not perturb the absorbance spectrum of Y129F glycolate oxidase. Moreover oxalate does not inhibit the enzymatic activity of the mutant enzyme. Typical features of wild-type glycolate oxidase that are related to a positively charged lysine side chain near the flavin N(1)-C(2 = O), such as stabilization of the anionic flavin semiquinone and formation of tight N(5)-sulfite adducts, are all conserved in the Y129F mutant protein. Y129F glycolate oxidase exhibited about 3.5% of the wild-type activity. The lower turnover number for the mutant of 0.74 s-1 versus 20 s-1 for the wild-type enzyme amounts to an increase of the energy of the transition state of about 7.8 kJ/mol. Steady-state analysis gave Km values of 1.5 mM and 7 microM for glycolate and oxygen, respectively. The Km for glycolate is slightly higher than that found for wild-type glycolate oxidase (1 mM) whereas the Km for oxygen is much lower. As was the case for wild-type glycolate oxidase, reduction was found to be the rate-limiting step in catalysis, with a rate of 0.63 s-1. The kinetic properties of Y129F glycolate oxidase provide evidence that the main function of the hydroxyl group of Tyr129 is the stabilization of the transition state.  相似文献   

4.
Kinetic and structural changes in recombinant human aldose reductase (AR) due to modification by S-nitrosoglutathione (GSNO) were investigated. Incubation of the enzyme with 10-50 microM GSNO led to a time- and concentration-dependent inactivation of the enzyme, with a second-order rate constant of 0.087 +/- 0.009 M-1 min-1. However, upon exhaustive modification, 30-40% of the enzyme activity was retained. The non-inactivated enzyme displayed a 2-3-fold change in Km for NADPH and Km fordl-glyceraldehyde, whereas the Km for the lipid peroxidation product, 4-hydroxy-2-trans nonenal (HNE), was comparable to that of the untreated enzyme. The residual activity of the enzyme after GSNO treatment was less sensitive to inhibition by the active site inhibitor sorbinil or to activation by sulfate. Significantly higher catalytic activity was retained when the enzyme was modified in the presence of NADPH, suggesting relatively low reactivity of the E-NADPH complex with GSNO. The modification site was identified using site-directed mutants in which each of the solvent-exposed cysteines of the enzyme was replaced individually by serine. The mutant C298S was insensitive to GSNO, whereas the sensitivity of the mutants C303S and C80S was comparable to that of the wild-type enzyme. Electrospray ionization mass spectroscopy of the GSNO-modified enzyme revealed a major modified species (70% of the protein) with a molecular mass that was 306 Da higher than that of the untreated enzyme, which is consistent with the addition of a single glutathione molecule to the enzyme. The remaining 30% of the protein displayed a molecular mass that was not significantly different from that of the native enzyme. No nitrosated forms of the enzyme were observed. These results suggest that inactivation of AR by GSNO is due to the selective formation of a single mixed disulfide between glutathione and Cys-298 located at the NADP(H)-binding site of the enzyme.  相似文献   

5.
1. Purification of horse-liver glutathione reductase was obtained by affinity chromatography on N6-(6-aminohexyl)-adenosine-1'5'-bisphosphate Sepharose (N6-2'5'-ADP-Sepharose) and Reactive Red-120-Agarose, and chromatography on DEAE-Sephadex and Sephacryl S-300. 2. The final preparation had 248 U/mg specific activity after 11,174-fold purification with 47% final recovery, and was homogeneous by SDS-electrophoresis. It showed charge heterogeneity in non-denaturing electrophoresis and chromatofocusing, with several peaks of pI between 5.7 and 6.7. 3. The enzyme was homodimeric (107,000 native MW), with S20w = 6.31 S, and 41.22 A of hydrodynamic radius. It showed absorption peaks at 270, 370 and 462 nm, a characteristic of flavoproteins. 4. When NADPH was substituted by deamino-NADPH or NADH the enzyme showed 69 and 8.5% activity, respectively, while with glutathione-CoA mixed disulfide the enzyme had 23% of the activity shown with GSSG. Apparent Km values of 8.8, 680, 59, and 560 microM were measured for NADPH, NADH, GSSG and ferricyanide, respectively.  相似文献   

6.
The Syk protein-tyrosine kinase is expressed in many hematopoietic cells and is involved in signaling from various receptors for antigen and Fc portions of IgG and IgE. After cross-linking of these receptors, Syk is rapidly phosphorylated on tyrosine residues. We have previously reported that Syk expressed in COS cells is predominantly phosphorylated at both Tyr518 and Tyr519 at its putative autophosphorylation site. In this study, we have examined the role of each of these two residues for the catalytic activity of Syk in vitro and for the Syk-induced phosphorylation of cellular proteins in intact cells. Mutation of either residue had minor effects on the catalytic activity of Syk, and even the double mutant [F518, F519]Syk was about 60% as active as the wild-type enzyme. In intact cells, however, all three mutants consistently failed to induce the extensive tyrosine phosphorylation of cellular proteins typically observed with wild-type Syk. We have recently shown that the doubly phosphorylated Y518/Y519 site is also the site for association of Syk with the SH2 domain of the Lck kinase, which suggests that although phosphates at Y518/Y519 may enhance the catalytic activity of Syk, its interaction with Src family protein-tyrosine kinases is at least equally important for the induction of downstream substrate phosphorylation.  相似文献   

7.
NADH oxidase from Amphibacillus xylanus is a potent alkyl hydroperoxide reductase in the presence of the small disulfide-containing protein (AhpC) of Salmonella typhimurium. In the presence of saturating AhpC, kcat values for reduction of hydroperoxides are approximately 180 s-1, and the double mutant flavoprotein enzyme C337S/C340S cannot support hydroperoxide reduction (Niimura, Y., Poole, L. B., and Massey, V. (1995) J. Biol. Chem. 270, 25645-25650). Kinetics of reduction of wild-type and mutant enzymes are reported here with wild-type enzyme; reduction by NADH was triphasic, with consumption of 2.6 equivalents of NADH, consistent with the known composition of one FAD and two disulfides per subunit. Rate constants for the first two phases (each approximately 200 s-1) where FAD and one disulfide are reduced are slightly greater than kcat values for AhpC-linked hydroperoxide reduction. The rate constant for the third phase (reduction to the 6-electron level) is too small for catalysis. Only the first phase of the wild-type enzyme occurs with the mutant enzyme. These results and the stoichiometry of NADH consumption indicate Cys337 and Cys340 as the active site disulfide of the flavoprotein and that electrons from FADH2 must pass through this disulfide to reduce the disulfide of AhpC.  相似文献   

8.
Soybean lipoxygenase-mediated cooxidation of reduced glutathione (GSH) and concomitant superoxide generation was examined. The oxidation of GSH was dependent on the concentration of linoleic acid (LA), GSH, and the enzyme. The optimal conditions to observe maximal enzyme velocity included the presence of 0.42 mM LA, 2 mM GSH, and 50 pmole of enzyme/mL. The GSH oxidation was linear up to 10 minutes and exhibited a pH optimum of 9.0. The reaction displayed a Km of 1.49 mM for GSH and Vmax of 1.35 +/- 0.02 mumoles/min/nmole of enzyme. Besides LA, arachidonic and gamma-linolenic acids also supported the lipoxygenase-mediated GSH oxidation. Hydrogen peroxide and 13-hydroperoxylinoleic acid supported GSH cooxidation, but to a very limited extent. Oxidized glutathione (GSSG) was identified as the major product of the reaction based on the depletion of nicotinamide-adenine dinucleotide 3'-phosphate (NADPH) in the presence of glutathione reductase. The GSH oxidation was accompanied by the reduction of ferricytochrome c, which can be completely abolished by superoxide dismutase (SOD), suggesting the generation of superoxide anion radicals. Under optimal conditions, the rate of superoxide generation (measured as the SOD-inhibitable reduction of ferricytochrome c) was 10 +/- 1.0 nmole/min/nmole of enzyme. These results clearly suggest that lipoxygenase is capable of oxidizing GSH to GSSG and simultaneously generating superoxide anion radicals, which may contribute to oxidative stress in cells under certain conditions.  相似文献   

9.
Crystallographic studies have demonstrated two flavin conformations for p-hydroxybenzoate hydroxylase (PHBH) [Gatti, D. L., Palfey, B. A. , Lah, M. S., Entsch, B., Massey, V., Ballou, D. P., & Ludwig, M. L. (1994) Science 266, 110-114. Schreuder, H. A., Mattevi, A., Obmolova, G., Kalk, K. H., Hol, W. G. J., van der Bolt, F. J. T., & van Berkel, W. J. H. (1994) Biochemistry 33, 10161-10170]. The isoalloxazine ring system of one conformation (the "out" conformation) is significantly more exposed to solvent and is not in position for necessary catalytic reactions, but when the natural substrate is bound to the enzyme, the isoalloxazine is in the correct position (the "in" conformation) for its chemical function. In this study, several aspects of the function of the conformational change in catalysis were explored using the wild-type and Tyr222Phe forms of PHBH substituted with 6-azido FAD. This flavin served as both a spectral probe and a photolabel. The enzyme containing 6-azido FAD was a relatively effective catalyst for the hydroxylation of p-hydroxybenzoate. However, the intermediate reduced 6-azido enzyme was chemically unstable, and a small fraction converted to 6-amino PHBH by the elimination of N2 during each catalytic cycle. The reduction of 6-azido FAD PHBH by NADPH was almost as fast as the reduction of the natural enzyme. The characteristic spectral change caused by NADPH binding prior to hydride transfer strongly suggests that flavin movement from the "in" to the "out" conformation precedes flavin reduction. Irradiation of 6-azido PHBH with visible light covalently labeled proline 293, an active site residue, under conditions in which the flavin adopted the "in" conformation, while no protein labeling occurred under conditions in which the flavin was "out". The labeled protein exchanged substrate and was reduced by NADPH much more slowly than before photolysis. It is therefore concluded that isoalloxazine movement is required for pyridine nucleotide to gain access to the active site and for the exchange of aromatic ligands.  相似文献   

10.
The mammalian ubiquitin conjugating enzyme known as E2-25K catalyzes the synthesis of polyubiquitin chains linked exclusively through K48-G76 isopeptide bonds. The properties of truncated and chimeric forms of E2-25K suggest that the polyubiquitin chain synthesis activity of this E2 depends on specific interactions between its conserved 150-residue core domain and its unique 50-residue tail domain [Haldeman, M. T., Xia, G., Kasperek, E. M., and Pickart, C. M. (1997) Biochemistry 36, 10526-10537]. In the present study, we provide strong support for this model by showing that a point mutation in the core domain (S86Y) mimics the effect of deleting the entire tail domain: the ability to form an E2 approximately ubiquitin thiol ester is intact, while conjugation activity is severely inhibited (>/=100-fold reduction in kcat/Km). The properties of E2-25K enzymes carrying the S86Y mutation indicate that this mutation strengthens the interaction between the core and tail domains: both free and ubiquitin-bound forms of S86Y-25K are completely resistant to tryptic cleavage at K164 in the tail domain, whereas wild-type enzyme is rapidly cleaved at this site. Other properties of S86Y-26K suggest that the active site of this mutant enzyme is more occluded than the active site of the wild-type enzyme. (1) Free S86Y-25K is alkylated by iodoacetamide 2-fold more slowly than the wild-type enzyme. (2) In assays of E2 approximately ubiquitin thiol ester formation, S86Y-25K shows a 4-fold reduced affinity for E1. (3) The ubiquitin thiol ester adduct of S86Y-25K undergoes (uncatalyzed) reaction with dithiothreitol 3-fold more slowly than the wild-type thiol ester adduct. One model to accommodate these findings postulates that an enhanced interaction between the core and tail domains, induced by the S86Y mutation, causes a steric blockade at the active site which prevents access of the incoming ubiquitin acceptor to the thiol ester bond. Consistent with this model, the S86Y mutation inhibits ubiquitin transfer to macromolecular acceptors (ubiquitin and polylysine) more strongly than transfer to small-molecule acceptors (free lysine and short peptides). These results suggest that unique residues proximal to E2 active sites may influence specific function by mediating intramolecular interactions.  相似文献   

11.
It is thought that only a limited number of residues in the extended binding pocket of coagulation proteases are critical for substrate and inhibitor specificity. A candidate residue from the crystal structures of thrombin and factor Xa (FXa) that may be critical for specificity at the S2 subsite is residue 99. Residue 99 is Tyr in FXa and Thr in activated protein C (APC). To determine the role of residue 99 in S2 specificity, a Gla-domainless mutant of protein C (GDPC) was prepared in which Thr99 was replaced with Tyr of FXa. GDPC T99Y bound Ca2+ and was activated by the thrombin-thrombomodulin complex normally. The T99Y mutant, similar to FXa, hydrolyzed the chromogenic substrates with a Gly at the P2 positions. This mutant was also inhibited by antithrombin (AT) (k2 = 4.2 +/- 0.2 x 10(1) M-1 s-1), and heparin accelerated the reaction >350-fold (k2 = 1.5 +/- 0.1 x 10(4) M-1 s-1). The T99Y mutant, however, did not activate prothrombin but inactivated factor Va approximately 2-fold better than wild type. To try to switch the specificity of FXa, both Tyr99 and Gln192 of FXa were replaced with those of APC in the Gla-domainless factor X (GDFX Y99T/Q192E). This mutant was folded correctly as it bound Ca2+ with a similar affinity as GDFX and was also activated by the Russell's viper venom at similar rate, but it cleaved the chromogenic substrates with a Gly at the P2 positions poorly. The mutant, instead, cleaved the APC-specific chromogenic substrates efficiently. The Y99T/Q192E mutant became resistant to inhibition by AT in the absence of heparin but was inhibited by AT almost normally in the presence of heparin (k2 = 3.4 +/- 0.5 x 10(5) M-1 s-1). The Y99T/Q192E mutant did not inactivate factor Va, and prothrombin activation by this mutant was impaired. These results indicate that 1) residue 99 is critical for enzyme specificity at the S2 subsite, 2) a role for heparin in acceleration of FXa inhibition by AT may involve the S2-P2 modulation, and 3) the exchange of residues 99 and 192 in FXa and APC may switch the enzyme specificity with the chromogenic substrates and inhibitors but not with the natural substrates.  相似文献   

12.
To investigate the functional role of the cysteine residues present in the spinach ferredoxin-NADP+ oxidoreductase, we individually replaced each of the five cysteine residues with serine using site-directed mutagenesis. All of the mutant reductases were correctly assembled in Escherichia coli except for the C42S mutant protein. C114S and C137S mutant enzymes apparently showed structural and kinetic properties very similar to those of the wild-type reductase. However, C272S and C132S mutations yielded enzymes with a decreased catalytic activity in the ferredoxin-dependent reaction (14 and 31% of the wild type, respectively). Whereas the C132S was fully competent in the diaphorase reaction, the C272S mutant flavoprotein showed a 35-fold reduction in catalytic efficiency with respect to the wild-type enzyme (0.4 versus 14.28 microM-1 s-1) due to a substantial decrease of kcat. NADP+ binding by the C272S mutant enzyme was apparently quantitatively the same (Kd = 37 microM) but qualitatively different, as shown by the differential spectrum. Stopped-flow experiments showed that the enzyme-FAD reduction rate was considerably decreased in the C272S mutant reductase, along with a much lower yield of the charge-transfer transient species. It is inferred from these data that the charge transfer (FAD-NADPH) between the reductase and NADPH is required for hydride transfer from the pyridine nucleotide to flavin to occur with a rate compatible with catalysis.  相似文献   

13.
A cephalosporin acylase from Pseudomonas strain N176 hydrolyses both 7-beta-(4-carboxybutanamido)-cephalosporanic acid (glutarylcephalosporanic acid) and cephalosporin C to 7-amino-cephalosporanic acid. However, its productivity in the original host was low and its activity against cephalosporin C was not sufficient for direct large-scale production of 7-amino-cephalosporanic acid. In order to overcome these problems, we established a high-level expression system for the acylase in Escherichia coli. Tyr270 in the acylase is reported to play an important role in the interaction with glutarylcephalosporanic acid, as determined from the reaction with an affinity-label reagent, 7 beta-(6-bromohexanoylamido) cephalosporanic acid [Ishii, Y., Saito, Y., Sasaki, H., Uchiyama, F., Hayashi, M., Nakamura, S. & Niwa, M. (1994) J. Ferment. Bioeng. 77, 598-603] and modification with tetranitromethane [Nobbs, T. J., Ishii, Y., Fujimura, T., Saito, Y. & Niwa, M. (1994) J. Ferment. Bioeng. 77, 604-609]. From carbamoylation with potassium cyanate and site-directed point mutagenesis of the cephalosporin C acylase, we have deduced that Tyr270 exists at a position where it can interact with a residue (possibly Ser239) corresponding to inactivation by carbamoylation. We mutated Met269 and Ala271 of the acylase and found that mutation of Met269 to Tyr or Phe caused a 1.6-fold and 1.7-fold increase, respectively, of specific activity against cephalosporin C as compared to that of the wild-type enzyme. Kinetic studies of these mutants revealed that their kcat values increased, although their Km values against cephalosporin C were not changed. These data indicate that the mutation of Met269 near Tyr270 induces a minor conformational change to increase the stability of the activated complex with the enzyme and cephalosporin C. In particular, a mutant in which Met269 was replaced by Tyr was 2.5-fold more efficient in converting cephalosporin C to 7-amino-cephalosporanic acid than the wild-type enzyme under conditions similar to those in a bio-reactor system.  相似文献   

14.
Conditions are described under which the nonphysiological substrate mercuric bromide (HgBr2) is rapidly turned over, both by the wild type (CCCC) and by an active site double mutant (CCAA) of mercuric reductase in which the C-terminal cysteines 557' and 558' are replaced by alanine and only the redox-active pair Cys135 and Cys140 are available for catalysis. A maximum rate of turnover kcatapp of approximately 18 s-1 (at 3 degreesC) for both enzymes is observed, and at high [HgBr2]/[enzyme] ratios, inhibition is found. The UV-vis spectral changes during turnover are closely similar in both enzymes, indicating that catalysis follows the same enzymatic mechanism. Single-turnover analysis of the mutant enzyme shows that after binding of HgBr2, two further rapid events ensue, followed by reduction of the metal ion (kobs approximately 23.5 s-1). It is shown that under multiple-turnover conditions, completion of the catalytic cycle must occur via an ordered mechanism where rapid binding of a new molecule of HgBr2 to EH2.NADP+ precedes exchange of the pyridine nucleotide. Binding of HgBr2 to the active site triple mutant C135A/C557A/C558A (ACAA) is ca. 100-fold slower compared to that of the CCAA mutant and results in no detectable turnover. It is concluded that in the reducible enzyme.Hg(II) complex, the metal ion is coordinated to Cys135 and Cys140 and that for efficient catalysis both residues are required. Furthermore, the data imply that binding to EH2.NADPH occurs via initial rate-limiting attack of Cys135, followed by reaction with Cys140.  相似文献   

15.
The flavoenzyme glutathione reductase catalyzes the NADPH-dependent reduction of glutathione disulfide, yielding two molecules of glutathione. The oxidation-reduction potentials, Eox/EH2 (two-electron reduced enzyme), for yeast, Escherichia coli, and human glutathione reductase have been determined between pH 6.0 and 9.8 relative to the nonphysiological substrate couple NAD+/NADH and were found to be -237, -243, and -227 mV (+/-5 mV) at pH 7.0 and 20 degreesC, respectively. The potential as a function of pH demonstrated slopes of -51, -45, and -42 mV/pH unit, respectively, at low pH and -37, -31, and -34 mV/pH unit, respectively, at high pH. The change in slope indicated pKa values of 7.4, 8.5, and 7.6, respectively. The slopes indicate that two protons are associated with the two-electron reduction of Eox at low pH and that only one proton is involved with the two-electron reduction of Eox at high pH, provided that the effects of nearby titratable residues are considered in the data analysis. The influence of four such groups, Cys50, Cys45, His456', and either Tyr107 or the flavin-(N3), has been included (residue numbering refers to the yeast sequence). The enzyme loses activity upon deprotonation of the acid-base catalyst at high pH. Since the pKa ascribed to the EH2-to-EH- ionization is lower than the pKa of the acid-base catalyst, both the EH2 and EH- forms of glutathione reductase must be catalytically active, in contrast to the closely related enzyme lipoamide dehydrogenase, for which only EH2 is active.  相似文献   

16.
We employed site-directed mutagenesis based on sequence comparisons and characterization of purified mutant enzymes to identify Glu558 and Asp766 of Syrian hamster 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase (EC 1.1.1.34) as essential for catalysis. Mutant enzymes E558D, E558Q, and D766N had wild-type Km values for (S)-HMG-CoA and NADPH, but exhibited less than 0.5% of the wild-type catalytic activity. The inactive mutant polypeptides E558Q and D766N nevertheless can associate to generate an active enzyme. In vitro, 6% of the wild-type activity was observed when mutant polypeptides E558D and D766N were mixed in the absence of chaotropic agents. When mutant polypeptides E558Q and D766N were co-expressed in Escherichia coli, the resulting purified enzyme had 25% of wild-type activity. Hamster HMG-CoA reductase thus is a two-site, dimeric enzyme whose subunits associate to form an active site in which each monomer contributes at least one residue (e.g. Glu558 from one monomer and Asp766 from the other). The wild-type enzyme behaves as a dimer during size exclusion chromatography and has one HMG-CoA binding site per monomer. Syrian hamster HMG-CoA reductase thus appears to be a homodimer with two active sites which are located at the subunit interface.  相似文献   

17.
Transgenic tobacco seedlings that overexpress a cDNA encoding an enzyme with both glutathione S-transferase (GST) and glutathione peroxidase (GPX) activity had GST- and GPX-specific activities approximately twofold higher than wild-type seedlings. These GST/GPX overexpressing seedlings grew significantly faster than control seedlings when exposed to chilling or salt stress. During chilling stress, levels of oxidized glutathione (GSSG) were significantly higher in transgenic seedlings than in wild-types. Growth of wild-type seedlings was accelerated by treatment with GSSG, while treatment with reduced glutathione or other sulfhydryl-reducing agents inhibited growth. Therefore, overexpression of GST/GPX can stimulate seedling growth under chilling and salt stress, and this effect could be caused by oxidation of the glutathione pool.  相似文献   

18.
The CorA transport system is the major Mg2+ influx pathway for bacteria and the Archaea. CorA contains three C-terminal transmembrane segments. No conserved charged residues are apparent within the membrane, suggesting that Mg2+ influx does not involve electrostatic interactions. We have mutated conserved residues within the third transmembrane segment to identify sites involved in transport. Mutation of conserved aromatic residues at either end of the membrane segment to alternative aromatic amino acids did not affect total cation uptake or cation affinity. Mutation to alanine greatly diminished uptake with little change in cation affinity implying that the conserved aromatic residues play a structural role in stabilizing this membrane segment of CorA at the interface between the bilayer and the aqueous environment. In contrast, mutation of Tyr292, Met299, and Tyr307 greatly altered the transport properties of CorA. Y292F, Y292S, Y292C, or Y292I mutations essentially abolished transport, without effect on expression or membrane insertion. M299C and M299A mutants exhibited a decrease in cation affinity for Mg2+, Co2+, or Ni2+ of 10-50-fold without a significant change in uptake capacity. Mutations at Tyr307 had no significant effect on cation uptake capacity; however, the affinity of Y307F and Y307A mutations for Mg2+ and Co2+ was decreased 3-10-fold, while affinity for Ni2+ was unchanged compared with the wild type CorA. In contrast, the affinity of the Y307S mutant for all three cations was decreased 2-5-fold. Projection of the third transmembrane segment as an alpha-helix suggests that Tyr292, Met299, and Tyr307 all reside on the same face of the alpha-helix. We interpret the transport data to suggest that a hydroxyl group is important at Tyr307, and that these three residues interact with Mg2+ during transport, forming part of the cation pore or channel within CorA.  相似文献   

19.
The phosphotriesterase from Pseudomonas diminuta catalyzes the hydrolysis of a wide array of phosphotriesters and related phosphonates, including organophosphate pesticides and military nerve agents. It has now been shown that this enzyme can also catalyze the hydrolysis of phosphodiesters, albeit at a greatly reduced rate. However, the enzymatic hydrolysis of ethyl-4-nitrophenyl phosphate (compound I) by the wild-type enzyme was >10(8) times faster than the uncatalyzed reaction (kcat = 0.06 s-1 and Km = 38 mM). Upon the addition of various alkylamines to the reaction mixture, the kcat/Km for the phosphodiester (compound I) increased up to 200-fold. Four mutant enzymes of the phosphotriesterase were constructed in a preliminary attempt to improve phosphodiester hydrolysis activity of the native enzyme. Met-317, which is thought to reside in close proximity to the pro-S-ethoxy arm of the paraoxon substrate, was mutated to arginine, alanine, histidine, and lysine. These mutant enzymes showed slight improvements in the catalytic hydrolysis of organophosphate diesters. The M317K mutant enzyme displayed the most improvement in catalytic activity (kcat = 0.34 s-1 and Km = 30 mM). The M317A mutant enzyme catalyzed the hydrolysis of the phosphodiester (compound I) in the presence of alkylamines up to 200 times faster than the wild-type enzyme in the absence of added amines. The neutralization of the negative charge on the oxygen atom of the phosphodiester by the ammonium cation within the active site is thought to be responsible for the rate enhancement by these amines in the hydrolytic reaction. These results demonstrate that an active site optimized for the hydrolysis of organophosphate triesters can be made to catalyze the hydrolysis of organophosphate diesters.  相似文献   

20.
Previous studies [Kondo, T., Dale, G. L. and Beutler, E. (1981) Biochim. Biophys. Acta, 645, 132-136] have shown evidence for the existence of two different active-transport processes for glutathione disulphide (GSSG) in human erythrocytes (the high-Km and low-Km processes). In the present investigation adenosine-triphosphate-dependent transport of glutathione S-conjugate was characterized in comparison with active glutathione transport using inside-out vesicles from human erythrocytes. Incubation of the vesicles with glutathione S-conjugate (S-2,4-dinitrophenylglutathione) was found to inhibit competitively the high-Km process of GSSG transport but not significantly affect the low-Km process. The glutathione S-conjugate transport required ATP. A lineweaver-Burk plot of the transport rate as a function of the conjugate concentration gave an apparent Km value of 0.94 mM. The Km value of ATP-Mg was 0.76 mM. The transport of glutathione S-conjugate was dependent on temperature. Preincubation of vesicles with dithiothreitol resulted in an increase of the transport rate while thiol reagents, such as iodoacetamide, N-ethylmaleimide and p-chloromercuribenzoate inhibited the transport. Addition of nucleotides, such as CTP, UTP or GTP had no effect on the transport. These findings suggest that glutathione S-conjugate formed by the catalytic reaction of glutathione S-transferase in erythrocytes under the exposure to electrophilic compounds, is eliminated via the same transport process for GSSG elevated under oxidative stress.  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