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 31 毫秒
1.
The effect of the antiviral, antitumoural xanthate D609 on the activity of phospholipase A2, C (PC- and Pi-specific) and D was investigated. D609 is the first model substance of a new concept of antiviral therapy that interferes with cellular regulation mechanisms, rather than with virus coded enzymes. Exclusively phosphatidylcholine (PC) specific phospholipase C (PC-PLC) was found to be inhibited in a dose-dependent manner. Enzyme activity was determined either as the rate of acid release from PC or as the rate of phosphorylcholine production form 3H labelled PC. Lineweaver-Burk plots revealed D609 as a competitive inhibitor of PC-PLC with a Ki of 6.4 microM. In addition, D609 competitively inhibited PC-PLC mediated cleavage of P-nitrophenylphosphorylcholine (p-NPP), a pseudo-substrate of PC-PLC with a Ki of 8.8 microM. These data suggest that D609 competes with the phosphorylcholine residue of PC for binding to PC-PLC.  相似文献   

2.
Listeria monocytogenes, a gram-positive facultative intracellular pathogen, produces two distinct phospholipases C. PC-PLC, encoded by plcB, is a broad-range phospholipase, whereas PI-PLC, encoded by plcA, is specific for phosphatidylinositol. It was previously shown that PI-PLC plays a role in efficient escape of L. monocytogenes from the primary phagosome. To further understand the function of PI-PLC in intracellular growth, site-directed mutagenesis of plcA was performed. Two potential active-site histidine residues were mutated independently to alanine, serine, and phenylalanine. With the exception of the activity of the enzyme containing H38F, which was unstable, the PI-PLC enzyme activities of culture supernatants containing each mutant enzyme were <1% of wild-type activity. In addition, the levels of expression of the mutant PI-PLC proteins were equivalent to wild-type expression. Derivatives of L. monocytogenes containing these specific plcA mutations were found to have phenotypes similar to that of the plcA deletion strain in an assay for escape from the primary vacuole, in intracellular growth in a murine macrophage cell line, and in a plaquing assay for cell-to-cell spread. Thus, catalytic activity of PI-PLC is required for all its intracellular functions.  相似文献   

3.
The role of amino acid residues located in the active site pocket of phosphatidylinositol-specific phospholipase C (PI-PLC) from Bacillus cereus[Heinz, D. W., Ryan, M., Bullock, T., & Griffith, O. H. (1995) EMBO J. 14, 3855-3863] was investigated by site-directed mutagenesis, kinetics, and crystal structure analysis. Twelve residues involved in catalysis and substrate binding (His32, Arg69, His82, Gly83, Lys115, Glu117, Arg163, Trp178, Asp180, Asp198, Tyr200, and Asp274) were individually replaced by 1-3 other amino acids, resulting in a total number of 21 mutants. Replacements in the mutants H32A, H32L, R69A, R69E, R69K, H82A, H82L, E117K, R163I, D198A, D198E, D198S, Y200S, and D274S caused essentially complete inactivation of the enzyme. The remaining mutants (G83S, K115E, R163K, W178Y, D180S, Y200F, and D274N) exhibited reduced activities up to 57% when compared with wild-type PI-PLC. Crystal structures determined at a resolution ranging from 2.0 to 2.7 A for six mutants (H32A, H32L, R163K, D198E, D274N, and D274S) showed that significant changes were confined to the site of the respective mutation without perturbation of the rest of the structure. Only in mutant D198E do the side chains of two neighboring arginine residues move across the inositol binding pocket toward the newly introduced glutamic acid. An analysis of these structure-function relationships provides new insight into the catalytic mechanism, and suggests a molecular explanation of some of the substrate stereospecificity and inhibitor binding data available for this enzyme.  相似文献   

4.
The I1-imidazoline receptor is expressed in the rostral ventrolateral medulla (RVLM) where it mediates vasodepression, and in PC12 pheochromocytoma cells where it elicits generation of diacylglycerol independent of phosphatidylinositol turnover or activation of phospholipase D. We hypothesized that the I1-imidazoline receptor couples to a phosphatidylcholine-selective phospholipase C (PC-PLC). The I1-agonist moxonidine elicited diacyglyceride accumulation and release of [3H]phosphocholine from PC12 cells prelabeled with [3H]choline. The PC-PLC inhibitor D609 abolished both responses. Microinjection of D609 into the RVLM of hypertensive rats blocked the vasodepressor response to intravenous moxonidine. These data implicate PC-PLC in cellular and organismic responses to I1-receptor stimulation.  相似文献   

5.
In Alcaligenes eutrophus H16 the hyp gene complex consists of six open reading frames hypA1, B1, F1, C, D and E whose products are involved in maturation of the two NiFe hydrogenases: an NAD-reducing cytoplasmic enzyme (SH) and a membrane-bound electron-transport-coupled protein (MBH). hypB1 and hypF1 were originally considered to form a single open reading frame designated hypB [Dernedde, J., Eitinger, M. & Friedrich, B. (1993) Arch. Microbiol. 159, 545-553]. Re-examination of the relevant sequence identified hypB1 and hypF1 as two distinct genes. Non-polar in-frame deletions in the individual hyp genes were constructed in vitro and transferred via gene replacement to the wild-type strain. The resulting mutants fall into two classes. Deletions in hypC, D and E (class I) gave a clear negative phenotype, while hypA1, B1 and F1 deletion mutants (class II) were not impaired in hydrogen metabolism. Class I mutants were unable to grow on hydrogen under autotrophic conditions. The enzymatic activities of SH and MBH were disrupted in all three class I mutants. Immunoblot analysis showed the presence of the H2-activating SH subunit (HoxH) at levels comparable to those observed in the wild-type strain whereas the other three subunits (HoxF, U and Y) were only detectable in trace amounts, probably due to proteolytic degradation. Likewise, MBH was less stable in hypC, D and E deletion mutants and was not attached to the cytoplasmic membrane. In the wild-type strain, HoxH and the MBH large subunit (HoxG) undergo C-terminal proteolytic processing before attaining enzymatic activity. In class I mutants this maturation was blocked. 63Ni-incorporation experiments identified both hydrogenases as nickel-free apoproteins in these mutants. Although class II mutants bearing deletions in hypA1, B1 and F1 showed no alteration of the wild-type phenotype, a role for these genes in the incorporation of nickel and hence hydrogenase maturation cannot be excluded, since there is experimental evidence that this set of genes is duplicated in A. eutrophus.  相似文献   

6.
The final step in the biosynthesis of the plant signaling molecule ethylene is catalyzed by 1-aminocyclopropane-1-carboxylate (ACC) oxidase, a member of the non-heme iron(II) dependent family of oxygenases and oxidases, which has a requirement for ascorbate as a co-substrate and carbon dioxide as an activator. ACC oxidase (tomato) has a particularly short half-life under catalytic conditions undergoing metal-catalyzed oxidative (MCO) fragmentation. Sequence comparisons of ACC oxidases with isopenicillin N synthase (IPNS) and members of the 2-oxoglutarate Fe(II) dependent dioxygenases show an aspartate and two of six ACC oxidase conserved histidine residues are completely conserved throughout this subfamily of Fe(II) dependent oxygenases/oxidases. Previous mutagenesis, spectroscopic, and crystallographic studies on IPNS indicate that the two completely conserved histidine and aspartate residues act as Fe(II) ligands. To investigate the role of the conserved aspartate and histidine residues in ACC oxidase (tomato fruit), they were substituted via site-directed mutagenesis. Modified ACC oxidases produced were H39Q, H56Q, H94Q, H177Q, H177D, H177E, D179E, D179N, H177D&D179E, H211Q, H234Q, H234D, and H234E. Among those histidine mutants replaced by glutamine, H39Q, H56Q, H94Q, and H211Q were catalytically active, indicating these histidines are not essential for catalysis. Mutant enzymes H177D, H177Q, D179N, H177D&D179E, H234Q, H234D, and H234E were catalytically inactive consistent with the assignment of H177, D179, and H234 as iron ligands. Replacement of H177 with glutamate or D179 with glutamate resulted in modified ACC oxidases which still effected the conversion of ACC to ethylene, albeit at a very low level of activity, which was stimulated by bicarbonate. The H177D (inactive), H177E (low activity), D179E (low activity), and H234Q (inactive) modified ACC oxidases all underwent MCO fragmentation, indicating that they can bind iron, dioxygen, ACC, and ascorbate. The results suggest that MCO cleavage results from active site-mediated reactions and imply that, while H177, D179, and H234 are all involved in metal ligation during catalysis, ligation to H234 is not required for fragmentation. It is possible that MCO fragmentation results from reaction of incorrectly folded or "primed" ACC oxidase.  相似文献   

7.
A universal protocol for PCR detection of 13 species of foodborne pathogens in foods was developed. The protocol used a universal culture medium and the same PCR conditions with 13 sets of specific primers. The 13 species of foodborne pathogens examined were Escherichia coli, E. coli-ETEC, E. coli-O157:H7, Shigella spp., Salmonella spp., Yersinia enterocolitica, Y. pseudotuberculosis, Vibrio cholerae, V. parahaemolyticus, V. vulnificus, Listeria monocytogenes, Staphylococcus aureus and Bacillus cereus. No interference was observed using the PCR assay when food sample was artificially inoculated with each individual bacterial species. Twelve different seafood samples and two soft cheese samples without artificial inoculation were examined by this protocol. Vibrio vulnificus, Salmonella spp., E. coli, Listeria monocytogenes and Bacillus cereus were detected in some foods. Internal probe hybridization and nested PCR procedures were used to confirm the above findings.  相似文献   

8.
Pigeon liver malic enzyme was inactivated by ferrous sulfate in the presence of ascorbate. Manganese and some other divalent metal ions provided complete protection of the enzyme against the Fe(2+)-induced inactivation. The inactivated enzyme was subsequently cleaved by the Fe(2+)-ascorbate system at Asp258-Ile259, which was presumably the Mn(2+)-binding site of the enzyme [Wei, C. H., Chou, W. Y., Huang, S. M., Lin, C. C., & Chang, G. G. (1994) Biochemistry 33, 7793-7936]. For identification of Asp258 as the putative metal-binding site of the enzyme, we prepared four mutant enzymes substituted at Asp258 with glutamate (D258E), asparagine (D258N), lysine (D258K), or alanine (D258A), respectively. These mutant proteins were recombinantly expressed in a bacterial expression system (pET-15b) with a stretch of histidine residues attached at the N-terminus and were successfully purified to apparent homogeneity by a single Ni-chelated affinity column. Among the four mutants, only D258E possessed 0.8% residual activity after purification; all other purified mutants had < 0.0001% residual activity in catalyzing the oxidative decarboxylation of L-malate. The D258E mutant was susceptible to inactivation by the Fe(2+)-ascorbate system, albeit with much slower inactivation rate, and was protected by the Mn2+ to a lesser extent as compared to the wild-type enzyme. None of the mutants were cleaved by the Fe(2+)-ascorbate system under conditions that cleaved the natural or wild-type enzyme at Asp258.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
This paper presents the first structural analysis of the cytoplasmic domain of neurofascin, which is highly conserved among the L1CAM family of cell adhesion molecules, and describes sequence requirements for neurofascin-ankyrin interactions in living cells. The cytoplasmic domain of neurofascin dimerizes in solution, has an asymmetric shape, and exhibits a reversible temperature-dependent beta-structure. Residues Ser56-Tyr81 are necessary for ankyrin binding but do not contribute to either dimerization or formation of structure. Transfected neurofascin recruits GFP-tagged 270-kDa ankyrinG to the plasma membrane of human embryo kidney 293 cells. Deletion mutants demonstrate that the sequence Ser56-Tyr81 contains the major ankyrin-recruiting activity of neurofascin. Mutations of the FIGQY tyrosine (Y81H/A/E) greatly impair neurofascin-ankyrin interactions. Mutation of human L1 at the equivalent tyrosine (Y1229H) is responsible for certain cases of mental retardation (Van Camp, G., Fransen, E., Vits, L., Raes, G., and Willems, P. J. (1996) Hum. Mutat. 8, 391). Mutations F77A and E73Q greatly impair ankyrin binding activity, whereas mutation D74N and a triple mutation of D57N/D58N/D62N result in less loss of ankyrin binding activity. These results provide evidence for a highly specific interaction between ankyrin and neurofascin and suggest that ankyrin association with L1 is required for L1 function in humans.  相似文献   

10.
We have used phage display to generate a panel of closely related catalytic antibodies. Seeking to improve the catalytic activity of an esterolytic antibody, we displayed libraries derived from the humanized Fab fragment of the antibody 17E8 (h17E8) on filamentous phage and sorted for binding to an immobilized transition-state analog (TSA). Previous work had suggested that residues outside the antibody active site contribute to TSA binding and catalytic efficiency, and we tested this notion by generating libraries containing such "second sphere" residues. Selected variants of h17E8 retained esterolytic activity and showed variations in affinity within 40-fold and kinetic parameters within tenfold of wild-type antibody, indicating that residues remote from the active site do modulate catalytic activity. In order to understand which mutations were responsible for the properties of phage-selected variants, we designed a series of site-directed mutants. From this series, we identified a double mutant in which Tyr97 was changed to Arg in the heavy chain (Y97HR) and the heavy chain Tyr100a was mutated to Asn (Y100aHN). This variant showed a tenfold improvement in catalytic efficiency (kcat/KM) relative to wild-type h17E8. These mutations were additive; Y97HR increases the catalytic turnover (kcat) by three- to fourfold, while Y100aHN has been shown to lower the Michaelis constant (KM) by three- to fivefold. TSA binding was correlated with catalytic turnover for variants that differed by single mutations, but less so for variants that differed by many mutations. Thus, future selections based on TSA binding should focus on mutating a small number of residues at a time.  相似文献   

11.
We have studied mu-conotoxin (mu-CTX) block of rat skeletal muscle sodium channel (rSkM1) currents in which single amino acids within the pore (P-loop) were substituted with cysteine. Among 17 cysteine mutants expressed in Xenopus oocytes, 7 showed significant alterations in sensitivity to mu-CTX compared to wild-type rSkM1 channel (IC50 = 17.5 +/- 2.8 nM). E758C and D1241C were less sensitive to mu-CTX block (IC50 = 220 +/- 39 nM and 112 +/- 24 nM, respectively), whereas the tryptophan mutants W402C, W1239C, and W1531C showed enhanced mu-CTX sensitivity (IC50 = 1.9 +/- 0.1, 4.9 +/- 0.9, and 5.5 +/- 0.4 nM, respectively). D400C and Y401C also showed statistically significant yet modest (approximately twofold) changes in sensitivity to mu-CTX block compared to WT (p < 0.05). Application of the negatively charged, sulfhydryl-reactive compound methanethiosulfonate-ethylsulfonate (MTSES) enhanced the toxin sensitivity of D1241C (IC50 = 46.3 +/- 12 nM) while having little effect on E758C mutant channels (IC50 = 199.8 +/- 21.8 nM). On the other hand, the positively charged methanethiosulfonate-ethylammonium (MTSEA) completely abolished the mu-CTX sensitivity of E758C (IC50 > 1 microM) and increased the IC50 of D1241C by about threefold. Applications of MTSEA, MTSES, and the neutral MTSBN (benzyl methanethiosulfonate) to the tryptophan-to-cysteine mutants partially or fully restored the wild-type mu-CTX sensitivity, suggesting that the bulkiness of the tryptophan's indole group is a determinant of toxin binding. In support of this suggestion, the blocking IC50 of W1531A (7.5 +/- 1.3 nM) was similar to W1531C, whereas W1531Y showed reduced toxin sensitivity (14.6 +/- 3.5 nM) similar to that of the wild-type channel. Our results demonstrate that charge at positions 758 and 1241 are important for mu-CTX toxin binding and further suggest that the tryptophan residues within the pore in domains I, III, and IV negatively influence toxin-channel interaction.  相似文献   

12.
B*2704 and B*2706 are closely related HLA-B27 subtypes of which the former but not the latter is associated to ankylosing spondylitis. Their peptide specificity relative to other disease-associated subtypes was analyzed by testing binding of self-peptides naturally presented by B*2705 or B*2702, and synthetic analogs, to B*2704, B*2706, and site-specific mutants mimicking their changes. Peptides with basic, aliphatic or aromatic C-terminal residues bound to B*2705 with similar affinity. In B*2704 C-terminal aliphatic/ aromatic residues were preferred. B*2706 discriminated drastically between polar and nonpolar C-terminal residues, showing strong preference for Leu and Phe, and less than B*2704 for basic and Tyr residues. Loss of single acidic charges (D > S77, D > Y116) increased preference for C-terminal Leu and Phe, but allowed efficient binding of peptides with basic residues or Tyr. Their gain (V > E152, H > D114) maintained wide C-terminal specificity, but severely impaired binding, presumably by disrupting interactions with internal peptide residues. This was compensated by Y116 in the double D114Y116 mutant. The specificity of B*2704 and B*2706 was explained only partially by the separate effects of single mutations, indicating that novel properties arise from concomitant changes at various positions. For instance, specificity of B*2706 for nonpolar C-terminal residues required simultaneous removal of Asp77 and Asp116. B*2706 differed from B*2705, B*2702, and B*2704 in its lower suitability for C-terminal Tyr, suggesting that this feature might be relevant for HLA-B27 association to spondyloarthropathy.  相似文献   

13.
Shiga-like toxin I (SLT-I), the potent cytotoxin produced by certain pathogenic strains of Escherichia coli, is a member of a burgeoning family of ribosome-in-activating proteins (RIPs), which share common structural and mechanistic features. The prototype of the group is the plant toxin ricin. Recently we proposed a structural model for the Slt-IA active site, based in part on the known geometry of the enzymatic subunit of the ricin toxin. The model places three aromatic residues within the putative Slt-IA active site cleft: tyrosine 77, tyrosine 114, and tryptophan 203. Here we present biochemical and biophysical data regarding, the phenotypes of conservative point mutants of Slt-IA in which tyrosine 114 is altered. We used oligonucleotide-directed mutagenesis to replace tyrosine 114 with either phenylalanine (Y114F) or serine (Y114S). Periplasmic extracts of E. coli containing wild-type or mutant Slt-IA were tested for their ability to inhibit protein synthesis in vitro. Relative to wild-type, the activity of mutant Y114F was attenuated about 30-fold, while the mutant Y114S was attenuated about 500 to 1000-fold. In order to address the possibility that differential activation of the mutants rather than local effects at the active site might account for their diminished activity, we engineered the same mutations into a truncated slt-IA cassette that directs expression of a product corresponding to the activated A1 form of Slt-IA (wild-type-delta). The same general relationships held: relative to wild type-delta, Y114F-delta was attenuated about 7-fold, and Y114S-delta about 300-fold.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
The high-affinity streptavidin-biotin complex is characterized by an extensive hydrogen-bonding network. A study of hydrogen-bonding energetics at the ureido oxygen of biotin has been conducted with site-directed mutations at Asn 23, Ser 27, and Tyr 43. A new competitive biotin binding assay was developed to provide direct equilibrium measurements of the alterations in Kd. S27A, Y43F, Y43A, N23A, and N23E mutants display DeltaDeltaG degrees at 37 degrees C relative to wild-type streptavidin of 2.9, 1.2, 2.6, 3.5, and 2.6 kcal/mol, respectively. The equilibrium-binding enthalpies for all of the mutants were measured by isothermal titration calorimetry, and the Y43A and N23A mutants display large decreases in the equilibrium binding enthalpy at 25 degrees C of 8.9 and 6.9 kcal/mol, respectively. The S27A and N23E mutants displayed small decreases in binding enthalpy of 1.6 and 0.9 kcal/mol relative to wild-type, while the Y43F mutant displayed a -2.6 kcal/mol increase in the binding enthalpy at 25 degrees C. At 37 degrees C, the Y43A and N23A mutants display decreases of 7.8 and 7.9 kcal/mol, respectively, while the S27A, N23E, and Y43F mutants displayed decreases of 4.9, 3.7, and 1.2 kcal/mol relative to wild-type. Kinetic analyses were also conducted to probe the contributions of the hydrogen bonds to the activation barrier. Wild-type streptavidin at 37 degrees C displays a koff of (4.1 +/- 0.3) x 10(-5) s-1, and the conservative Y43F, S27A, and N23A mutants displayed increases in koff to (20 +/- 1) x 10(-5) s-1, (660 +/- 40) x 10(-5) s-1, and (1030 +/- 220) x 10(-)5 s-1, respectively. The Y43A and N23E mutants displayed 93-fold and 188-fold increases in koff, respectively. Activation energies and enthalpies for each of the mutants were determined by transition-state analysis of the dissociation rate temperature dependence. All of the mutants except Y43F display large reductions in the activation enthalpy. The Y43F mutant has a more positive activation enthalpy, and thus a more favorable activation entropy that underlies the overall reduction in the activation barrier. For the most conservative mutant at each ureido oxygen hydrogen-bonding position, bound-state alterations account for most of the energetic changes in a single transition-state model, suggesting that the ureido oxygen hydrogen-bonding interactions are broken in the dissociation transition state.  相似文献   

15.
Structural models of subunit epsilon of the ATP synthase from Escherichia coli have been determined recently by NMR [Wilkens et al. (1995) Nat. Struct. Biol. 2, 961-967] and by X-ray crystallography [Uhlin et al. (1997) Structure 5, 1219-1230], revealing a two-domain protein. In this study, six new epsilon mutants were constructed and analyzed: Y63A, D81A, T82A, and three truncated mutants, tr80(S), tr94(LAS), and tr117(AS). Seven mutants constructed previously were also analyzed: E31A, E59A, S65A, E70A, T77A, R58A, and D81A/R85A. Subunits were purified by isoelectric focusing from extracts of cells that overproduced these 13 mutants. F1 was prepared lacking subunit epsilon by immobilized-Ni affinity chromatography. Three mutants, E70A, S65A, and E31A, showed somewhat higher affinities and extents of inhibition than the wild type. Three mutants, T82A, R85A, and tr94(LAS), showed both lower affinities and extents of inhibition, over the concentration range tested. Two showed no inhibition, D81A and tr80(S). The others, T77A, Y63A, E59A, and tr117(AS), showed lower affinities than wild type, but the extents of inhibition were nearly normal. Results indicate that the C-terminal domain of subunit epsilon contributes to inhibition of ATP hydrolysis, but it is not necessary for ATP-driven proton translocation. Interactions with subunit gamma are likely to involve a surface containing residues S65, E70, T77, D81, and T82, while residues R85 and Y63 are likely to be important in the conformation of subunit epsilon.  相似文献   

16.
The EcoRI endonuclease is an important recombinant DNA tool and a paradigm of sequence-specific DNA-protein interactions. We have isolated temperature-sensitive (TS) EcoRI endonuclease mutants (R56Q, G78D, P90S, V97I, R105K, M157I, C218Y, A235E, M255I, T261I and L263F) and characterized activity in vivo and in vitro. Although the majority were TS for function in vivo, all of the mutant enzymes were stably expressed and largely soluble at both 30 degrees C and 42 degrees C in vivo and none of the mutants was found to be TS in vitro. These findings suggest that these mutations may affect folding of the enzyme at elevated temperature in vivo. Both non-conservative and conservative substitutions occurred but were not correlated with severity of the mutation. Of the 12 residues identified, 11 are conserved between EcoRI and the isoschizomer RsrI (which shares 50% identity), a further indication that these residues are critical for EcoRI structure and function. Inspection of the 2.8 A resolution X-ray crystal structure of the wild-type EcoRI endonuclease-DNA complex revealed that: (1) the TS mutations cluster in one half of the globular enzyme; (2) several of the substituted residues interact with each other; (3) most mutations would be predicted to disrupt local structures; (4) two mutations may affect the dimer interface (G78D and A235E); (5) one mutation (P90S) occurred in a residue that is part of, or immediately adjacent to, the EcoRI active site and which is conserved in the distantly related EcoRV endonuclease. Finally, one class of mutants restricted phage in vivo and was active in vitro, whereas a second class did not restrict and was inactive in vitro. The two classes of mutants may differ in kinetic properties or cleavage mechanism. In summary, these mutations provide insights into EcoRI structure and function, and complement previous genetic, biochemical, and structural analyses.  相似文献   

17.
Site-directed mutagenesis was used to identify key amino acid residues of the cholesterol oxidase from Streptomyces sp., which catalyzes the oxidation of cholesterol and the isomerization of 5-cholesten-3-one. Eight mutant enzymes were constructed and the following amino acid substitutions were identified: N318A, N318H, E356A, E356D, H441A, H441N, N480A and N480Q. Mutants N318A and N318H retained both oxidation and isomerization activities. The mutant E356D retained oxidation but not isomerization activity. On the other hand, mutants N480A and N480Q showed no oxidation activity but retained their isomerization activities. The two catalytic reactions, oxidation and isomerization, in cholesterol oxidase were thus successfully separated. When the H441A or H441N mutation was introduced, both the oxidase and isomerase activities were completely lost. The H441, E356 and N480 residues thus appear to participate in the catalysis of cholesterol oxidase, whereas N318 does not. An analysis of the products of these mutant enzymes suggested that the previously proposed 6-hydroxylation reaction by cholesterol oxidase is actually autooxidation from 5-cholesten-3-one. Kinetic studies of the purified wild-type and mutant enzymes showed that the k(cat)/Km values for oxidation in E356D and for isomerization in N480A increased six- and threefold, respectively, over those in the wild-type. These mutational effects and the reaction mechanisms are discussed in terms of the three-dimensional structure of the enzyme constructed on the basis of homology modeling.  相似文献   

18.
Pyridine nucleotide transhydrogenase (EC 1.6.1.1) from Escherichia coli was investigated with respect to the role of glutamic and aspartic acid residues reactive to N,N'-dicyclohexylcarbodiimide (DCCD) and potentially involved in the proton-pumping mechanism of the enzyme. The E. coli transhydrogenase consists of an alpha (510 residues) and a beta (462 residues) subunit. DCCD reacts with the enzyme to inhibit catalytic activity and proton pumping. This reagent modifies Asp alpha 232, Glu alpha 238, and Glu alpha 240 as well as amino acid residue(s) in the beta subunit. Using the cloned and overexpressed E. coli transhydrogenase genes (Clarke, D. M., and Bragg, P. D. (1985) J. Bacteriol. 162, 367-373), Asp alpha 232 and Glu alpha 238 were replaced independently by site-specific mutagenesis. In addition, Asp alpha 232, Glu alpha 238, and Glu alpha 240 were replaced to generate triple mutants. The specific catalytic activities of the mutant transhydrogenases alpha D232N, alpha D232E, alpha D232K, alpha D232H, alpha E238K, and alpha E238Q as well as of the triple mutants alpha D232N, alpha E238Q, alpha E240Q and alpha D232H, alpha E238Q, alpha E240Q were in the range of 40-90% of the wild-type activity. Proton-pumping activity was present in all mutants. Examination of the extent of subunit modification by [14C]DCCD revealed that the label was still incorporated into both alpha and beta subunits in the Asp alpha 232 mutants, but that the alpha subunit was not labeled in the triple mutants. Catalytic and proton-pumping activities were nearly insensitive to DCCD in the triple mutants. This suggests that loss of catalytic and proton-pumping activities is associated with modification of the aspartic and glutamic acid residues of the alpha subunit. In the presence of the substrate NADPH, the rate of modification of the beta subunit by [14C]DCCD was increased, and there was a greater extent of enzyme inactivation. By contrast, NADH and 3-acetylpyridine-NAD+ protected the catalytic activity of the transhydrogenase from inhibition by DCCD. The protection was particularly marked in the E238Q and E238K mutants. It is concluded that the Asp alpha 232, Glu alpha 238, and Glu alpha 240 residues are not essential for catalytic activity or proton pumping. The inactivation by DCCD is likely due to the introduction of a sterically hindering group that reacts with the identified acidic residues close to the NAD(H)-binding site.  相似文献   

19.
Previous studies demonstrated that at least two mechanisms are involved in the epoxidation of styrene and stilbene by myoglobin and H2O2 (Ortiz de Montellano, P. R., and Catalano, C. E. (1985) J. Biol. Chem. 260, 9265-9271). One mechanism, reaction of the olefin with the ferryl oxygen, preserves the olefin stereochemistry and incorporates an oxygen from H2O2 into the epoxide. The second mechanism, proposed to be a protein-mediated co-oxidation process, results in loss of stereochemistry and incorporation of an atom of molecular oxygen. To examine the role of individual residues in olefin epoxidation, we have examined the catalytic activities of the possible Tyr-->Phe mutants, the His-64-->Val mutant, and a protein combining all the tyrosine and histidine mutations. The latter protein is less stable than the other mutants and is the only one for which a protein radical is not detected in the reaction with H2O2. The Km and Vmax for styrene epoxidation range, respectively, from 0.3-8 mM and 12-35 pmol/min/nmol of protein. Incubation with H(2)18O2 results in 20-30% incorporation of labeled oxygen into the epoxide with all the mutants except Y103F/Y151F, Y146F/Y151F, H64V, and H64V/Y103F/Y146F/Y151F, for which 52, 58, 89, and 96% of the epoxide oxygen, respectively, is labeled. Oxidation of cis-beta-methylstyrene by wild-type myoglobin yielded a 54:46 ratio of cis- and trans-beta-methylstyrene oxides. The cis-isomer accounts for 47-100% of the epoxide produced by the mutant hemoproteins, with the two H64V mutants yielding almost exclusively the cis-epoxide. The oxygen in the cis-epoxide derives primarily or exclusively from H2O2 and that in the trans-epoxide from an alternative source. These results indicate that tyrosine residues may participate in, but are not essential for, protein-mediated epoxidation. In contrast, His-64 appears to be essential for co-oxidative epoxidation because in its absence olefin epoxidation is mediated almost exclusively by ferryl oxygen transfer.  相似文献   

20.
Attempts to modify the guanine specificity of ribonuclease T1 (RNase T1) by rationally designed amino acid substitutions failed so far. Therefore, we applied a semirational approach by randomizing the guanine binding site. A combinatorial library of approximately 1.6 million RNase T1 variants containing permutations of 6 amino acid positions within the recognition loop was screened on RNase indicator plates. The specificity profiles of 180 individual clones showing RNase activity revealed that variant K41S/N43W/N44H/Y45A/E46D (RNaseT1-8/3) exhibits an altered preference toward purine nucleotides. The ApC/GpC preference in the cleavage reaction of this variant was increased 4000-fold compared to wild-type. Synthesis experiments of dinucleoside monophosphates from cytidine and the corresponding 2'3'-cyclic diesters using the reverse reaction of the transesterification step showed a 7-fold higher ApC synthesis rate of RNase 8/3 than wild-type, whereas the GpC synthesis rates for both enzymes were comparable. This study shows that site-directed random mutagenesis is a powerful additional tool in protein design in order to achieve new enzymatic specificities.  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