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.
Three mutant Escherichia coli aspartate aminotransferases in which Val39 was changed to Ala, Leu, and Phe by site-directed mutagenesis were prepared and characterized. Among the three mutant and the wild-type enzymes, the Leu39 enzyme had the lowest Km values for dicarboxylic substrates. The Km values of the Ala39 enzyme for dicarboxylates were essentially the same as those of the wild-type (Val39) enzyme. These two mutant enzymes showed essentially the same kcat values for dicarboxylic substrates as did the wild-type enzyme. On the other hand, incorporation of a bulky side-chain at position 39 (Phe39 enzyme) decreased both the affinity (1/Km) and catalytic ability (kcat) toward dicarboxylic substrates. These results show that the position 39 residue is involved in the modulation of both the binding of dicarboxylic substrates to enzyme and the catalytic ability of the enzyme. Although the replacement of Val39 with other residues altered both the kcat and Km values toward various substrates including dicarboxylic and aromatic amino acids and the corresponding oxo acids, it did not alter the ratio of the kcat/Km value of the enzyme toward a dicarboxylic substrate to that for an aromatic substrate. The affinity for aromatic substrates was not affected by changing the residue at position 39. These data indicate that, although the side chain bulkiness of the residue at position 39 correlates well with the activity toward aromatic substrates in the sequence alignment of several aminotransferases [Seville, M., Vincent M.G., & Hahn, K. (1988) Biochemistry 27, 8344-8349], the residue does not seem to be involved in the recognition of aromatic substrates.  相似文献   

2.
The role of the C-terminal region of Staphylococcal nuclease (SNase) was examined by deletion mutation. Deletions up to eight residues do not affect the structure and function. The structure and enzymatic activity were partially lost by deleting Ser141-Asn149 (Delta141-149), and deletion of Trp140-Asn149 (Delta140-149) resulted in further loss of structure and activity. A 13-residue deletion showed the same effect as the 10-residue deletion. Both Ser141Gln and Ser141Ala mutations for an eight-residue deletion mutant did not alter properties as well as Ser141A1a for full-length SNase. In contrast, Trp140Ala mutation for Delta141-149 shows the same effect as the deletion of Trp140. Trp140Ala mutation for full-length SNase causes the loss of native structure. These observations indicate the significance of the 140th and the 141st residues. The side-chain of the 140th residue is required to be tryptophan; however, the backbone of the 141st residue is solely critical for foldability, but the side-chain information is not crucial. All of the mutants that take a non-native conformation show enzymatic activity and inhibitor-induced folding, suggesting that foldability is required for the activity.  相似文献   

3.
To assess the structural requirements for G(s) coupling by prostaglandin E receptors (EPs), the G(s)-coupled EP2 and G(i)-coupled EP3beta receptors were used to generate hybrid receptors. Interchanging of the whole i2 loop and its N-terminal half (i2N) had no effect on the binding of both receptors expressed in HEK293 cells. Agonist-induced cAMP formation was observed in wild type EP2 but not in the i2 loop- or i2N-substituted EP2. Wild type EP3beta left cAMP levels unaffected, whereas i2 loop- and i2N-substituted EP3 gained agonist-induced adenylyl cyclase stimulation. In EP2, the ability to stimulate cAMP formation was lost by mutation of Tyr(143) into Ala but retained by mutations into Phe, Trp, and Leu. Consistent with this observation, substitution of the equivalent His(140) enabled EP3beta to stimulate cAMP formation with the rank order of Phe > Tyr > Trp > Leu. The point mutation of His(140) into Phe was effective in another EP3 variant in which its C-terminal tail is different or lacking. Simultaneous mutation of the adjacent Trp(141) to Ala but not at the following Tyr(142) weakened the acquired ability to stimulate cAMP levels in the EP3 mutant. Mutation of EP2 at adjacent Phe(144) to Ala but not at Tyr(145) reduced the efficiency of agonist-induced cAMP formation. In Chinese hamster ovary cells stably expressing G(s)-acquired EP3 mutant, an agonist-dependent cAMP formation was observed, and pertussis toxin markedly augmented cAMP formation. These results suggest that a cluster of hydrophobic aromatic amino acids in the i2 loop plays a key role for G(s) coupling.  相似文献   

4.
Chitinase B (ChiB) of S. marcescens has five exposed aromatic residues linearly aligned toward the catalytic cleft, Tyr481 and Trp479 in the C-terminal domain, and Trp252, Tyr240 and Phe190 in the catalytic domain. To determine the contribution of these residues to the hydrolysis of crystalline beta-chitin, site-directed mutagenesis, to replace them by alanine, was carried out. The Y481A, W479A, W252A, and Y240A mutations all decreased the binding activity and hydrolyzing activity toward beta-chitin microfibrils. Substitution of Trp residues affected the binding activity more severely than that of Tyr residues. The F190A mutation decreased neither the binding activity nor the hydrolyzing activity. None of the mutations decreased the hydrolyzing activity toward soluble substrates. These results suggest that ChiB hydrolyzes crystalline beta-chitin via a mechanism in which four exposed aromatic residues play important roles, similar to the mechanism of hydrolysis by ChiA of this bacterium, although the directions of hydrolysis of the two chitinases are opposite.  相似文献   

5.
Two notable features of the thermophilic CYP119, an Arg154-Glu212 salt bridge between the F-G loop and the I helix and an extended aromatic cluster, were studied to determine their contributions to the thermal stability of the enzyme. Site-specific mutants of the salt bridge (Arg154, Glu212) and aromatic cluster (Tyr2, Trp4, Trp231, Tyr250, Trp281) were expressed and purified. The substrate-binding and kinetic constants for lauric acid hydroxylation are little affected in most mutants, but the E212D mutant is inactive and the R154Q mutant has higher K(s),K(m), and k(cat) values. The salt bridge mutants, like wild-type CYP119, melt at 91+/-1 degrees C, whereas mutation of individual residues in the extended aromatic cluster lowers the T(m) by 10-15 degrees C even though no change is observed on mutation of an unrelated aromatic residue. The extended aromatic cluster, but not the Arg154-Glu212 salt bridge, contributes to the thermal stability of CYP119.  相似文献   

6.
The transport of [125I]triiodothyronine ([125I]T3) and [3H]tryptophan ([3H]Trp) by washed rat erythrocytes was studied at 25 degrees C in the presence of leucine in order to block the neutral amino acid transport system L. Eadie-Hofstee plots of initial velocity data gave the following values of Km (micromolar) and Vmax (nanomole/min/10(8) cells): 0.122 +/- 0.007 and 0.140 +/- 0.021 for T3, and 558 +/- 28 and 17.4 +/- 2.3 for Trp (n = 5). The Trp transport system in rat erythrocytes is similar to the human erythrocyte aromatic amino acid-specific system T described by Rosenberg et al. (Rosenberg, R., Young, J. D., and Ellory, J. C. (1980) Biochim. Biophys. Acta 598, 375-384). Unlabeled aromatic amino acids (e.g. Trp, phenylalanine, tyrosine) competitively inhibited [125I]T3 uptake and unlabeled iodothyronine analogues (e.g. T3, D-T3, thyroxine, thyronine) competitively inhibited [3H]Trp uptake. The inhibition constants of these competitors measured with each labeled substrate were highly correlated. N-Ethylmaleimide irreversibly inhibited T3 and Trp transport and each substrate protected the transport system of the other from inactivation by N-ethylmaleimide. The Vmax of T3 and Trp transport by human erythrocytes were 500 and 120 times lower, respectively, than those of rat erythrocytes (0.30 and 126 pmol/min/10(8) cells, respectively). The T3 and Trp transport activities of sheep erythrocytes were undetectable. These results indicate that T3 and Trp either share a common multi-specific transport system or are transported by closely linked systems which interact in the erythrocyte membrane.  相似文献   

7.
We generated four point mutations in the DNA-binding protein (DBP) gene of adenovirus type 5 by oligonucleotide-directed site-specific mutagenesis. The sites mutated were in the three conserved regions (CR; amino acids 178-186 [CR1], 322-330 [CR2], and 464-475 [CR3]) identified previously by comparative sequence analysis (G. R. Kitchingman, Virology 146:90-101, 1985). The mutations resulted in changes in amino acids 181 (Trp to Leu), 323 (Arg to Leu), 324 (Trp to Leu), and 469 (Phe to Ile). The mutated DBP genes were put under the control of the simian virus 40 early promoter and analyzed by transfection for their ability to help adeno-associated virus replicate its DNA in COS-1 monkey cells. Mutations in the aromatic amino acids 324 and 469 reduced the amount of AAV DNA replication approximately 10-fold, while the mutation in Arg 323 produced a reduction of approximately fourfold. The Trp-to-Leu mutation in amino acid 181 had no effect on AAV DNA replication. The decreased helper activity of the 323, 324, and 469 mutations was not caused by any effect of the mutation on the stability of the DBP. These results suggest that CR2 and CR3 are involved in AAV helper activity, specifically in AAV DNA replication. The relevance of these findings to the identification of residues important for the functions of DBP in adenovirus infection is discussed.  相似文献   

8.
An SV40-based shuttle vector system was used to identify the types of mutational changes and the sites of mutation within the supF DNA sequence generated by the four stereoisomers of benzo[c]phenanthrene 3,4-dihydrodiol 1,2-epoxide (B[c]PhDE), by racemic mixtures of bay or fjord region dihydrodiol epoxides (DE) of 5-methylchrysene, of 5,6-dimethylchrysene, of benzo[g]chrysene and of 7-methylbenz[a]anthracene and by two direct acting polycyclic aromatic hydrocarbon carcinogens, 7-bromomethylbenz[a]anthracene (7-BrMeBA) and 7-bromomethyl-12-methylbenz[a]anthracene (7-BrMe-12-MeBA). The results of these studies demonstrated that the predominant type of mutation induced by these compounds is the base substitution. The chemical preference for reaction at deoxyadenosine (dAdo) or deoxyguanosine (dGuo) residues in DNA, which is in general correlated with the spatial structure (planar or non-planar) of the reactive polycyclic aromatic hydrocarbon, is reflected in the preference for mutation at AT or GC pairs. In addition, if the ability to react with DNA in vivo is taken into account, the relative mutagenic potencies of the B[c]PhDE stereoisomers are consistent with the higher tumorigenic activity associated with non-planar polycyclic aromatic hydrocarbons and their extensive reaction with dAdo residues in DNA. Comparison of the types of mutations generated by polycyclic aromatic hydrocarbons and other bulky carcinogens in this shuttle vector system suggests that all bulky lesions may be processed by a similar mechanism related to that involved in replication past apurinic sites. However, inspection of the distribution of mutations over the target gene induced by the different compounds demonstrated that individual polycyclic aromatic hydrocarbons induce unique patterns of mutational hotspots within the target gene. A polymerase arrest assay was used to determine the sequence specificity of the interaction of reactive polycyclic aromatic hydrocarbons with the shuttle vector DNA. The results of these assays revealed a divergence between mutational hotspots and polymerase arrest sites for all compounds investigated, i.e., sites of mutational hotspots do not correspond to sites where high levels of adduct formation occur, and suggested that some association between specific adducts and sequence context may be required to constitute a premutagenic lesion. A site-specific mutagenesis system employing a single-stranded vector (M13mp7L2) was used to investigate the mutational events a single benzo[a]pyrene or benzo[c]phenanthrene dihydrodiol epoxide–DNA adduct elicits within specific sequence contexts. These studies showed that sequence context can cause striking differences in mutagenic frequencies for given adducts. In addition, these sequence context effects do not originate only from nucleotides immediately adjacent to the adduct, but are also modulated by more distal nucleotides. The implications of these results for mechanisms of polycyclic aromatic hydrocarbon-induced mutagenesis and carcinogenesis are discussed.  相似文献   

9.
Three human lysozymes containing a mutation either at Asp-53 to Glu or at Tyr-63 to Trp or Phe were synthesized and examined for their immunological and enzymatical activities in comparison with the native one. All mutants were immunologically indistinguishable from native human lysozyme. The [Trp63] and [Phe63] mutants catalysed the hydrolysis of Micrococcus lysodeikticus cell wall and glycol chitin effectively, while the [Glu53] mutant displayed very low activity toward M. lysodeikticus cells and no detectable activity toward glycol chitin.  相似文献   

10.
In this study, the selectivity and specificity of aldose reductase (AR) for glutathionyl aldehydes was examined. Relative to free aldehydes, AR was a more efficient catalyst for the reduction of glutathiolated aldehydes. Reduction of glutathionyl propanal [gammaGlu-Cys(propanal)-Gly] was more efficient than that of Gly-Cys(propanal)-Gly and gamma-aminobutyric acid-Cys(propanal)-Gly suggesting a possible interaction between alpha-carboxyl of the conjugate and AR. Two active site residues, Trp20 or Ser302, were identified by molecular modeling as potential sites of this interaction. Mutations containing tryptophan-to-phenylalanine (W20F) and serine-to-alanine (S302A) substitutions did not significantly affect reduction of free aldehydes but decreased the catalytic efficiency of AR for glutathiolated aldehydes. Combined mutations indicate that both Trp20 and Ser302 are required for efficient catalysis of the conjugates. The decrease in efficiency due to W20F mutation with glutathionyl propanal was not observed with gamma-aminobutyric-Cys(propanal)-Gly or Gly-Cys-(propanal)-Gly, indicating that Trp20 is involved in binding the alpha-carboxyl of the conjugate. The effect of the S302A mutation was less severe when gammaGlu-Cys(propanal)-Glu rather than glutathionyl propanal was used as the substrate, consistent with an interaction between Ser302 and Gly-3 of the conjugate. These observations suggest that glutathiolation facilitates aldehyde reduction by AR and enhances the range of aldehydes available to the enzyme. Because the N-terminal carboxylate is unique to glutathione, binding of the conjugate with the alpha-carboxyl facing the bottom of the alpha/beta-barrel may assist in the exclusion of unrelated peptides and proteins.  相似文献   

11.
Raman spectroscopy was used to determine structural features of the native toxin alpha from Naja nigricollis, which contains only one Trp and one Tyr, and of chemically modified toxins having chromophores added to these two conserved aromatic amino acids. The percentages of secondary structure were determined by using amide I polypeptidic vibration analysis and are in agreement with X-ray structure [Low et al. (1976) Proc. Natl. Acad Sci. U.S.A. 73, 2991-2994] as well as with the geometry of the disulfide bridges estimated by using the v(S-S) vibrations. In the native toxin alpha, the single invariant tyrosine 25 appears to be buried in the structure and involved in a strong hydrogen bond. We have chemically modified these two invariant aromatic side chains by addition of chromophores. The presence of a (nitrophenyl)sulfenyl (NPS) chromophore bound to the Trp does not perturb the secondary structure of the toxin as shown by the analysis of the polypeptidic amide I vibrations; however, the environment of this Trp and the geometry of a disulfide bridge seem to be modified. The secondary structure is not affected by the presence of the NPS chromophore; therefore, the decrease in binding affinity observed after modification of Trp-29 by the reagent NPS-Cl [Faure et al. (1983) Biochemistry 22, 2068-2076] is due to an alteration of the environment of this aromatic amino acid and/or a steric hindrance and not to an overall modification of the toxin structure. The binding assays of [nitrotyrosyl]toxin show that after nitration the affinity toward the monoclonal antibody M alpha 1 is unchanged and that the affinity toward the cholinergic receptor (AcChR) from Torpedo marmorata remains high. We concluded that the structure of toxin alpha after adding the NO2 chromophore to Tyr-25 is the same as it is in native toxin.  相似文献   

12.
Achromobacter protease I (API), a lysine-specific serine-protease of the trypsin family, has an aromatic-ring stacking Trp 169-His 210 in close proximity to the reactive site. In order to investigate the role of this novel aromatic stacking, several mutants of the two residues were constructed and their kinetic parameters were determined. Three His 210 mutants showed lower activity by one order of magnitude than the wild-type with a peptide substrate of Ala-Ala-Lys-MCA (4-methylcoumaryl-7-amide), but 30-170% activity towards Val-Leu-Lys-MCA, suggesting that His 210 plays a role in keeping high activity toward various substrates by maintaining the active form of the substrate-binding subsite. Kinetic results of eight Trp 169 variants showed a roughly linear relation between k(cat) or K(m) values and the surface area at residue 169. With increasing size of the side-chain, k(cat) values increased, while K(m) values decreased. A systematic kinetic analysis of the activities of Trp 169 mutants toward Lys-MCA, Ala-Lys-MCA, and Ala-Ala-Lys-MCA peptide substrates revealed that large side-chain, rather than aromaticity, plays an important role in retaining the high catalytic activity of API. Due to the presence of the aromatic stacking, API shows one order of magnitude higher activity than bovine trypsin.  相似文献   

13.
Versatile peroxidases (VP), a recently described family of ligninolytic peroxidases, show a hybrid molecular architecture combining different oxidation sites connected to the heme cofactor. High-resolution crystal structures as well as homology models of VP isoenzymes from the fungus Pleurotus eryngii revealed three possibilities for long-range electron transfer for the oxidation of high redox potential aromatic compounds. The possible pathways would start either at Trp164 or His232 of isoenzyme VPL, and at His82 or Trp170 of isoenzyme VPS1. These residues are exposed, and less than 11 A apart from the heme. With the purpose of investigating their functionality, two single mutations (W164S and H232F) and one double mutation (W164S/P76H) were introduced in VPL that: (i) removed the two pathways in this isoenzyme; and (ii) incorporated the absent putative pathway. Analysis of the variants showed that Trp164 is required for oxidation of two high redox potential model substrates (veratryl alcohol and Reactive Black 5), whereas the two other pathways (starting at His232 and His82) are not involved in long-range electron transfer (LRET). None of the mutations affected Mn2+ oxidation, which would take place at the opposite side of the enzyme. Substitution of Trp164 by His also resulted in an inactive variant, indicating that an indole side-chain is required for activity. It is proposed that substrate oxidation occurs via a protein-based radical. For the first time in a ligninolytic peroxidase such an intermediate species could be detected by low-temperature electron paramagnetic resonance of H2O2-activated VP, and was found to exist at Trp164 as a neutral radical. The H2O2-activated VP was self-reduced in the absence of reducing substrates. Trp164 is also involved in this reaction, which in the W164S variant was blocked at the level of compound II. When analyzing VP crystal structures close to atomic resolution, no hydroxylation of the Trp164 Cbeta atom was observed (even after addition of several equivalents of H2O2). This is in contrast to lignin peroxidase Trp171. Analysis of the crystal structures of both peroxidases showed differences in the environment of the protein radical-forming residue that could affect its reactivity. These variations would also explain differences found for the oxidation of some high redox potential aromatic substrates.  相似文献   

14.
Deficient activity of beta-hexosaminidase A (Hex A), resulting from mutations in the HEXA gene, typically causes Tay-Sachs disease. However, healthy individuals lacking Hex A activity against synthetic substrates (i.e., individuals who are pseudodeficient) have been described. Recently, an apparently benign C739-to-T (Arg247Trp) mutation was found among individuals with Hex A levels indistinguishable from those of carriers of Tay-Sachs disease. This allele, when in compound heterozygosity with a second "disease-causing" allele, results in Hex A pseudodeficiency. We examined the HEXA gene of a healthy 42-year-old who was Hex A deficient but did not have the C739-to-T mutation. The HEXA exons were PCR amplified, and the products were analyzed for mutations by using restriction-enzyme digestion or single-strand gel electrophoresis. A G805-to-A (Gly269Ser) mutation associated with adult-onset GM2 gangliosidosis was found on one chromosome. A new mutation, C745-to-T (Arg249Trp), was identified on the second chromosome. This mutation was detected in an additional 4/63 (6%) non-Jewish and 0/218 Ashkenazi Jewish enzyme-defined carriers. Although the Arg249Trp change may result in a late-onset form of GM2 gangliosidosis, any phenotype must be very mild. This new mutation and the benign C739-to-T mutation together account for approximately 38% of non-Jewish enzyme-defined carriers. Because carriers of the C739-to-T and C745-to-T mutations cannot be differentiated from carriers of disease-causing alleles by using the classical biochemical screening approaches, DNA-based analyses for these mutations should be offered for non-Jewish enzyme-defined heterozygotes, before definitive counseling is provided.  相似文献   

15.
Modulators are proposed to be novel ether aminophosphoglycerides that stabilize unoccupied and occupied glucocorticoid receptor steroid binding and inhibit glucocorticoid receptor complex activation. Two isoforms, modulator 1 and modulator 2, have been purified from rat liver cytosol [Bodine, P.V., & Litwack, G. (1990) J. Biol. Chem. 265, 9544-9554]. Since the mineralocorticoid receptor is relatively resistant to activation, modulator's effect on rat distal colon mineralocorticoid receptor function was examined. Warming of unoccupied receptor decreased residual specific [3H]aldosterone binding by 86 +/- 2%. Both modulator isoforms completely prevented this destabilization with Km's of 2 +/- 1 microM modulator 1 and 24 +/- 5 microM modulator 2. Warming of occupied mineralocorticoid receptors decreased [3H]aldosterone binding by 56 +/- 3%. Modulator only partially stabilized occupied receptor binding with Km's of 10 +/- 2 microM modulator 1 and 68 +/- 8 microM modulator 2. Modulator inhibited receptor activation with Km's of 3 +/- 1 microM modulator 1 and 33 +/- 10 microM modulator 2. Double-reciprocal analysis showed linear kinetics, and mixing modulator isoforms together had additive effects on unoccupied and occupied receptor steroid binding stabilization and activation inhibition. Colon cytosol contained a low molecular weight, heat-stable factor(s) which inhibited receptor activation and stabilized occupied receptor steroid binding. Molybdate completely stabilized unoccupied mineralocorticoid receptor steroid binding and inhibited activation with half-maximal effects at 3-4 mM but only stabilized occupied receptor binding by approximately 40%. These data indicate that (i) apparent physiologic concentrations of modulator stabilize mineralocorticoid receptor steroid binding and inhibit receptor activation, (ii) an aldosterone-responsive tissue contains a modulator-like activity, and (iii) molybdate mimics the effects of modulator.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The effects of hypoxia on metabolism of 5-hydroxytryptamine (5-HT or serotonin) and 3,4-dihydroxyphenylethylamine (DA or dopamine) were compared with those on open-field activity in male CD-1 mice. Chemical hypoxia was induced with NaNO2. Hypoxia did not alter striatal concentrations of DA, 5HT, Trp, Tyr, 5-hydroxyindoleacetic acid, or homovanillic acid. However, NaNO2 (75 mg/kg) reduced the rates of conversion of [3H]Tyr to [3H]DA (-41%) and [3H]Trp to [3H]5-HT (-39%). Hypoxia also reduced dihydroxyphenylacetic acid (DOPAC) levels (-27%) and DOPAC/DA ratios (-20%). Open-field behavior, as measured in an automated activity monitor, decreased in a dose-dependent fashion with 75-150 mg/kg of NaNO2 (-35 to -90%). Comparison with previous studies suggests that the syntheses of dopamine, serotonin, and the amino acids are equally vulnerable to hypoxic insults but may be less sensitive than the synthesis of acetylcholine.  相似文献   

17.
The fate of the acetyl-CoA units released during peroxisomal fatty acid oxidation was studied in isolated hepatocytes from normal and peroxisome-proliferated rats. Ketogenesis and hydrogen peroxide generation were employed as indicators of mitochondrial and peroxisomal fatty acid oxidation, respectively. Butyric and hexanoic acids were employed as mitochondrial substrates, 1, omega-dicarboxylic acids as predominantly peroxisomal substrates, and lauric acid as a substrate for both mitochondria and peroxisomes. Ketogenesis from dicarboxylic acids was either absent or very low in normal and peroxisome-proliferated hepatocytes, but free acetate release was detected at rates that could account for all the acetyl-CoA produced in peroxisomes by dicarboxylic and also by monocarboxylic acids. Mitochondrial fatty acid oxidation also led to free acetate generation but at low rates relative to ketogenesis. The origin of the acetate released was confirmed employing [1-14C]dodecanedioic acid. Thus, the activity of peroxisomes might contribute significantly to the free acetate generation known to occur during fatty acid oxidation in rats and possibly also in humans.  相似文献   

18.
Xenobiotic metabolism and mutation in a human lymphoblastoid cell line   总被引:2,自引:0,他引:2  
Aryl hydrocarbon hydroxylase-1 (AHH-1) cells are a human lymphoblastoid cell line competent in some aspects of xenobiotic metabolism. This cell line contains stable mixed function oxidase activity which is inducible by polycyclic aromatic hydrocarbons (PAHs) but not by phenobarbital or Arochlor 1254. Two substrates for the cellular mixed function oxidase activity, benzo[a]pyrene (B[a]P) and 7-ethoxyresorufin, have been examined. The basal and induced activities have different kinetic parameters toward these two substrates. In contrast, basal and induced activities had similar sensitivities to two cytochrome P-450 suicide substrates. B[a]P metabolism and mutagenicity were studied in this cell line. AHH-1 cells were found to produce predominantly B[a]P phenols and quinones. The major phenol metabolite cochromatographed with authentic 9-hydroxy B[a]P. AHH-1 cells were capable of forming glucuronic acid conjugates of B[a]P phenols; the major product after hydrolysis cochromatographed with 3-hydroxy B[a]P standard. AHH-1 cells did not contain detectable epoxide hydrolase activity using B[a]P-4,5-oxide as substrate. This observation is consistent with the absence of trans-dihydrodiol B[a]P metabolites in the metabolic profile. B[a]P-induced mutagenicity at the hypoxanthine guanine phosphoribosyl transferase (hgprt) locus in AHH-1 cells was found to be linearly related to phenol production during treatment and inhibited by alpha-naphthoflavone (ANF).  相似文献   

19.
20.
In steroid hydroxylation system in adrenal cortex mitochondria, NADPH-adrenodoxin reductase (AR) and adrenodoxin (Adx) form a short electron-transport chain that transfers electrons from NADPH to cytochromes P-450 through FAD in AR and [2Fe-2S] cluster in Adx. The formation of [AR/Adx] complex is essential for the electron transfer mechanism in which previous studies suggested that AR tryptophan (Trp) residue(s) might be implicated. In this study, we modified AR Trps by N-bromosuccinimide (NBS) and studied AR binding to Adx by a resonant mirror biosensor. Chemical modification of tryptophans caused inhibition of electron transport. The modified protein (AR*) retained the native secondary structure but showed a lower affinity towards Adx with respect to AR. Activity measurements and fluorescence data indicated that one Trp residue of AR may be involved in the electron transferring activity of the protein. Computational analysis of AR and [AR/Adx] complex structures suggested that Trp193 and Trp420 are the residues with the highest probability to undergo NBS-modification. In particular, the modification of Trp420 hampers the correct reorientation of AR* molecule necessary to form the native [AR/Adx] complex that is catalytically essential for electron transfer from FAD in AR to [2Fe-2S] cluster in Adx. The data support an incorrect assembly of [AR*/Adx] complex as the cause of electron transport inhibition.  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